Skript Algtopo
Skript Algtopo
GEOMETRIE IN POTSDAM
Algebraic Topology
Lecture Notes, Summer Term 2022
Preface 1
2. Homotopy Theory 15
2.1. Homotopic maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2. The fundamental group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.3. The fundamental group of the circle . . . . . . . . . . . . . . . . . . . . . . . 27
2.4. The Seifert-van Kampen theorem . . . . . . . . . . . . . . . . . . . . . . . . . 40
2.5. The fundamental group of surfaces . . . . . . . . . . . . . . . . . . . . . . . . 51
2.6. Higher homotopy groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
2.7. Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
3. Homology Theory 79
3.1. Singular homology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
3.2. Relative homology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
3.3. The Eilenberg-Steenrod axioms and applications . . . . . . . . . . . . . . . . 88
3.4. The degree of a continuous map . . . . . . . . . . . . . . . . . . . . . . . . . 96
3.5. Homological algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
3.6. Proof of the homotopy axiom . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
3.7. Proof of the excision axiom . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
3.8. The Mayer-Vietoris sequence . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
3.9. Generalized Jordan curve theorem . . . . . . . . . . . . . . . . . . . . . . . . 126
3.10. CW-complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
3.11. Homology of CW-complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
3.12. Betti numbers and the Euler number . . . . . . . . . . . . . . . . . . . . . . . 141
3.13. Incidence numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
3.14. Homotopy versus homology . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
3.15. Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
i
Contents
A. Appendix 157
A.1. Free module generated by a set . . . . . . . . . . . . . . . . . . . . . . . . . . 157
Bibliography 163
Index 165
ii
Preface
These are the lecture notes of an introductory course on algebraic topology which I taught at
the University of Potsdam during the summer term 2022. 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 back in 2010. I am grateful to
Volker Branding who wrote a first draft of those lecture notes and created most of the figures
and to Ramona Ziese who improved those notes considerably. Moreover, I would like to thank
the participants of the 2022 course, especially for the valuable feedback that they provided.
Christian Bär
1
1. Set Theoretic Topology
Here are four versions of a typical question that one asks in mathematics. Fix integers 1 ≤ 𝑛 < 𝑚.
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
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→∞
𝜑1 𝜑2 𝜑3
𝜑4 𝜑5 𝜑6
4
1.1. Typical problems in topology
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.
1.) Given two spaces, are they homeomorphic? To show that they are, construct a homeo-
morphism. To show that they are not, find topological invariants, which are different for given
spaces.
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:
𝑆2 (sphere), genus 0
𝑇 2 (torus), genus 1
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 .
2 Images from https://pixabay.com
5
1. Set Theoretic Topology
Example 1.6 (Brouwer fixed point theorem). Any continuous map 𝑓 : 𝐷 𝑛 → 𝐷 𝑛 with
𝐷 𝑛 = {𝑥 ∈ R𝑛 |k𝑥 k ≤ 1} has a fixed point, i.e., there exists an 𝑥 ∈ 𝐷 𝑛 such that
𝑓 (𝑥) = 𝑥.
By the intermediate value theorem we can find an 𝑥 with 𝑔(𝑥) = 0 which is equivalent to 𝑓 (𝑥) = 𝑥.
(i) ∅, R𝑛 ∈ TR𝑛 ;
Ð
(ii) 𝑈 𝑗 ∈ TR𝑛 , 𝑗 ∈ 𝐽 ⇒ 𝑗∈𝐽 𝑈 𝑗 ∈ 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
6
1.2. Some basic definitions
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 ;
Ð
(ii) 𝑈 𝑗 ∈ T𝑋 , 𝑗 ∈ 𝐽 ⇒ 𝑗∈𝐽 𝑈 𝑗 ∈ T𝑋 ;
(iii) 𝑈1 , 𝑈2 ∈ T𝑋 ⇒ 𝑈1 ∩ 𝑈2 ∈ T𝑋 .
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,
7
1. Set Theoretic Topology
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 = id𝑋 : (𝑋, Tdiscrete ) → (𝑋, Tcoarse ).
1.3. Compactness
Examples 1.22. 1.) Finite sets are always compact. Namely, et 𝑌 = {𝑦 1 , . . . , 𝑦 𝑛 } and let {𝑈𝑖 }𝑖∈ 𝐼
be an open cover of 𝑌 . Then choose 𝑖 𝑗 such that 𝑦 𝑖 ∈ 𝑈𝑖 𝑗 . Then {𝑈𝑖1 , . . . , 𝑈𝑖𝑛 } 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 < ∞.
8
1.4. Hausdorff spaces
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 𝑖 1 , . . . , 𝑖 𝑛 such that 𝑋 ⊂ 𝑈 ∪ 𝑈𝑖1 ∪ · · · ∪ 𝑈𝑖𝑛 and we
conclude that 𝑌 ⊂ 𝑈𝑖1 ∪ · · · ∪ 𝑈𝑖𝑛 .
Proof. Let {𝑈𝑖 }𝑖∈ 𝐼 be an open cover of 𝑓 (𝐾), i.e. 𝑓 (𝐾) ⊂ ∪𝑖∈ 𝐼 𝑈𝑖 . Then we have
9
1. Set Theoretic Topology
b b
𝑈1 𝑈2
Figure 5. Hausdorff property
Examples 1.27. 1.) Spaces with discrete topology are Hausdorff spaces, simply put 𝑈𝑖 = {𝑥𝑖 }.
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 𝑥1 ≠ 𝑥2 put
𝑟 := 𝑑 (𝑥1 , 𝑥2 ) > 0. Then the open balls 𝑈𝑖 = 𝐵(𝑥𝑖 , 𝑟/2) separate 𝑥1 and 𝑥2 .
4.) Let 𝑋 be Hausdorff and let 𝑌 ⊂ 𝑋 any subset. Then 𝑌 with its induced topology is again
Hausdorff.
Proof. Let 𝑝 ∈ 𝑋 \ 𝑌 . Then for every 𝑦 ∈ 𝑌 the Hausdorff property tells us that there exist open
subsets 𝑈 𝑦, 𝑝 , 𝑉𝑦, 𝑝 ⊂ 𝑋 such that 𝑦 ∈ 𝑈 𝑦, 𝑝 , 𝑝 ∈ 𝑉𝑦, 𝑝 and 𝑈 𝑦, 𝑝 ∩ 𝑉𝑦, 𝑝 = ∅. Since 𝑌 is compact
there exist 𝑦 1 , . . . , 𝑦 𝑛 ∈ 𝑌 such that 𝑌 ⊂ 𝑈 𝑦1 , 𝑝 ∪ · · · ∪ 𝑈 𝑦𝑛 , 𝑝 . Now put 𝑉 𝑝 := ∩𝑛𝑗=1𝑉𝑦 𝑗 , 𝑝 . Since
𝑉 𝑝 is a finite intersection of open subsets it is open itself. Moreover, 𝑝 ∈ 𝑉 𝑝 . Now
𝑌 ∩ 𝑉 𝑝 ⊂ (𝑈 𝑦1 , 𝑝 ∪ · · · ∪ 𝑈 𝑦𝑛 , 𝑝 ) ∩ 𝑉 𝑝 ⊂ (𝑈 𝑦1 , 𝑝 ∩ 𝑉𝑦1 , 𝑝 ) ∪ · · · ∪ (𝑈 𝑦𝑛 , 𝑝 ∩ 𝑉𝑦𝑛 , 𝑝 ) = ∅,
hence 𝑉 𝑝 ⊂ 𝑋 \ 𝑌 . Therefore
𝑋 \ 𝑌 = ∪ 𝑝∈𝑋\𝑌 𝑉 𝑝 ⊂ 𝑋 is open.
Thus 𝑌 ⊂ 𝑋 is closed.
10
1.5. Quotient spaces
Let 𝜋 : 𝑋 → 𝑋/∼ , 𝑥 ↦→ [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=
⑤
𝜋 ⑤⑤⑤
⑤ ¯
⑤ 𝑓
𝑋/∼
Example 1.30
Let 𝑋 = [0, 1] and let the equivalence relation be given by 𝑥 ∼ 𝑥 ∀𝑥 ∈ 𝑋 and 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.
11
1. Set Theoretic Topology
𝑥1 ∼ 𝑥2 ⇔ 𝑥1 = 𝑥2 or 𝑥1 , 𝑥2 ∈ 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 1.32. R/ (0,∞) is not a Hausdorff space because the points [0] and [1] cannot be
separated.
𝑋 := 𝑋1 × · · · × 𝑋𝑛
Definition 1.33. A subset 𝑈 ⊂ 𝑋 is called open (for the product topology) iff for all
𝑝 = ( 𝑝 1 , . . . , 𝑝 𝑛 ) ∈ 𝑈 there exist open subsets 𝑈𝑖 ⊂ 𝑋𝑖 with 𝑝 𝑖 ∈ 𝑈𝑖 and 𝑈1 × · · · × 𝑈𝑛 ⊂ 𝑈.
Examples 1.34. 1.) If all 𝑋𝑖 carry the discrete topology then 𝑋 carries the discrete topology.
Namely, the sets {( 𝑝 1 , . . . , 𝑝 𝑛 )} are open, hence all subsets of the product space are open.
2.) If all 𝑋𝑖 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.
1.) The projection maps 𝜋𝑖 : 𝑋 → 𝑋𝑖 , 𝜋𝑖 (𝑥) = 𝑥𝑖 , are continuous because for any open subset
𝑈𝑖 ⊂ 𝑋𝑖
𝜋𝑖−1 (𝑈𝑖 ) = 𝑋1 × · · · × 𝑋𝑖−1 × 𝑈𝑖 × 𝑋𝑖+1 × · · · × 𝑋𝑛
is open in 𝑋.
2.) Fix 𝑥𝑖 ∈ 𝑋𝑖 where 𝑖 ∈ {1, . . . , 𝑖 0 − 1, 𝑖 0 + 1, . . . , 𝑛}. Then the map 𝜄 : 𝑋𝑖0 → 𝑋 is continuous
where 𝜄(𝜉) = (𝑥1 , . . . , 𝑥𝑖0 −1 , 𝜉, 𝑥𝑖0 +1 , . . . , 𝑥 𝑛 ).
12
1.7. Exercises
3.) Universal property: for all topological spaces 𝑌 and for all maps
𝑓 = ( 𝑓 1 , . . . , 𝑓 𝑛 ) : 𝑌 → 𝑋 = 𝑋1 × · · · × 𝑋 𝑛
1.7. Exercises
1.1. Determine all topologies on the set {1, 2, 3} and investigate which ones 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.
a) are not homeomorphic when equipped with the standard topologies induced by R;
1.5. Let 𝐷 𝑛 be as in Exercise 1.2. Let 𝑥, 𝑦 ∈ 𝐷˚ 𝑛 , i.e., k𝑥 k, k𝑦k < 1. Construct a homeomorphism
𝜑 : 𝐷 𝑛 → 𝐷 𝑛 with 𝜑(𝑥) = 𝑦 and 𝜑(𝑧) = 𝑧 for all 𝑧 ∈ 𝜕𝐷 𝑛 , i.e., k𝑧k = 1.
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.
a) 𝐷 𝑛 /𝜕𝐷 𝑛 ≈ 𝑆 𝑛 ;
b) 𝑊 𝑛 /𝜕𝑊 𝑛 ≈ 𝑆 𝑛 .
13
1. Set Theoretic Topology
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 .
b) Σ𝐷 𝑛 ≈ 𝐷 𝑛+1 .
c) Σ𝑆 𝑛 ≈ 𝑆 𝑛+1 .
14
2. Homotopy Theory
Definition 2.1. Let 𝑋 and 𝑌 be topological spaces. Let 𝐴 ⊂ 𝑋 be a subset. Two continuous
maps 𝑓0 , 𝑓1 : 𝑋 → 𝑌 are called homotopic relative to 𝐴 iff there exists a continuous map
𝐹 : 𝑋 × [0, 1] → 𝑌 such that
Remark 2.3. If 𝐴 = ∅, then we say “ 𝑓0 and 𝑓1 are homotopic” instead of “ 𝑓0 and 𝑓1 are homotopic
relative to ∅”. Similarly, we write “ 𝑓0 ≃ 𝑓1 ” instead of “ 𝑓0 ≃ ∅ 𝑓1 ”.
2.) Let 𝑓0 : R𝑛 → 𝑌 be continuous. Put 𝐹 (𝑥, 𝑡) := 𝑓0 ((1 − 𝑡)𝑥), then 𝑓0 ≃ const map.
3.) Lef 𝑓 = Exp : R → 𝑆1 ∈ C, Exp(𝑥) = 𝑒2 𝜋𝑖𝑥 . 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}.
15
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), 0 ≤ 𝑡 ≤ 1/2
𝐻 (𝑥, 𝑡) :=
𝐺 (𝑥, 2𝑡 − 1), 1/2 ≤ 𝑡 ≤ 1
Remark 2.8. This defines an equivalence relation on the class of topological spaces.
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.
16
2.1. Homotopic maps
Definition 2.12. Let 𝐴 ⊂ 𝑋 and let 𝜄 : 𝐴 → 𝑋 be the inclusion map. Then 𝐴 is called
Examples 2.13. 1.) Let 𝑋 any topological space, let 𝐴 = {𝑥0 } ⊂ 𝑋 consist of just one point.
Then 𝑟 : 𝑋 → 𝐴, 𝑟 (𝑥) = 𝑥0 , 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) = 𝑡𝑥 + (1 − 𝑡)
k𝑥 k
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.
17
2. Homotopy Theory
𝐴
Example 2.14 1
We consider the comb space given by
n o
𝑋 = (𝑥, 𝑦) ∈ R2 | 0 ≤ 𝑦 ≤ 1 and (𝑥 = 0 or 𝑥 = 1
for some 𝑛 ∈ N) ......
𝑛
∪ (𝑥, 𝑦) ∈ R2 | 𝑦 = 0 and 0 ≤ 𝑥 ≤ 1
Therefore id𝑋 ≃ 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
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
k𝐺 (𝑥, 𝑦, 𝑡) − 𝐺 (𝑥 ′ , 𝑦 ′ , 𝑡 ′ ) k < 1
2
18
2.1. Homotopic maps
𝐴
1 1
Now choose 𝑛 so large that 𝑛 < 𝛿 and consider
(𝑥, 𝑦) = 1
𝑛, 1 , (𝑥 ′ , 𝑦 ′ ) = (0, 1), 𝑡 = 𝑡′.
Then
1
𝐺 𝑛 , 1, 𝑡 − 𝐺 (0, 1, 𝑡) < 21 , ∀𝑡 ∈ [0, 1] .
On the other hand the mapping 𝑡 ↦→ 𝐺 ( 𝑛1 , 1, 𝑡) is a continuous path in 𝑋 from ( 𝑛1 , 1) to (0, 1) and
must take values in 𝑍 for some 𝑡.
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 = {𝑥0 } 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
“ ⇒ ”.
19
2. Homotopy Theory
𝑥1
𝑋
𝜂 𝑥0 𝑥2
𝜔
0 1 0 1
𝑡 → 2𝑡 𝑡 → 2𝑡 − 1
0 1
Figure 9. Concatenation of paths
20
2.2. The fundamental group
Definition 2.17. For 𝜔 ∈ Ω(𝑋; 𝑥0 ) denote by [𝜔] the homotopy class of 𝜔 relative to {0, 1}.
Then
𝜋1 (𝑋; 𝑥0 ) := {[𝜔] | 𝜔 ∈ Ω(𝑋; 𝑥0 )}
is called the fundamental group of (𝑋, 𝑥0 ).
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
gets mapped to 𝑥0 .
𝜔 𝜂
Figure 10. Concatenations of homotopic paths are homotopic.
In formulas, if we denote the homotopy between 𝜔 and 𝜔′ by 𝐹 and the one between 𝜂 and 𝜂 ′ by
𝐺, then the homotopy 𝐻 : [0, 1] × [0, 1] → 𝑋 between 𝜔 ★ 𝜂 and 𝜔′ ★ 𝜂 ′ is given by
(
𝐹 (2𝑡, 𝑠), 0 ≤ 𝑡 ≤ 12 ,
𝐻 (𝑡, 𝑠) =
𝐺 (2𝑡 − 1, 𝑠), 21 ≤ 𝑡 ≤ 1.
(ii) The first diagram in Figure ?? shows 𝜀 𝑥0 ★𝜔 ≃ {0,1} 𝜔, the second proves that 𝜔 ≃ {0,1} 𝜔★𝜀 𝑥0 .
21
2. Homotopy Theory
𝜔 𝜔
𝜀 𝑥0 𝜔 𝜔 𝜀 𝑥0
Figure 11. Concatenation with constant path is homotopic to original path
𝜔(1 − 𝑠)
𝑠
𝜔 𝜔 −1
Figure 12. Inverse modulo homotopy
For any 𝑠 ∈ [0, 1] the corresponding ”blue“ line segment gets mapped to 𝜔(1 − 𝑠). In formulas,
the homotopy 𝐻 : [0, 1] × [0, 1] → 𝑋 between 𝜔 ★ 𝜔 −1 and 𝜀 𝑥0 is given by
𝜔(2𝑡), 0 ≤ 𝑡 ≤ 1−𝑠
2 ,
1−𝑠 1+𝑠
𝐻 (𝑡, 𝑠) = 𝜔(1 − 𝑠), 2 ≤ 𝑡 ≤ 2 ,
𝜔(2 − 2𝑡),
1+𝑠
2 ≤ 𝑡 ≤ 1.
22
2.2. The fundamental group
𝜔 𝜂 𝜁
𝜔 𝜂 𝜁
Figure 13. Associativity
Hence 𝜋1 (𝑋; 𝑥0 ) 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 𝑥0 ∈ 𝑋.
We put 𝑓 (𝑥0 ) =: 𝑦 0 ∈ 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; 𝑥0 ) → 𝜋1 (𝑌 ; 𝑦 0 ), 𝑓# ([𝜔]) = [ 𝑓 ◦ 𝜔].
23
2. Homotopy Theory
𝑓# : 𝜋1 (𝑋, 𝑥0 ) → 𝜋1 (𝑌 , 𝑦 0 )
is a group isomorphism.
( 𝑓 −1 )# ◦ ( 𝑓# ) = ( 𝑓 −1 ◦ 𝑓 )# = (id𝑋 )# = id 𝜋1 (𝑋;𝑥0 )
Now we want to deal with the question to what extent 𝜋1 (𝑋; 𝑥0 ) depends on the choice of 𝑥0 .
𝑋
To study this question let 𝑥0 , 𝑥1 ∈ 𝑋 and assume 𝜔
there exists a path 𝛾 ∈ Ω(𝑋; 𝑥0 , 𝑥1 ). If such a
𝑥1
path does not exist 𝜋1 (𝑋; 𝑥0 ) and 𝜋1 (𝑋; 𝑥1 ) are not
related. 𝛾
𝑥0
Figure 14. Dependence on base point
Look at the map Φ𝛾 : Ω(𝑋; 𝑥1 ) → Ω(𝑋; 𝑥0 ) where 𝜔 ↦→ (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
is well defined.
24
2.2. The fundamental group
5.) For any [𝜔] ∈ 𝜋1 (𝑋; 𝑥1 ) we have Φ̂𝛾′ ([𝜔]) = 𝜅 · Φ̂𝛾 ([𝜔]) · 𝜅 −1 where 𝜅 = [𝛾 ′ ★ 𝛾 −1 ] ∈
𝜋1 (𝑋; 𝑥0 ).
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
Φ̂𝛾 [𝜔] · [𝜂] = Φ̂𝛾 ([𝜔 ★ 𝜂])
= (𝛾 ★ (𝜔 ★ 𝜂)) ★ 𝛾 −1
= 𝛾 ★ ((𝜔 ★ (𝛾 −1 ★ 𝛾)) ★ 𝜂) ★ 𝛾 −1
= (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 = Φ̂ 𝜀𝑥0 = id 𝜋1 (𝑋;𝑥0 ) .
Proposition 2.22. Let 𝑋, 𝑌 be topological spaces and 𝑥0 ∈ 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 (𝑥0 ), 𝑔(𝑥0 )) by
25
2. Homotopy Theory
𝑌
𝑔◦𝜔
𝜂
𝑠 𝐹
𝑓 ◦𝜔
𝑡
𝜋1 (𝑋; 𝑥0 )P
PPP
𝑔#
PPP𝑓#
PPP
PP(
𝜋1 (𝑌 ; 𝑔(𝑥0 )) / 𝜋1 (𝑌 ; 𝑓 (𝑥 0 ))
Φ̂ 𝜂
Proof. Let [𝜔] ∈ 𝜋1 (𝑋; 𝑥0 ) and define 𝐹 : [0, 1] × [0, 1] → 𝑌 by 𝐹 (𝑡, 𝑠) := 𝐻 (𝜔(𝑡), 𝑠).
| |
1 2
3 3
Figure 16. The deformation
Now we can improve Corollary 2.20 and show that homotopy equivalent spaces have isomorphic
fundamental groups.
𝑓# : 𝜋1 (𝑋; 𝑥0 ) → 𝜋1 (𝑌 ; 𝑓 (𝑥0 ))
26
2.3. The fundamental group of the circle
𝑔# ◦ 𝑓# = (𝑔 ◦ 𝑓 )# = Φ̂ 𝜂 ◦ (id𝑋 )# = Φ̂ 𝜂 ◦ id 𝜋1 (𝑋;𝑥0 ) = Φ̂ 𝜂 .
Example 2.25. If 𝑋 is contractible then the one-point set 𝐴 = {𝑥0 } ⊂ 𝑋 is a deformation retract.
Hence 𝜋1 (𝑋; 𝑥0 ) ≃ 𝜋1 ( 𝐴; 𝑥0 ) = {[𝜀 𝑥0 ]} = {1}. Thus contactible spaces have trivial fundamental
group.
Lemma 2.26. Let 𝑡0 ∈ R and 𝑧0 = Exp(𝑡0 ) ∈ 𝑆1 . Then for all 𝑓 ∈ 𝐶 (𝑆1 , 𝑆1 ) with 𝑓 (1) = 𝑧0
there exists a unique 𝑓ˆ ∈ 𝐶 (R, R) with 𝑓ˆ(0) = 𝑡0 such that the following diagram commutes:
𝑓ˆ
R /R
Exp Exp
𝑆1 / 𝑆1
𝑓
27
2. Homotopy Theory
Moreover, the restriction of Exp to each of these intervals is a homeomorphism onto its image,
1 ≈
Exp [𝑡1 +𝑘,𝑡1 +𝑘+ 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.
𝑓ˆ := (Exp | 𝐼0 ) −1 ◦ 𝑓 ◦ Exp,
1
where 𝐼0 ⊂ R is the compact interval of length 2 with Exp(𝐼0 ) = 𝐶0 and 𝑡0 ∈ 𝐼0 . This insures
that
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 12 with Exp(𝐼1 ) = 𝐶1 and 𝑡1 ∈ 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.
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 .
28
2.3. The fundamental group of the circle
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.
3.) The map 𝑡 ↦→ 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.
Hence 𝑓ˆ + 𝑔ˆ is a lift of 𝑓 𝑔 and we get the following formula for the degree
deg( 𝑓 𝑔) = 𝑓ˆ(1) + 𝑔(1)
ˆ − 𝑓ˆ(0) + 𝑔(0)
ˆ = deg( 𝑓 ) + deg(𝑔) .
29
2. Homotopy Theory
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
Thus 𝑓 is onto.
Example 2.31. For the map 𝑓𝑛 : 𝑆1 → 𝑆1 with 𝑓𝑛 (𝑧) = 𝑧 𝑛 a lift is given by 𝑓ˆ(𝑡) = 𝑛𝑡 so that its
degree is deg( 𝑓𝑛 ) = 𝑓ˆ𝑛 (1) − 𝑓ˆ𝑛 (0) = 𝑛.
is continuous and not surjective because −1 is not contained in the image. Hence, by Remark 2.30,
𝐹 (·,𝑠)
deg 𝐹 (·,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
where 0 = 𝑠0 < 𝑠1 < · · · < 𝑠𝑟 = 1 is a partition of the unit interval [0, 1] satisfying
|𝑠𝑖+1 − 𝑠𝑖 | < 𝛿.
30
2.3. The fundamental group of the circle
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.
Since dividing by 𝑎 𝑛 does not change the roots, we assume without loss of generality that 𝑎 𝑛 = 1.
Now assume that 𝑝 has no roots. Then the map 𝑓 : C → 𝑆1 with 𝑓 (𝑧) = | 𝑝𝑝 (𝑧)
(𝑧) | is a well-defined
continuous map. To compute deg( 𝑓 𝑆 1 ) consider
The map 𝐹 ∈ 𝐶 (𝑆1 × [0, 1], 𝑆1 ) is a homotopy from 𝑓𝑛 (𝑧) = 𝑧 𝑛 to 𝑓 𝑆 1 . Computing its degree
we find deg( 𝑓 𝑆 1 ) = deg( 𝑓𝑛 ) = 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.
Moreover,
− 𝑓 Exp(𝑡) = − Exp 𝑓ˆ(𝑡) = Exp 21 Exp 𝑓ˆ(𝑡) = Exp 𝑓ˆ(𝑡) + 21 .
31
2. Homotopy Theory
From 𝑓 (− Exp(𝑡)) = − 𝑓 (Exp(𝑡)) we conclude Exp 𝑓ˆ 𝑡 + 12 = Exp 𝑓ˆ(𝑡) + 12 and hence
1 1
𝑓ˆ 𝑡 + 2 − 𝑓ˆ(𝑡) + 2 =: 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 + 1 ) − 𝑓ˆ(𝑡) = 𝑘 + 1 for all 𝑡 ∈ R. Now we can compute
2 2
deg( 𝑓 ) = 𝑓ˆ(1) − 𝑓ˆ(0) = 𝑓ˆ(1) − 𝑓ˆ 12 + 𝑓ˆ 21 − 𝑓ˆ(0)
= (𝑘 + 21 ) + (𝑘 + 12 ) = 2𝑘 + 1
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
p
is defined and continuous. Moreover, 𝑔 satisfies 𝑔(−𝑥) = −𝑔(𝑥) for all 𝑥 ∈ 𝑆 2 . Now define a
map 𝐺 : 𝐷 2 → 𝑆1 by 𝐺 (𝑦) = 𝑔(𝑦, 1 − k𝑦k 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 𝑥0 ∈ 𝑆2 with 𝑓 (−𝑥0 ) = 𝑓 (𝑥0 ).
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 𝑥0 ∈ 𝑆2 with
0 = 𝑔(𝑥0 ) = 𝑓 (𝑥0 ) − 𝑓 (−𝑥0 ), 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 .
32
2.3. The fundamental group of the circle
𝐻 𝑥,𝑡 ′
𝐻 𝑥,𝑡
33
2. Homotopy Theory
vol( 𝐴)
𝛼𝑥
𝑡 𝑥− 𝑡 +𝑥
Figure 20. Monotonicity of the volume function
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
Thus the Borsuk-Ulam Theorem 2.36 applies and there exists a point 𝑥0 ∈ 𝑆2 with
Hence 𝛼(𝑥0 ) = 𝛽(𝑥0 ) = 𝛾(𝑥0 ). 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).
34
2.3. The fundamental group of the circle
𝑓𝜔
Exp 𝜔
!
a) For 𝜔 ∈ Ω(𝑆1 ; 1) and 𝐼 = [0, 1] consider the 𝑆1 a❈o 𝐼 / 𝑆1
❈❈ ④=
following 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
deg : Ω(𝑆1 ; 1) → Z.
c) This map is surjective because for 𝑛 ∈ Z we can consider the map 𝜔(𝑡) = Exp(𝑛𝑡). The
commutative diagram
𝑧↦→ 𝑧 𝑛 = 𝑓 𝑛 (𝑧)
𝑆1 f◆◆ / 𝑆1
O
◆◆◆
◆◆◆ 𝑡↦→ 𝜔 (𝑡 )=Exp (𝑛𝑡 )=Exp (𝑡 ) 𝑛
◆
𝑡↦→Exp(𝑡 ) ◆◆◆
◆
𝐼
shows 𝑓 𝜔 (𝑧) = 𝑧 𝑛 = 𝑓𝑛 (𝑧) and we get deg(𝜔) = deg( 𝑓𝑛 ) = 𝑛.
35
2. Homotopy Theory
Let 𝑓ˆ𝜔 , 𝑓ˆ𝜔 ′ be lifts of 𝑓 𝜔 , 𝑓 𝜔 ′ with 𝑓ˆ𝜔 ′ (0) = 𝑓ˆ𝜔 (1). Then we have
(
Exp( 𝑓ˆ𝜔 (2𝑡)), 0 ≤ 𝑡 ≤ 1/2
𝑓 𝜔★𝜔 ′ (Exp(𝑡)) =
Exp( 𝑓ˆ𝜔 ′ (2𝑡 − 1)), 1/2 ≤ 𝑡 ≤ 1
( !
𝑓ˆ𝜔 (2𝑡), 0 ≤ 𝑡 ≤ 1/2
= Exp
𝑓ˆ𝜔 ′ (2𝑡 − 1), 1/2 ≤ 𝑡 ≤ 1
| {z }
=: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′ .
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.
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)
36
2.3. The fundamental group of the circle
𝑆1 = 𝜕𝐷 2
Example 2.44 𝐷2
We show that 𝑆1 = 𝜕𝐷 2 is not a retract of 𝐷 2 . Suppose the
map 𝑟 : 𝐷 2 → 𝑆1 is a retraction and denote the inclusion by
𝜄 : 𝑆1 → 𝐷 2 .
Figure 21. The disk and its
boundary
Then we would have 𝑟 ◦ 𝜄 = id𝑆 1 , hence 𝑟 # ◦ 𝜄# = (𝑟 ◦ 𝜄)# = (id𝑆 1 )# = id 𝜋1 (𝑆 1 ;1) . We then get a
contradiction because of the following diagram
𝜄# 𝑟#
Z 𝜋1 (𝑆1 ; 1) / 𝜋1 (𝐷 2 ; 1) {1} / 𝜋1 (𝑆 1 ; 1) Z
:
id 𝜋 1
1 (𝑆 ;1)
As a corollary we get a proof of Brouwer’s fixed point theorem in dimension two. See page 93
for the theorem in general dimensions.
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
• 𝑟 (𝑥)
𝑓 (𝑥)
1 + 21 sin(𝑥)𝑦 − 2𝑥 = 0,
1 𝑥2
2 cos(𝑥)𝑦 + 2 − 2𝑦 = 0, (2.1)
37
2. Homotopy Theory
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
1 sin(𝑥)𝑦 cos(𝑥)𝑦 𝑥 2
k 𝑓 (𝑥, 𝑦) k 2 = + + +
2 4 4 4
1 sin(𝑥)𝑦 sin(𝑥) 𝑦 2 2 cos(𝑥) 2 𝑦 2 cos(𝑥)𝑦𝑥 2 𝑥 4
= + + + + +
4 4 16 16 8 16
1 sin(𝑥)𝑦 𝑦 2 cos(𝑥)𝑦𝑥 2 𝑥 4
= + + + +
4 4 16 8 16
1 |𝑦| 𝑦 2 |𝑦|𝑥 2 𝑥 4
≤ + + + +
4 4 16 8 16
1 1 1 1 1
≤ + + + +
4 4 16 8 16
3
=
4
≤ 1.
Example 2.47. Consider 𝑓𝑛 ∈ 𝐶 (𝑆1 , 𝑆1 ) with 𝑓𝑛 (𝑧) = 𝑧 𝑛 . We check that the diagram
deg
𝜋1 (𝑆1 ; 1)
/Z
( 𝑓 𝑛 )# 𝑛·
deg
𝜋1 (𝑆1 ; 1) /Z
Remark 2.48. From Exercise 2.5 we know that for 𝑋1 , 𝑋2 and 𝑥1 ∈ 𝑋1 , 𝑥2 ∈ 𝑋2 we have
38
2.3. The fundamental group of the circle
1.) We call 𝑋 connected iff 𝑋 and ∅ are the only subsets of 𝑋 which are both open and closed.
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 < 𝑡1 , the other case being analogous. We put 𝑇 := sup(𝐼 ∩ [0, 𝑡1 )). Then
0 ≤ 𝑡 0 ≤ 𝑇 ≤ 𝑡 1 ≤ 1. Since 𝐼 is closed 𝑇 ∈ 𝐼.
If 𝑇 = 1 then 𝑇 = 𝑡 1 which contradicts 𝑡1 ∉ 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.
Again, the converse implication does not hold in general. Consider for example the space
Remark 2.53. Let 𝑋 be path-connected. Then the following are equivalent (see Exercise 2.2):
39
2. Homotopy Theory
𝑔 · 𝐻 = 𝐻 · 𝑔 for all 𝑔 ∈ 𝐺 .
The condition on a subgroup of being normal can be reformulated in various ways. It is equivalent
to any of the following:
hence 𝑔 · ℎ · 𝑔 −1 ∈ ker(𝜑).
40
2.4. The Seifert-van Kampen theorem
(𝑔 · 𝐻) · ( 𝑔˜ · 𝐻) = (𝑔 𝑔)
˜ · 𝐻.
Normality of 𝐻 ensures that this multiplication is well defined. The group 𝐻 is then the kernel of
𝐺 → 𝐺/𝐻 with 𝑔 ↦→ 𝑔 · 𝐻. Thus the normal subgroups are exactly those which arise as kernels
of group homomorphisms.
Ù
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.
Definition 2.59
Let 𝐺 1 and 𝐺 2 be groups. A group 𝐺 is called free product of 𝐺1 ❆
𝐺 1 and 𝐺 2 iff there exist homomorphisms 𝑖 𝑗 : 𝐺 𝑗 → 𝐺 such 𝑖1 ⑥⑥ ❆❆ 𝜑
⑥ ❆❆ 1
⑥⑥ ❆❆
that for all groups 𝐻 and for all homomorphisms 𝜑 𝑗 : 𝐺 𝑗 → 𝐻 ~ ⑥⑥⑥
𝜑1★𝜑2
❆
there exists a unique homomorphism 𝐺 `❆ /𝐻
❆❆ >
❆❆ ⑥⑥⑥
❆ ⑥
𝜑1 ★ 𝜑2 : 𝐺 → 𝐻 𝑖2 ❆❆ ⑥⑥ 𝜑
⑥⑥ 2
𝐺2
such that the diagram to the right commutes.
41
2. Homotopy Theory
𝐺1 ❇
Namely, let 𝐺 ′ be another free product of 𝐺 1 and 𝐺 2 with 𝑖1 ⑥⑥⑥ ❇❇ 𝑖′
❇❇1
𝑖 ′𝑗 : 𝐺 𝑗 → 𝐺 ′ the corresponding homomorphisms. By the ⑥⑥⑥ ❇❇
~⑥⑥ ′ ′
𝑖1 ★𝑖2 !
universal property of 𝐺 with 𝐻 = 𝐺 ′ and 𝜑 𝑗 = 𝑖 ′𝑗 we get the 𝐺 `❆ / 𝐺′
❆❆ ⑤=
following commutative diagram: ❆❆
❆ ⑤⑤⑤
𝑖2 ❆❆ ⑤⑤ ′
⑤⑤ 𝑖2
𝐺2
𝐺1 ❆❆
𝑖1′ ⑤⑤ ❆❆ 𝑖1
⑤ ❆❆
⑤ ⑤⑤ ❆❆
} ⑤ 𝑖1★𝑖2
Interchanging the roles of 𝐺 and 𝐺 ′ we get another commutative ′
𝐺 a❇ /𝐺
diagram: ❇❇ ⑥⑥>
❇❇ ⑥⑥
❇ ⑥⑥𝑖2
𝑖2′ ❇❇ ⑥⑥
𝐺2
𝐺1 ❆
𝑖1 ⑥⑥ ′ ❆❆❆ 𝑖1
⑥⑥⑥ 𝑖1 ❆❆
⑥ ❆
Combining both diagrams we obtain ~⑥⑥𝑖1′ ★𝑖2′ 𝑖1★𝑖2 ❆
𝐺 `❆ / 𝐺O ′ /𝐺
❆❆ ⑥⑥>
❆❆ ′ ⑥
❆ 𝑖 ⑥⑥
𝑖2 ❆❆2 ⑥⑥⑥ 𝑖2
𝐺2
𝐺1 ❆
𝑖1 ⑥⑥ ❆❆ 𝑖
⑥⑥ ❆❆1
On the other hand, this diagram commute as well. ⑥⑥ ❆❆
❆
~⑥
⑥ id
𝐺 `❆ /𝐺
❆❆ ⑥⑥>
❆❆ ⑥⑥
❆ ⑥⑥
𝑖2 ❆❆ ⑥⑥ 𝑖2
𝐺1
By the uniqueness of the induced homomorphisms we have
(𝑖 1 ★ 𝑖 2 ) ◦ 𝑖 1′ ★ 𝑖 2′ = id𝐺
and similarly
𝑖 1′ ★ 𝑖 2′ ◦ (𝑖 1 ★ 𝑖2 ) = id𝐺 ′ .
42
2.4. The Seifert-van Kampen theorem
𝐺 1 ∩ 𝐺 2 = ∅.2 We define
𝐺 1 ★ 𝐺 2 := (𝑥1 , . . . 𝑥 𝑛 ) | 𝑛 ∈ N0 , 𝑥 𝑗 ∈ 𝐺 1 \ {1𝐺1 )} ∪ (𝐺 2 \ {1𝐺2 } such that
if 𝑥𝑖 ∈ 𝐺 1 then 𝑥𝑖+1 ∈ 𝐺 2 or conversely .
:= (𝑥1 , . . . , 𝑥 𝑛−1 ) · (𝑥 𝑛+2 , . . . , 𝑥 𝑛+𝑚 ) if (𝑥 𝑛 , 𝑥 𝑛+1 ∈ 𝐺 1 or 𝑥 𝑛 , 𝑥 𝑛+1 ∈ 𝐺 2 )
and 𝑥 𝑛 · 𝑥 𝑛+1 = 1,
(𝑥1 , . . . , 𝑥 𝑛 , 𝑥 𝑛+1 , . . . , 𝑥 𝑛+𝑚 )
otherwise.
This turns 𝐺 1 ★𝐺 2 into a group with neutral element the empty sequence (). The inverse element
for (𝑥1 , . . . 𝑥 𝑛 ) is given by (𝑥 𝑛−1 , . . . , 𝑥1−1 ) because of
(𝑥1 , . . . , 𝑥 𝑛 ) · 𝑥 𝑛−1 , . . . , 𝑥1−1 = (𝑥1 , . . . , 𝑥 𝑛−1 ) · 𝑥 𝑛−1
−1
, . . . , 𝑥1−1 = · · · = (𝑥1 ) (𝑥1−1 ) = ().
Remark 2.62
1.) The subset 𝑖 𝑗 (𝐺 𝑗 ) ⊂ 𝐺 1 ★ 𝐺 2 is a subgroup isomorphic to 𝐺 𝑗 . The intersection 𝑖1 (𝐺 1 ) ∩
𝑖2 (𝐺 2 ) = {1} is trivial. The union 𝑖1 (𝐺 1 ) ∪ 𝑖 2 (𝐺 2 ) generates 𝑖1 (𝐺 1 ) ★ 𝑖 2 (𝐺 2 ) as a group.
3.) If 𝐺 1 ≠ {1} and 𝐺 2 ≠ {1} then we may choose 𝑥 ∈ 𝐺 1 \ {1} and 𝑦 ∈ 𝐺 2 \ {1}. Then
43
2. Homotopy Theory
Remark 2.63. Usually one identifies 𝐺 1 with 𝑖 1 (𝐺 1 ) and 𝐺 2 with 𝑖2 (𝐺 2 ) and writes
𝑥1 · 𝑥2 · . . . · 𝑥 𝑛 instead of (𝑥1 , 𝑥2 , . . . , 𝑥 𝑛 ).
Example 2.64. Consider 𝐺 1 = Z/2Z = {1, −1} and 𝐺 2 = Z/2Z = {1′ , −1′ }. Elements of
Z/2Z ★ Z/2Z are for example
Now we calculate
Proposition 2.65. Let 𝑋 be a topological space and let 𝑈, 𝑉 ⊂ 𝑋 be open such that 𝑈 ∪𝑉 = 𝑋.
Let 𝑥0 ∈ 𝑈 ∩ 𝑉. Furthermore, assume that 𝑈, 𝑉 and 𝑈 ∩ 𝑉 are path-connected. Then 𝑋 is
path-connected and the map
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 𝑥0 by a path.
The statement of the proposition is equivalent to saying that
𝑖 # 𝜋1 (𝑈; 𝑥0 ) ∪ 𝑗 # 𝜋1 (𝑉; 𝑥0 )
generates 𝜋1 (𝑋; 𝑥0 ) as a group. Now let [𝜔] ∈ 𝜋1 (𝑋; 𝑥0 ) and subdivide the unit interval 𝐼 = [0, 1]
by 0 = 𝑡 0 ≤ 𝑡 1 · · · < 𝑡 𝑛 = 1 such that 𝜔([𝑡𝑖 , 𝑡 𝑖+1 ]) ⊂ 𝑈 or ⊂ 𝑉.
44
2.4. The Seifert-van Kampen theorem
𝑋
𝜔0 𝜔2
𝑈
𝑥0
𝑈 ∩𝑉 𝜂1 𝜂2 𝜂3 𝑈 ∩𝑉
𝜔(𝑡1 ) 𝜔(𝑡2 ) 𝜔(𝑡 3 )
𝜔1
45
2. Homotopy Theory
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; 𝑥0 ) ★ 𝜋1 (𝑉; 𝑥0 ) → 𝜋1 (𝑋; 𝑥0 )
is onto we get that 𝜋1 (𝑋; 𝑥0 ) = {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.
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 𝑥0 ∈ 𝑈 ∩ 𝑉. Let 𝑈, 𝑉 and 𝑈 ∩ 𝑉 be path connected. Then
the map 𝑖# ★ 𝑗 # induces an isomorphism
′ 𝜋1 (𝑋; 𝑥0 ).
𝜋1 (𝑈; 𝑥0 ) ★ 𝜋1 (𝑉; 𝑥0 )
N 𝑖 # (𝛼) · 𝑗 #′ (𝛼) −1 | 𝛼 ∈ 𝜋1 (𝑈 ∩ 𝑉; 𝑥0 )
46
2.4. The Seifert-van Kampen theorem
ker(𝑖 # ★ 𝑗 # ) ⊂ N 𝑖#′ (𝛼) · 𝑗 #′ (𝛼) −1 | 𝛼 ∈ 𝜋1 (𝑈 ∩ 𝑉; 𝑥0 ) .
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
1 −
| | |
𝜔1 𝜔2 𝜔𝑁
Figure 26. Subdividing the homotopy
47
2. Homotopy Theory
𝑢
>
𝑙 𝑟
∧ ∧
Considering the edges of the subsquares we
get homotopies
>
𝑑
48
2.4. The Seifert-van Kampen theorem
𝑢1 𝑢2 𝑢𝑁
> > >
𝜀 𝑥0 = 𝑙 1 ∧ ∧ ∧ ∧ ∧ 𝑟 𝑁 = 𝜀 𝑥0
> > >
𝑑1 𝑑2 𝑑𝑁
In
𝜋1 (𝑈; 𝑥0 ) ★ 𝜋1 (𝑉; 𝑥0 )
𝐺 :=
N ({𝑖 #′ (𝛼) · 𝑗 #′ (𝛼) −1 | 𝛼 ∈ 𝜋1 (𝑈 ∩ 𝑉; 𝑥0 )})
the computation (2.4) is still possible. By induction on all squares in the row from right to the
left we find that
z }| {
=1
[𝐷 1 ] 𝑊1 · . . . · [𝐷 𝑁 ] 𝑊𝑁 = [𝐿 1 ] 𝑊1 ·[𝐷 1 ] 𝑊1 · . . . · [𝐷 𝑁 ] 𝑊𝑁 = [𝑈1 ] 𝑊1 · . . . · [𝑈 𝑁 ] 𝑊𝑁 .
and hence
ker(𝑖 # ★ 𝑗# ) ⊂ N 𝑖 #′ (𝛼) · 𝑗 #′ (𝛼) −1 | 𝛼 ∈ 𝜋1 (𝑈 ∩ 𝑉; 𝑥0 ) .
49
2. Homotopy Theory
Corollary 2.69. Let 𝑋 be a topological space. Let 𝑈, 𝑉 ⊂ 𝑋 be open subsets such that
𝑈 ∪ 𝑉 = 𝑋 and let 𝑥0 ∈ 𝑈 ∩ 𝑉. Assume that 𝑈 and 𝑉 are path connected and that 𝑈 ∩ 𝑉 is
1-connected. Then
𝜋1 (𝑋; 𝑥0 ) 𝜋1 (𝑈; 𝑥0 ) ★ 𝜋1 (𝑉; 𝑥0 )
where the isomorphism is induced by the inclusion maps.
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
𝑈 ∩ 𝑉 ≃ point
𝑀
𝑉
b
𝑝
𝑈 ∩ 𝑉 ≈ R𝑛 \ {0} ≃ 𝑆 𝑛−1
50
2.5. The fundamental group of surfaces
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. Let 𝑋 = 𝑀#𝑁.
𝑀 𝑁
Now choose 𝑈 and 𝑉 as in Figure 32. Then we have that 𝑈 ≈ 𝑀 \ {𝑝} and 𝑉 ≈ 𝑁 \ {𝑞}.
𝑈 ∩𝑉
| {z }
=𝑈 | {z }
=𝑉
Figure 32. ... and consider their connected sum.
Since
𝑈 ∩ 𝑉 ≈ 𝑆 𝑛−1 × (0, 1) ≃ 𝑆 𝑛−1
is 1-connected we find by Corollary 2.69 once again that
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.3 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.
51
2. Homotopy Theory
𝐹𝑔 = ···
| {z }
𝑔−times
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.
𝐹𝑔 = 𝑇 2 # · · · #𝑇 2 ≈ 𝐷 2 /∼
| {z }
𝑔−times
𝑏 𝑔−1 𝑎1
𝑎 𝑔−1 𝑏1
𝑎1−1
𝑏1−1
··· ≈ 𝐷2
𝑎2
𝑏2
𝑎2−1
52
2.5. The fundamental group of surfaces
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
𝑏1−1 𝑎1 𝑎1
≈ 𝑏1−1 𝑏1 ≈ 𝑏1−1 𝑏1 ≈
𝑎1−1 𝑏1
𝑎1−1
Figure 35. Starting the induction with the torus
𝑏 𝑔−1 𝑎1 𝑎1
𝑏1 𝑏1
𝑎 𝑔−1 𝑏 𝑔−1
𝑎1−1 𝑎 𝑔−1 𝑎1−1
𝑐
𝑏1−1 𝑏1−1
≈ 𝑐
𝑎2
𝑎2
𝑏2
𝑏2
𝑎2−1
𝑎2−1
Figure 36. Induction step by cutting off a segment
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.
53
2. Homotopy Theory
𝑎1
𝑏1
𝑏 𝑔−1
𝑎1−1
𝑎 𝑔−1 𝑏 𝑔−1 𝑎2
𝑎 𝑔−1 𝑏2
𝑐 𝑏1−1 𝑎2−1 𝑏1−1 𝑎1
𝑐
𝑐 𝑏2−1 𝑐
≈
𝑎2
𝑏2 𝑎1−1 𝑏1
𝑎2−1
Figure 37. Closing the loop 𝑐
𝑐 𝑐 ≈
| {z } | {z }
(𝑔−1) −times 𝑔−times
(...(𝜑1 ★ 𝜑2 ) ★ 𝜑3 ) ★ · · · ★ 𝜑 𝑛 =: 𝜑1 ★ · · · ★ 𝜑 𝑛 : (...(Z ★ Z) ★ Z) ★ · · · ★ Z =: Z ★ · · · ★ Z → 𝐺
with
(𝜑1 ★ · · · ★ 𝜑 𝑛 ) ((𝑘 1 , 𝑖 1 ), ..., (𝑘 𝑟 , 𝑖𝑟 )) = 𝑔𝑖𝑘11 · · · 𝑔𝑖𝑘𝑟𝑟 ,
where 𝑖 𝑗 ∈ {1, ..., 𝑛} is an index used to make the Z-factors formally disjoint.
54
2.5. The fundamental group of surfaces
If the normal subgroup ker(𝜑1 ★ · · · ★ 𝜑 𝑛 ) is also finitely generated as a group, with generators
𝑥1 , ..., 𝑥 𝑚 , then we call 𝐺 finitely presentable and
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.
Z/2Z h𝑎 | 𝑎2 i.
Remark 2.80. A word of caution: isomorphic groups may have several, quite different presen-
tations. For example
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.
Proof. Recall that 𝐹𝑔 ≈ 𝐷 2 /∼ as in Proposition 2.76. To apply the Seifert-van Kampen theo-
rem 2.68 we put 𝑈 := 𝐷˚ 2 and 𝑉 := (𝐷 2 /∼) \ 𝐷 2 ( 21 ). We have 𝐹𝑔 = 𝑈 ∪ 𝑉.
55
2. Homotopy Theory
𝑏 𝑔−1 𝑎1
𝑏1
𝑎 𝑔−1
𝑎1−1
𝑏1−1
𝑉
𝑎2
𝑏2
𝑎2−1
Figure 39. Applying the Seifert-van Kampen theorem to the disk representation
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 𝑥0 ∈ 𝑋,
𝑦 0 ∈ 𝑌 such that 𝑋 ∨ 𝑌 := 𝑋 ∪ 𝑌 /{𝑥0 ∼ 𝑦 0 }. 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 ∨ · · · ∨ 𝑆1 .
| {z }
2𝑔 times
| {z }
𝜋1 (𝑉) Z ★ Z ★ · · · ★ Z
2𝑔 times
56
2.5. The fundamental group of surfaces
𝑏 𝑔−1 𝑎1
𝑏1
𝑎 𝑔−1
𝑎1−1
𝑏1−1
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 𝑏1−1 · · · 𝑎 𝑔 𝑏 𝑔 𝑎 𝑔−1 𝑏 𝑔−1 .
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 , ..., 𝑎 𝑔 , 𝑏 𝑔 | 𝑎1 𝑏1 𝑎1−1 𝑏1−1 · · · 𝑎 𝑔 𝑏 𝑔 𝑎 𝑔−1 𝑏 𝑔−1 i.
Example 2.82. For the two-dimensional torus 𝑇 2 we find with Example 2.79
57
2. Homotopy Theory
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 , ..., 𝑎 𝑔 , 𝑏 𝑔 | 𝑎1 𝑏1 𝑎1−1 𝑏1−1 · · · 𝑎 𝑔 𝑏 𝑔 𝑎 𝑔−1 𝑏 𝑔−1 , 𝑎1 𝑏1 𝑎1−1 𝑏1−1 , ...
[𝜋1 (𝐹𝑔 ), 𝜋1 (𝐹𝑔 )]
..., 𝑎1 𝑎2 𝑎1−1 𝑎2−1 , ...i
= h𝑎1 , 𝑏1 , ..., 𝑎 𝑔 , 𝑏 𝑔 | 𝑎1 𝑏1 𝑎1−1 𝑏1−1 , ..., 𝑎1 𝑎2 𝑎1−1 𝑎2−1 , ...i
Z2𝑔 ,
where the second equality follows because the simple commutators 𝑎𝑖 𝑏𝑖 𝑎𝑖−1 𝑏𝑖−1 , 𝑖 = 1, ..., 𝑔
generate the relation 𝑎1 𝑏1 𝑎1−1 𝑏1−1 · · · 𝑎 𝑔 𝑏 𝑔 𝑎 𝑔−1 𝑏 𝑔−1 .
Hence if 𝐹𝑔 ≃ 𝐹𝑔′ then 𝜋1 (𝐹𝑔 ) 𝜋1 (𝐹𝑔′ ) and thus
Remark 2.84. This proves the uniqueness part of the classification for surfaces, see Example 1.5.
Definition 2.85. Let 𝑊 𝑛 = [0, 1] × . . . × [0, 1] be the 𝑛-cube. Let 𝑋 be a topological space
| {z }
𝑛 times
and 𝑥0 ∈ 𝑋. Then
is called the 𝑛-th homotopy group of 𝑋 with base point 𝑥0 . Here [𝜎] 𝜕𝑊 𝑛 denotes the homotopy
class of 𝜎 relative to 𝜕𝑊 𝑛 .
58
2.6. Higher homotopy groups
𝑊𝑛 𝑊𝑛 𝑊𝑛
𝜎 𝜎★𝜏
𝑋 𝑡2 , . . . , 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 𝜋 𝑛 (𝑋; 𝑥0 ) 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 (𝑡1 , . . . , 𝑡 𝑛 ) = 𝑥0 , and [𝜎] 𝜕𝑊
−1 is
𝑛
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 𝑥0 ∈ 𝑋. Then for 𝑛 ≥ 2 the group
𝜋 𝑛 (𝑋; 𝑥0 ) is abelian.
Proof. For 𝜎, 𝜏 ∈ 𝐶 (𝑊 𝑛, 𝑋) with 𝜎 (𝜕𝑊 𝑛 ) = 𝜏(𝜕𝑊 𝑛 ) = {𝑥0 } 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 𝜋 𝑛 (𝑋; 𝑥0 ).
𝑓# : 𝜋 𝑛 (𝑋; 𝑥0 ) → 𝜋 𝑛 (𝑌 ; 𝑦 0 )
defined by [𝜎] 𝜕𝑊 𝑛 ↦→ [ 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 𝜋 𝑛 (𝑋; 𝑥0 ). In particular, if 𝑓 : 𝑋 → 𝑌 is a homotopy equivalence
then the map 𝑓# : 𝜋 𝑛 (𝑋; 𝑥0 ) → 𝜋 𝑛 (𝑌 ; 𝑓 (𝑥0 )) is an isomorphism. For 𝛾 ∈ Ω(𝑋; 𝑥0 , 𝑥1 ) there is
an isomorphism
Φ𝛾 : 𝜋 𝑛 (𝑋; 𝑥1 ) → 𝜋 𝑛 (𝑋; 𝑥0 )
given by [𝜎] 𝜕𝑊 𝑛 ↦→ [𝜎 ′ ] 𝜕𝑊 𝑛 where
𝛾
( 𝑥0 𝑥1 𝜎
𝜎 (2𝑡), k𝑡 k max ≤ 12 ,
𝜎 ′ (𝑡) = 1
𝛾(2 − 2k𝑡 k max ), 2 ≤ k𝑡 k max ≤ 1.
60
2.6. Higher homotopy groups
𝜋 𝜑
𝑊𝑛 / 𝑊 𝑛 /𝜕𝑊 𝑛
≈
/ 𝑆𝑛
Remark 2.90. The Seifert-van-Kampen theorem for 𝜋 𝑛 works only under very restrictive as-
sumptions. For this reason the computation of 𝜋 𝑛 for explicit examples can be very difficult. We
will be able to compute 𝜋 𝑛 (𝑆 𝑚 ) for 𝑛 ≤ 𝑚 but many of the 𝜋 𝑛 (𝑆 𝑚 ) for 𝑛 > 𝑚 are actually still
unknown.
But there is no (natural) group structure on 𝜋0 (𝑋; 𝑥0 ). More precisely, 𝜋0 (𝑋; 𝑥0 ) is a pointed set,
i.e., a set with a distinguished point, namely the path component containing 𝑥0 . This corresponds
to the neutral element in 𝜋 𝑛 (𝑋; 𝑥0 ) 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 (HLP for short) 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
b
𝑒0
Example 2.93. Consider the spaces 𝐸 = {point},
𝐵 = R, 𝑊 = 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
Figure 45. Failure of HLP due to
lack of surjectivity
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 = {𝑤 0 }
| 𝐵
ℎ
Figure 46. Surjective but HLP still fails
The map 𝑝 is surjective but still does not have the HLP for 𝑊 = 𝑊 0 . For example, choose
ℎ(𝑤 0 , 𝑡) = 𝑡, 𝑓 (𝑤 0 ) = (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 = 𝑝𝑟 1 , 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
Now put 𝐻 (𝑤, 𝑡) := (ℎ(𝑤, 𝑡), 𝜑(𝑤)). Then 𝐻 ∈ 𝐶 (𝑊 × [0, 1], 𝐵 × 𝐹) and
62
2.6. Higher homotopy groups
𝑊 𝑛 × [0, 1]
[0, 1]
𝑄 11
𝑊𝑛
Figure 47. Subdivision for which the fiber bundle is trivial over each subcube
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
𝑝𝑟1 ttt
t
t tt 𝑝 | 𝑝 −1 (𝑈)
y t
𝑈
commutes.
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] into small subcubes such that ℎ maps each subcube entirely into an
open subset 𝑈 as in the definition of the fiber bundle, see Figure 47.
By Example 2.96, products have the HLP, hence we can extend the map 𝑓 to a continuous map
𝐻11 : (𝑊 𝑛 × {0}) ∪ 𝑄 11 → 𝐸 such that 𝑝 ◦ 𝐻11 = ℎ.
Next we want to extend the lift to 𝑄 12 , see Figure 48. Now there seems to be a problem because
the required lift 𝐻12 need not only coincide with 𝑓 along the edge 𝑄 12 ∩ (𝑊 𝑛 × {0}) but also
with 𝐻11 along 𝑄 11 ∩ 𝑄 12 .
But there are homeomorphisms of a cube onto itself mapping two edges onto one as indicated in
Figure 49.
63
2. Homotopy Theory
𝑊 𝑛 × [0, 1]
[0, 1]
𝑄 11 𝑄 12
𝑊𝑛
Figure 48. Lift over second cube
64
2.6. Higher homotopy groups
Thus we can apply the HLP and extend the lift to (𝑊 𝑛 × {0}) ∪ 𝑄 11 ∪ 𝑄 12 . 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 ↦→ 𝑔 · 𝐻
is a fiber bundle with fiber 𝐻. The proof of this fact requires some technical work, see [8, p. 120
ff].
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# : 𝜋 𝑛 (𝐹; 𝑒0 ) → 𝜋 𝑛 (𝐸; 𝑒0 ),
𝑝 # : 𝜋 𝑛 (𝐸; 𝑒0 ) → 𝜋 𝑛 (𝐵; 𝑏0 ).
Now we construct a map 𝜕 : 𝜋 𝑛 (𝐵; 𝑏0 ) → 𝜋 𝑛−1 (𝐹; 𝑒0 ). We define
Box0 := {1} and Box 𝑘 := (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} .
65
2. Homotopy Theory
b
𝑥
b
𝜂 𝑘 (𝑥)
b
( 21 , . . . , 12 , −1)
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′ (𝑡1 , . . . , 𝑡 𝑛 ).
From ℎ(Box𝑛 ) = {𝑏0 } we have 𝐻 (Box𝑛 ) ⊂ 𝐹. Then we get a homotopy in 𝐹 relative to 𝜕𝑊 𝑛−1
−1 to 𝜎
˜ = Σ ◦ 𝜂 𝑛−1
from 𝜎 ˜ ′ = Σ′ ◦ 𝜂 𝑛−1
−1 as shown in Figure 52.
66
2.6. Higher homotopy groups
𝐹 𝐸
𝑡𝑛
b 𝑒0
𝜎 𝑝
b
𝑏0
𝑡 1 , . . . , 𝑡 𝑛−1
Σ′
Σ −1
𝑡 𝑛+1 𝑡𝑛
𝑡1 , . . . , 𝑡 𝑛−1
67
2. Homotopy Theory
𝜎′
𝐵
ℎ
b
𝑏0
𝑡 𝑛+1 𝑡𝑛
𝑡 1 , . . . , 𝑡 𝑛−1
𝜎
Figure 53. Homotopy invariance
yields a lift 𝐻 : 𝑊 𝑛+1 → 𝐸 of ℎ with 𝐻 (𝑡 1 , . . . , 𝑡 𝑛−1 , 0, 𝑡 𝑛+1 ) = 𝑒0 , see Figure 53. We thus obtain
a homotopy
−1
𝐻ˆ (𝑡1 , . . . , 𝑡 𝑛−1 , 𝑠) = 𝐻 (𝜂 𝑛−1 (𝑡 1 , . . . , 𝑡 𝑛−1 ), 𝑠)
in 𝐹 from 𝜎 ˜ ′ relative to 𝜕𝑊 𝑛−1 and hence
˜ to 𝜎
˜ ′ ] 𝜕𝑊 𝑛−1 .
˜ 𝜕𝑊 𝑛−1 = [ 𝜎
[ 𝜎]
We have shown that the map 𝜕 : 𝜋 𝑛 (𝐵; 𝑏0 ) → 𝜋 𝑛−1 (𝐹, 𝑒0 ) is well defined.
Lemma 2.101. For 𝑛 ≥ 2 the map 𝜕 : 𝜋 𝑛 (𝐵; 𝑏0 ) → 𝜋 𝑛−1 (𝐹, 𝑒0 ) is a group homomorphism.
68
2.6. Higher homotopy groups
Σ ★𝑇 𝐻
−1 ) ★ (𝑇 ◦ 𝜂 −1 )
(Σ ◦ 𝜂 𝑛−1 −1
𝐻 ◦ 𝜂 𝑛−1
𝑛−1
b b b
𝜕 𝜄# 𝑝# 𝜕 𝜄#
... / 𝜋 𝑛 (𝐹; 𝑒 𝑜 ) / 𝜋 𝑛 (𝐸; 𝑒 0 ) / 𝜋 𝑛 (𝐵; 𝑏 0 ) / 𝜋 𝑛−1 (𝐹; 𝑒 0 ) / ...
𝜕 𝜄# 𝑝#
... / 𝜋0 (𝐹; 𝑒 0 ) / 𝜋0 (𝐸; 𝑒 0 ) / 𝜋0 (𝐵; 𝑏 0 )
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.
69
2. Homotopy Theory
[0, 1]
𝑊𝑛
Figure 55. Bottom-to-box homotopy
We obtain a homotopy in 𝐸 relative to 𝜕𝑊 𝑛 from 𝜏 to 𝐻 ◦ 𝜂 𝑛−1 . Hence 𝜏 ≃𝜕𝑊 𝑛 𝐻 ◦ 𝜂 𝑛−1
and we conclude that
[𝜏] 𝐸,𝜕𝑊 𝑛 = [𝐻 ◦ 𝜂 𝑛−1 ] 𝐸,𝜕𝑊 𝑛 = 𝜄# ([𝐻 ◦ 𝜂 𝑛−1 ] 𝐹,𝜕𝑊 𝑛 ) ∈ im 𝜄# .
ii) ker 𝜄# ⊂ im 𝜕:
Let [𝜏] 𝜕𝑊 𝑛 ∈ 𝜋 𝑛 (𝐹; 𝑒0 ) with 0 = 𝜄# ([𝜏] 𝜕𝑊 𝑛 ). Hence 𝜏 ≃𝜕𝑊 𝑛 𝜀 𝑛𝑒0 in 𝐸. Let 𝐻 be a
homotopy in 𝐸 relative to 𝜕𝑊 𝑛 from 𝜀 𝑛𝑒0 to 𝜏, see Figure 57. 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
𝜋 𝑛+1 (𝐵; 𝑏0 ). By definition of 𝜕, we have [𝐻 ◦ 𝜂 𝑛−1 ] 𝜕𝑊 𝑛 = 𝜕 ([ℎ] 𝜕𝑊 𝑛+1 ).
In the image of Box𝑛 under 𝜂 𝑛 we let the interior cube grow and thereby obtain a homotopy
in 𝐹 relative to 𝜕𝑊 𝑛 from 𝐻 ◦𝜂 𝑛−1 to 𝜏, see Figure 58. Therefore [𝜏] 𝜕𝑊 𝑛 = [𝐻 ◦𝜂 𝑛−1
−1 ]
𝜕𝑊 𝑛 =
𝜕 [ℎ] 𝜕𝑊 𝑛+1 ∈ im(𝜕).
70
2.6. Higher homotopy groups
𝜀 𝑛𝑒0
Figure 56. Bottom-to-box homotopy, again
[0, 1]
𝜀 𝑛𝑒0
𝑊𝑛
Figure 57. Homotopy between 𝜏 and constant map
−1
𝜕 ( 𝑝 # ([𝜏] 𝜕𝑊 𝑛 ) = 𝜕 ([ 𝑝 ◦ 𝜏] 𝜕𝑊 𝑛 )) = [𝜏 ◦ 𝜂 𝑛−1 ] 𝜕𝑊 𝑛−1 = [𝜀 𝑛−1
𝑒0 ] 𝜕𝑊 𝑛−1 = 0.
ii) ker 𝜕 ⊂ im 𝑝 # :
Let [𝜎] 𝜕𝑊 𝑛 ∈ 𝜋 𝑛 (𝐵; 𝑏0 ) with 𝜕 ([𝜎] 𝜕𝑊 𝑛 ) = 0. Let Σ be a lift of 𝜎 with initial condition
𝜀 𝑛−1 −1
𝑒0 . Then Σ ◦ 𝜂 𝑛−1 represents 𝜕 ([𝜎] 𝜕𝑊 𝑛 ) = 0. Hence we have in 𝐹
−1
Σ ◦ 𝜂 𝑛−1 ≃𝜕𝑊 𝑛−1 𝜀 𝑛−1
𝑒0
71
2. Homotopy Theory
𝑒0
𝑒0
𝑒0 𝜏 𝑒0 −→ 𝑒0
deform 𝜏 𝑒0 deform
−→ 𝜏
𝑒0
𝑒0
𝑒0
maps to 𝐹
𝑒0 𝐹 𝐹 𝑒0
Σ
𝑒0
Figure 59. Extension of Σ
𝑝 # : 𝜋 𝑛 (𝐸, 𝑒0 ) → 𝜋 𝑛 (𝐵; 𝑏0 )
72
2.6. Higher homotopy groups
𝑏0
deform deform
−→ −→
𝑏0 𝜎 𝑏0 𝜎 𝜎
𝜄# 𝑝# 𝜕
{0} = 𝜋 𝑛 (𝐹; 𝑒 𝑜 ) / 𝜋 𝑛 (𝐸; 𝑒 0 ) / 𝜋 𝑛 (𝐵; 𝑏 0 ) / 𝜋 𝑛−1 (𝐹; 𝑒 0 ) = {0}.
Example 2.106. The map Exp : R → 𝑆1 with 𝑡 ↦→ 𝑒2 𝜋𝑖𝑡 is a covering with fiber Z. Hence
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 ↦→ [𝑥] ∼ is a covering with fiber
Z/2Z. Hence 𝜋 𝑘 (RP𝑛 ) 𝜋 𝑘 (𝑆 𝑛 ) for all 𝑘 ≥ 2. For 𝑛 ≥ 2 we investigate the sequence:
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
We know that 𝜋 𝑘 (𝑆1 ) = {0} and 𝜋 𝑘−1 (𝑆1 ) = {0}. Hence we find
73
2. Homotopy Theory
𝑥 2 + 𝑦 2 − 𝑢2 − 𝑣 2
© ª
2(𝑦𝑢 − 𝑥𝑣) 2(𝑦𝑣 − 𝑥𝑢)
𝑓 (𝑥, 𝑦, 𝑢, 𝑣) = 2(𝑦𝑢 + 𝑥𝑣) 2 2
𝑥 −𝑦 +𝑢 −𝑣2 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
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
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.
2.7. Exercises
2.1. Let 𝑋 be a set and let 𝑥0 ∈ 𝑋. Determine 𝜋1 (𝑋; 𝑥0 ) where
2.2. Let 𝑋 be a topological space and let 𝜔 : 𝑆1 → 𝑋 be continuous. Show that the following
are equivalent:
74
2.7. Exercises
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.
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.
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) Show that each homeomorphism 𝑓 : 𝐷 2 → 𝐷 2 maps the boundary onto itself, 𝑓 (𝜕𝐷 2 ) =
𝜕𝐷 2 .
75
2. Homotopy Theory
a) 𝐺 = Z/2Z × Z/3Z;
b) Z ∗ Z;
c) 𝐺 = (Z ∗ Z) × (Z ∗ Z);
d) 𝐺 = Z ∗ Z/[Z ∗ Z, Z ∗ Z].
2.15. Decide whether or not the groups 𝐺 and 𝐻 are isomorphic where
2.16. Let 𝑋 be a path-connected topological space. Show that the suspension Σ𝑋 (see Exer-
cise 1.11) is also path-connected.
76
2.7. Exercises
2.17. Show that the map C → C, 𝑧 ↦→ 𝑧2 , has the homotopy lifting property for 𝑊 0 but not for
𝑊 1.
𝜋1 (𝑋, {0}) Z
𝜋1 (RP𝑛 ) Z/2Z.
2.20. 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.
2.21. Let 𝐸 = 𝐵 × 𝐹 and 𝑝 = pr1 : 𝐸 → 𝐵 be the product fibration. Show that the boundary
map 𝜕 : 𝜋 𝑛+1 (𝐵; 𝑏0 ) → 𝜋 𝑛 (𝐹; 𝑒0 ) is trivial in this case.
𝜋 𝑘 (SU(𝑛)) 𝜋 𝑘 (U(𝑛))
for all 𝑘 ≥ 2.
2.23. 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 ↦→ [𝑧] ∼ , is called the Hopf fibration.
Compute
𝜋 𝑘 (C𝑃 𝑛 )
for all 𝑘 ≤ 2𝑛.
Hint: You can use 𝜋 𝑘 (𝑆 𝑚 ) = {0} for all 1 ≤ 𝑘 < 𝑚 and 𝑚 ≥ 2.
77
2. Homotopy Theory
𝑖 𝑝
2.26. 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 ↦→ (𝑎, 0) and (𝑎, 𝑎′′ ) ↦→ 𝑎′′ , respectively.
If these conditions hold then we say that the exact sequence is split.
𝑖 𝑝
2.27. 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.
𝑒0 = (0, . . . , 0) ∈ R𝑛
𝑒1 = (1, . . . , 0) ∈ R𝑛
..
.
𝑒 𝑛 = (0, . . . , 1) ∈ R𝑛
Now we define
( )
Õ
𝑛 Õ
𝑛
Δ𝑛 := convex hull of 𝑒0 , . . . , 𝑒 𝑛 = 𝑡 𝑖 𝑒𝑖 𝑡 𝑖 ≥ 0, 𝑡𝑖 ≤ 1 .
𝑖=0 𝑖=0
Example 3.1. For 𝑛 = 0, 1, 2, and 3 the standard simplices are familiar, compare Figure 61:
79
3. Homology Theory
𝑒2 b b
Δ3
b b b b
𝑒0 𝑒1
Definition 3.3. We define the set of singular 𝑛-chains 𝑆 𝑛 (𝑋; 𝑅) as the free 𝑅-module generated
by 𝐶 (Δ𝑛 , 𝑋).
Í𝑚
Then 𝑆𝑛 (𝑋; 𝑅) is an 𝑅-module. Elements of 𝑆𝑛 (𝑋; 𝑅) are formal linear combinations 𝑖=1 𝛼𝑖 𝜎𝑖
where 𝛼𝑖 ∈ 𝑅 and 𝜎𝑖 ∈ 𝐶 (Δ𝑛 , 𝑋). See Appendix A.1 for more on free modules generated by
sets.
For 𝑛 ≥ 0 we consider the affine linear map 𝐹𝑛+1 𝑖 : Δ𝑛 → Δ𝑛+1 given by
(
𝑖 𝑒 𝑗, 𝑗 <𝑖
𝐹𝑛+1 (𝑒 𝑗 ) = (3.1)
𝑒 𝑗+1 , 𝑗 ≥ 𝑖
𝑋
𝑒2 𝜎
b
𝐹21
b
b
𝑒0 𝑒1 𝑒0 𝑒1
𝜎 (𝑖)
Figure 62. Face maps
80
3.1. Singular homology
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 𝑗
𝐹𝑛𝑖 ◦ 𝐹𝑛−1 = 𝐹𝑛 ◦ 𝐹𝑛−1
𝑖−1
. (3.2)
We compute
!
Õ
𝑛
𝜕𝜕𝜎 = 𝜕 (−1) 𝑖 𝜎 ◦ 𝐹𝑛𝑖
𝑖=0
Õ
𝑛
= (−1) 𝑖 𝜕 𝜎 ◦ 𝐹𝑛𝑖
𝑖=0
Õ𝑛 Õ
𝑛−1
𝑖 𝑗
= (−1) (−1) 𝑗 𝜎 ◦ 𝐹𝑛𝑖 ◦ 𝐹𝑛−1
Õ Õ
𝑖=0 𝑗=0
𝑗 𝑗
= (−1) 𝑖+ 𝑗 𝜎 ◦ 𝐹𝑛𝑖 ◦ 𝐹𝑛−1 + (−1) 𝑖+ 𝑗 𝜎 ◦ 𝐹𝑛𝑖 ◦ 𝐹𝑛−1
Õ Õ
0≤ 𝑗<𝑖≤𝑛 0≤𝑖≤ 𝑗 ≤𝑛−1
𝑗 ′ ′ −1 𝑗′ ′
= (−1) 𝑖+ 𝑗 𝜎 ◦ 𝐹𝑛 𝑖−1
◦ 𝐹𝑛−1 + (−1) 𝑗 +𝑖 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.
81
3. Homology Theory
Example 3.9. Assume that 𝑋 = {point}. Then there is only one singular 𝑛-simplex, namely
the constant map 𝜎𝑛 : Δ𝑛 → 𝑋. In other words, 𝑆𝑛 (𝑋; 𝑅) = 𝑅 · 𝜎𝑛 . Consequently,
𝜎𝑛(𝑖) = 𝜎𝑛 ◦ 𝐹𝑛𝑖 = 𝜎𝑛−1 and
Õ
𝑛
0, 𝑛 odd,
𝑖
𝜕𝜎𝑛 = (−1) 𝜎𝑛−1 = 𝜎𝑛−1 , 𝑛 even, 𝑛 ≠ 0,
0,
𝑖=0 𝑛 = 0.
This implies (
𝑅 · 𝜎𝑛 , 𝑛 odd or 𝑛 = 0,
𝑍 𝑛 (𝑋; 𝑅) =
0, 𝑛 even and 𝑛 ≠ 0,
and (
𝑅 · 𝜎𝑛 , 𝑛 odd,
𝐵𝑛 (𝑋; 𝑅) =
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
Õ
𝑚 Õ
𝑚
𝑆𝑛 ( 𝑓 ) 𝛼𝑖 · 𝜎𝑖 := 𝛼𝑖 · ( 𝑓 ◦ 𝜎𝑖 ).
𝑖=1 𝑖=1
82
3.2. Relative homology
𝑆𝑛 ( 𝑓 )
𝑆𝑛 (𝑋; 𝑅) / 𝑆 𝑛 (𝑌 ; 𝑅)
𝜕 𝜕
𝑆𝑛−1 ( 𝑓 )
𝑆𝑛−1 (𝑋; 𝑅) / 𝑆 𝑛−1 (𝑌 ; 𝑅)
Proof. We compute
𝜕 (𝑆𝑛 ( 𝑓 ) (𝜎)) = 𝜕 ( 𝑓 ◦ 𝜎)
Õ𝑛
= (−1) 𝑖 ( 𝑓 ◦ 𝜎) ◦ 𝐹𝑛𝑖
𝑖=0
Õ
𝑛
= (−1) 𝑖 𝑓 ◦ (𝜎 ◦ 𝐹𝑛𝑖 )
𝑖=0
Õ
𝑛
= 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
(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.
83
3. Homology Theory
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) .
Since 𝑍 𝑛 (𝑋, 𝐴) = 𝑍 𝑛′ (𝑋, 𝐴)/𝑆𝑛 ( 𝐴) and 𝐵𝑛 (𝑋, 𝐴) = 𝐵′𝑛 (𝑋, 𝐴)/𝑆𝑛 ( 𝐴) we obtain
𝑍 𝑛′ (𝑋, 𝐴)/𝑆𝑛 ( 𝐴) 𝑍 𝑛′ (𝑋, 𝐴)
𝐻𝑛 (𝑋, 𝐴) = = .
𝐵′𝑛 (𝑋, 𝐴)/𝑆𝑛 ( 𝐴) 𝐵′𝑛 (𝑋, 𝐴)
𝑍 𝑛′ (𝑋, ∅) = 𝑍 𝑛 (𝑋),
𝐵′𝑛 (𝑋, ∅) = 𝐵𝑛 (𝑋),
𝐻𝑛 (𝑋, ∅) = 𝐻𝑛 (𝑋).
Example 3.12. Let 𝑋 = 𝑆1 × [0, 1] be the cylinder over 𝑆1 and 𝐴 = 𝑆1 × {0} ⊂ 𝑋, see Figure 63.
To construct an element in the relative homology 𝐻1 (𝑋, 𝐴) take the 1-simplex
𝜎 : Δ1 → 𝑋,
(𝑡𝑒1 + (1 − 𝑡)𝑒0 ) ↦→ (cos (2𝜋𝑡), sin (2𝜋𝑡), 1),
see Figure 64. Since 𝜎 is a closed curve in 𝑋 we find for its boundary
𝜕𝜎 = 𝜎 (𝑒1 ) − 𝜎 (𝑒0 ) = 0.
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 Figure 65.
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).
84
3.2. Relative homology
𝐴
Figure 63. Cylinder relative to bottom
|
>
𝜎 𝑋
Δ1
𝐴
85
3. Homology Theory
𝜎
>
<
𝜎1 𝑋
Δ2
𝜎2
𝐴
Figure 65. 𝜎 is null-homologous
Let (𝑋, 𝐴) and (𝑌 , 𝐵) be pairs of spaces and 𝑓 ∈ 𝐶 ((𝑋, 𝐴), (𝑌 , 𝐵)). Then, for each 𝑛 ∈ N0 , we
have the commutative diagram
𝑆𝑛 ( 𝑓 )
𝑆𝑛 (𝑋) / 𝑆 𝑛 (𝑌 )
O O
? 𝑆𝑛 ( 𝑓 | 𝐴 ) ?
𝑆𝑛 ( 𝐴) / 𝑆 𝑛 (𝐵)
𝑆𝑛 ( 𝑓 )
𝑆𝑛 (𝑋) / 𝑆 𝑛 (𝑌 )
𝑆𝑛 ( 𝑓 )
𝑆𝑛 (𝑋, 𝐴) / 𝑆 𝑛 (𝑌 , 𝐵)
commutes. Combining with Lemma 3.10 and diagram (3.3) we get the commutative diagram
𝑆𝑛 ( 𝑓 )
𝑆𝑛 ( 𝑓 ) )
𝑆𝑛 (𝑋, 𝐴) o 𝑆𝑛 (𝑋) / 𝑆 𝑛 (𝑌 ) / 𝑆 𝑛 (𝑌 , 𝐵)
𝜕¯ 𝜕 𝜕 𝜕¯
𝑆𝑛−1 ( 𝑓 )
𝑆𝑛−1 (𝑋, 𝐴) o 𝑆𝑛−1 (𝑋) / 𝑆 𝑛−1 (𝑌 ) / 𝑆 𝑛−1 (𝑌 , 𝐵)
5
𝑆𝑛−1 ( 𝑓 )
86
3.2. Relative homology
𝐻𝑛 ( 𝑓 )
𝐻𝑛 (𝑋, 𝐴) / 𝐻𝑛 (𝑌 , 𝐵)
commutes.
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
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 ↦→ 𝜕 (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.
𝐻𝑛 ( 𝑓 )
𝐻𝑛 (𝑋, 𝐴) / 𝐻𝑛 (𝑌 , 𝐵)
𝜕 𝜕
𝐻𝑛−1 ( 𝑓 | 𝐴 )
𝐻𝑛−1 ( 𝐴) / 𝐻𝑛−1 (𝐵)
Proof. We compute
Õ Õ
𝐻𝑛−1 ( 𝑓 | 𝐴) 𝜕 𝛼𝑖 𝜎𝑖 = 𝐻𝑛−1 ( 𝑓 | 𝐴) 𝜕 𝛼𝑖 𝜎𝑖
Õ Õ
𝑖 𝑖
𝑗
= 𝐻𝑛−1 ( 𝑓 | 𝐴) 𝛼𝑖 (−1) 𝑗 𝜎 𝑗 ◦ 𝐹𝑛−1
Õ Õ
𝑖 𝑗
𝑗
= 𝛼𝑖 (−1) 𝑗 ( 𝑓 | 𝐴) ◦ (𝜎𝑖 ◦ 𝐹𝑛−1 )
Õ Õ
𝑖 𝑗
𝑗
= 𝛼𝑖 (−1) 𝑗 ( 𝑓 ◦ 𝜎𝑖 ) ◦ 𝐹𝑛−1
𝑖 𝑗
87
3. Homology Theory
Õ
= 𝜕 𝛼𝑖 ( 𝑓 ◦ 𝜎𝑖 )
Õ
𝑖
=𝜕 𝛼𝑖 ( 𝑓 ◦ 𝜎𝑖 )
Õ
𝑖
= 𝜕 𝐻𝑛 ( 𝑓 ) 𝛼𝑖 𝜎𝑖 .
𝑖
Dimension Axiom.
𝑅, 𝑛 = 0
𝐻𝑛 ({point}; 𝑅)
0, otherwise
The next axiom deals with homotopy invariance. Two continuous maps 𝑓0 , 𝑓1 : (𝑋, 𝐴) → (𝑌 , 𝐵)
of pairs of spaces are called homotopic (in symbols 𝑓0 ≃ 𝑓1 ) iff there exists an 𝐻 ∈ 𝐶 (𝑋×[0, 1], 𝑌 )
such that for all 𝑥 ∈ 𝑋
𝐻 (𝑥, 0) = 𝑓0 (𝑥),
𝐻 (𝑥, 1) = 𝑓1 (𝑥),
𝐻 ( 𝐴 × [0, 1]) ⊂ 𝐵.
Homotopy Axiom.
Let 𝑓0 , 𝑓1 ∈ 𝐶 ((𝑋, 𝐴), (𝑌 , 𝐵)) be homotopic, 𝑓0 ≃ 𝑓1 . Then the induced maps on homology
coincide, i.e.,
𝐻𝑛 ( 𝑓0 ) = 𝐻𝑛 ( 𝑓1 ) : 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).
88
3.3. The Eilenberg-Steenrod axioms and applications
Exactness Axiom.
For any pair of spaces (𝑋, 𝐴) and inclusion maps 𝑖 : 𝐴 ↩→ 𝑋, 𝑗 : (𝑋, ∅) ↩→ (𝑋, 𝐴), the sequence
𝐻𝑛+1 ( 𝑗)
𝐻𝑛+1 (𝑋, 𝐴) o ···
𝜕
𝐻𝑛 (𝑖) 𝐻𝑛 ( 𝑗)
𝐻𝑛 ( 𝐴) / 𝐻𝑛 (𝑋) / 𝐻𝑛 (𝑋, 𝐴)
𝜕
𝐻𝑛−1 (𝑖)
··· o 𝐻𝑛−1 ( 𝐴)
the closure of 𝐴.
Excision Axiom.
For every pair of spaces (𝑋, 𝐴) and every 𝑈 ⊂ 𝐴 with 𝑈¯ ⊂ 𝐴˚ the homomorphism
𝐻𝑛 ( 𝑗) : 𝐻𝑛 (𝑋 \ 𝑈, 𝐴 \ 𝑈) → 𝐻𝑛 (𝑋, 𝐴)
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 will show their
usefulness by studying some basic examples.
Í
1.) If 𝑋 ≠ ∅ is path-connected then 𝐻0 (𝑋; 𝑅) 𝑅 and generators are represented by every 0-
Í
simplex 𝜎 : Δ0 → 𝑋. In other words, the isomorphism 𝐻0 (𝑋; 𝑅) → 𝑅 is given by 𝑗 𝛼 𝑗 𝜎 𝑗 ↦→
𝑗 𝛼𝑗.
É
2.) If 𝑋 𝑘 , 𝑘 ∈ 𝐾, are the path-components of 𝑋 then 𝐻𝑛 (𝑋; 𝑅) 𝑘 ∈𝐾 𝐻 𝑛 (𝑋 𝑘 ; 𝑅).
89
3. Homology Theory
𝐻0 (𝑆0 ) −→ 𝐻0 (𝐷 1 ) −→ 𝐻0 (𝐷 1 , 𝑆0 ) −→ 0
𝑅2 𝑅
∈
Since the map (𝑎, 𝑏) ↦→ 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.
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
=
90
3.3. The Eilenberg-Steenrod axioms and applications
For 𝑛 ≥ 2 we have
0
𝐻1 (𝐷 𝑛 ) −→ 𝐻1 (𝐷 𝑛 , 𝑆 𝑛−1 ) −→ 𝐻0 (𝑆 𝑛−1 ) −→ 𝐻0 (𝐷 𝑛 )
=
0
𝑆𝑛
𝐷 𝑛−
e) Put 𝑈 −𝑛 := {𝑥 ∈ 𝑆 𝑛 | 𝑥0 < − 21 }.
𝑆𝑛
𝐷 𝑛−
𝑈 −𝑛
(𝑆 𝑛 \ 𝑈 −𝑛 , 𝐷 𝑛− \ 𝑈 −𝑛 ) ↩→ (𝑆 𝑛 , 𝐷 𝑛− )
1 For later use, we also define the upper hemisphere 𝐷 +𝑛 = {𝑥 ∈ 𝑆 𝑛 | 𝑥0 ≥ 0}.
91
3. Homology Theory
induces an isomorphism
𝐻𝑚 (𝑆 𝑛 \ 𝑈 −𝑛 , 𝐷 𝑛− \ 𝑈 −𝑛 ) 𝐻𝑚 (𝑆 𝑛 , 𝐷 𝑛− )
(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
)= (3.5)
0, otherwise
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 ) (3.6)
Induction over 𝑚 concludes the proof.
Remark 3.17. Let us describe a generator of 𝐻1 (𝑆1 ) Z geometrically. We use the isomor-
phisms in (3.5) and start with 𝐻1 (𝐷 1 , 𝑆0 ). The map 𝑐 : Δ1 → 𝐷 1 , 𝑡 ↦→ cos(𝜋(1 − 𝑡)), is a
singular 1-simplex with 𝜕𝑐 = const1 − const −1 . Under the isomorphism 𝐻0 (𝑆0 ) 𝑅 2 it maps
to (1, −1) which generates the kernel of the map 𝑅 2 → 𝑅 given by (𝑎, 𝑏) ↦→ 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′ : Δ1 → 𝐷 1+ , 𝑡 ↦→ 𝑒𝑖 𝜋 (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
92
3.3. The Eilenberg-Steenrod axioms and applications
𝐹 (𝑡, 0) = 𝑐′ (𝑡),
𝐹 (𝑡, 1) = 𝑒𝑖 𝜋 (1−2𝑡 ) =: 𝑐′′ (𝑡),
𝐹 (0, 𝑠) = −1 ∈ 𝐷 1− ,
𝐹 (1, 𝑠) = 𝑒 −𝑖 𝜋𝑠 ∈ 𝐷 1− .
Thus 𝑐′ ≃ 𝑐′′ as maps (Δ1 , {0, 1}) → (𝑆1 , 𝐷 1− ). Using the homotopy axiom we compute
[𝑐′ ] = [𝑐′ ◦ idΔ1 ] = 𝐻1 (𝑐′ ) [idΔ1 ] = 𝐻1 (𝑐′′ ) [idΔ1 ] = [𝑐′′ ◦ idΔ1 ] = [𝑐′′ ].
Therefore [𝑐′′ ] ∈ 𝐻1 (𝑆1 , 𝐷 1− ) is a generator. Since 𝜕𝑐′′ = const𝑐′′ (1) − const 𝑐′′ (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′′ ] ∈ 𝐻1 (𝑆1 ) is a generator.
Using the homotopy (𝑡, 𝑠) ↦→ 𝑒𝑖 𝜋 (𝑠−2𝑡 ) we see that 𝑡 ↦→ 𝑒 −2𝑖 𝜋𝑡 also represents a generator of
𝐻1 (𝑆1 ). Finally, playing the same game with lower and upper hemispheres interchanged shows
that 𝑡 ↦→ 𝑒2𝑖 𝜋𝑡 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.
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).
𝐷𝑛
𝐻𝑛−1 (id)=id
𝐻𝑛−1 (𝑆 𝑛−1 ; Z)❙ Z / 𝐻𝑛−1 (𝑆 𝑛−1 ; Z) Z
5
❙❙❙❙
❙❙❙❙ ❦❦ ❦❦❦❦❦
❙❙
𝐻𝑛−1 ( 𝜄) ❙❙❙❙) ❦❦❦❦
❦❦❦❦ 𝐻𝑛−1 (𝑔)
𝐻𝑛−1 (𝐷 𝑛 ; Z) = 0
93
3. Homology Theory
Since the identity Z → Z does not factor through 0 we run into a contradiction.
𝑠(𝑥0 , 𝑥1 , . . . , 𝑥 𝑛 ) = (−𝑥0 , 𝑥1 , . . . , 𝑥 𝑛 ).
Proof. The proof is given by induction. First consider the case 𝑛 = 1. Let 𝑐 : Δ1 → 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 as indicated in Figure 70.
94
3.3. The Eilenberg-Steenrod axioms and applications
𝑒2
b 𝜎1
b
∧
b b
𝑒0 𝑒1
b b
𝑒0 𝑒1
Figure 71. Second 2-simplex 𝜎2
where the horizontal isomorphisms are the ones in (3.6). 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− .
95
3. Homology Theory
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.
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 ∈ 𝑆 𝑛 .
𝑥
𝑣(𝑥)
b
Theorem 3.23 (Hairy ball theorem). The 𝑛-dimensional sphere 𝑆 𝑛 admits a continuous vec-
tor 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) := (−𝑥1 , 𝑥0 , −𝑥3 , 𝑥2 , . . . , −𝑥 𝑛 , 𝑥 𝑛−1 ). This defines a nowhere
vanishing continuous vector field.
Now let 𝑛 be even and let 𝑣 : 𝑆 𝑛 → R𝑛+1 be a continuous vector
field without a zero. Then we can put 𝑤 (𝑥) := | 𝑣𝑣 (( 𝑥)
𝑥) | . We define
𝑥
the continuous map 𝐹 : 𝑆 𝑛 × [0, 1] → 𝑆 𝑛 by
∧ b
𝐻𝑛 ( 𝑓 ) : 𝐻𝑛 (𝑆 𝑛 ; Z) Z → 𝐻𝑛 (𝑆 𝑛 ; Z) 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 ).
2.) For the antipodal map we have seen that deg(𝑎) = (−1) 𝑛+1 .
(i) deg(id) = 1;
(ii) deg(const) = 0;
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:
𝐻𝑛+1 ( 𝑓 )
𝐻𝑛+1 (𝐷 𝑛+1 , 𝑆 𝑛 ) / 𝐻𝑛+1 (𝐷 𝑛+1 , 𝑆 𝑛 )
𝜕 𝜕
𝐻𝑛+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.
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
(1 − 𝑡) 𝑓 (𝑥) − 𝑡𝑥
𝐹 : 𝑆 𝑛 × [0, 1] → 𝑆 𝑛 , 𝐹 (𝑥, 𝑡) := .
|(1 − 𝑡) 𝑓 (𝑥) − 𝑡𝑥|
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
(1 − 𝑡) 𝑓 (𝑥) + 𝑡𝑥
𝐺 (𝑥, 𝑡) := .
|(1 − 𝑡) 𝑓 (𝑥) + 𝑡𝑥|
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
𝑗1 : 𝑆𝑛 → 𝑆𝑛 × 𝑆𝑛 , 𝑥 ↦→ (𝑥, 𝑝),
𝑛 𝑛 𝑛
𝑗2 : 𝑆 → 𝑆 × 𝑆 , 𝑥 ↦→ ( 𝑝, 𝑥) .
(𝐻𝑛 ( 𝑗1 ),𝐻𝑛 ( 𝑗2 ) ) 𝐻𝑛 ( 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 𝑗 𝜈 and 𝑗 𝜈′ .
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 ◦ 𝑗 1 = id and thus
𝑑1 = deg(𝜇 ◦ 𝑗1 ) = deg(id) = 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 + ℎ1𝑖 + ℎ2 𝑗 + ℎ3 𝑘 .
99
3. Homology Theory
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 ↦→ 𝑧 −1 = 𝑧¯ 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
implies that deg( 𝑓 −1 ) = −1 in the quaternionic case too. The horizontal isomorphisms are
obtained as the composition of the isomorphisms
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 , 𝑓1 ))
deg( 𝑓 𝑘−1 )
= (1, 1) ·
deg( 𝑓1 )
= deg( 𝑓 𝑘−1 ) + deg( 𝑓1 )
= 𝑘 − 1 + 1 = 𝑘.
Now we can show that the fundamental theorem of algebra 2.34 also holds for quaternionic
polynomials.
100
3.4. The degree of a continuous map
Proof. Suppose the polynomial 𝑝 has no root. Then we can define the continuous map
𝑝ˆ : 𝑆3 → 𝑆3 with 𝑝(𝑧)
ˆ := | 𝑝𝑝 (𝑧)
(𝑧) 3 3 𝑝 (𝑡 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) = 𝑝(𝑧), ˆ hence we have 𝑝(𝑧)
ˆ ≃ const and conse-
quently deg( 𝑝)
ˆ = 0.
On the other hand, we can put for 𝑧 ∈ 𝑆3 and 𝑡 > 0
𝑡 𝑘 𝑝 𝑧𝑡
𝐺 (𝑧, 𝑡) := 𝑘 𝑧
|𝑡 𝑝 𝑡 |
and observe that the expression
𝑧
𝑡𝑘 𝑝 = 𝑧 𝑘 + 𝑡𝛼1 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) = 𝑝(𝑧).
ˆ Thus 𝑝ˆ ≃ 𝑓 𝑘 and hence deg( 𝑝)
ˆ = 𝑘. This contradicts 𝑘 > 0.
(𝑖) (𝑖𝑖)
Z 𝐻𝑛 (𝑆 𝑛 ; Z) / 𝐻𝑛 (𝑆 𝑛 , 𝑆 𝑛 \ 𝑓 −1 ( 𝑝); Z) o
𝐻𝑛 (𝑈, 𝑈 \ 𝑓 −1 ( 𝑝); Z) (3.7)
deg 𝑝 ( 𝑓 ) 𝐻𝑛 ( 𝑓 )
(𝑖𝑖𝑖)
Z 𝐻𝑛 (𝑆 𝑛 ; Z)
/ 𝐻𝑛 (𝑆 𝑛 , 𝑆 𝑛 \ {𝑝}; Z)
We observe:
1. Concerning (𝑖): This homomorphism is induced by the inclusion
𝑆 𝑛 = (𝑆 𝑛 , ∅) ↩→ (𝑆 𝑛 , 𝑆 𝑛 \ 𝑓 −1 ( 𝑝)).
101
3. Homology Theory
Definition 3.34. The number deg 𝑝 ( 𝑓 ) is called the local degree of 𝑓 over 𝑝.
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 (3.7) coincide and so do 𝐻𝑛 ( 𝑓 ) and (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 (3.7) are isomorphisms in this case, hence
multiplication by deg 𝑝 ( 𝑓 ) is an isomorphism, thus deg 𝑝 ( 𝑓 ) = ±1.
𝐻𝑛 (𝑆 𝑛 ) / 𝐻𝑛 (𝑆 𝑛 , 𝑆 𝑛 \ 𝐾) o 𝐻𝑛 (𝑉, 𝑉 \ 𝐾)
deg 𝑝 ( 𝑓 ) 𝐻𝑛 ( 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 )
◗◗◗ ❦❦❦
◗◗◗
◗◗◗ ❦❦❦❦❦❦❦
❦
❦❦❦❦ 𝐻𝑛 ( 𝑓 | 𝑉 )
◗◗◗
(
𝐻𝑛 (𝑆 𝑛 , 𝑆 𝑛 \ 𝐾) o 𝐻𝑛 (𝑉, 𝑉 \ 𝐾)
102
3.4. The degree of a continuous map
𝑓 (𝑥0 , 𝑥 )
′
if 𝑥0 ≥ 0,
′
′
′ ′ ′
𝐹 (𝑥0 , 𝑥 ) = 𝑥0 , k𝑥 k · 𝑓 (0, 𝑥 /k𝑥 k) if − 1 < 𝑥0 < 0,
(−1, 0)
if 𝑥0 = −1.
(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
𝐻𝑛 (𝐷 +𝑛 , 𝑆 𝑛−1 ) / 𝐻𝑛 (𝑆 𝑛 , 𝐷 𝑛 ) o
− 𝐻𝑛 (𝑆 𝑛 )
· deg( 𝑓 ) 𝐻𝑛 (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 (3.8) concludes the proof.
103
3. Homology Theory
Ê
𝐻𝑚 (𝑋) 𝐻𝑚 (𝑋𝑖 ),
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 ∩ 𝑋𝑖
Ê
𝐻𝑚 (𝑋, 𝐴) 𝐻𝑚 (𝑋𝑖 , 𝐴𝑖 ) .
𝑖
Proposition 3.41 (Additivity of the local degree). Let 𝑓 : 𝑈 → 𝑆 𝑛 be a continuous map. Let
𝑝 ∈ 𝑆 𝑛 be such that 𝑓 −1 ( 𝑝) is compact. Let 𝑈𝜆 ⊂ 𝑈 be open and put 𝑓𝜆 := 𝑓 |𝑈𝜆 , 𝜆 = 1, . . . , 𝑟.
Assume that 𝑓 −1 ( 𝑝) is the disjoint union of the 𝑓𝜆−1 ( 𝑝), i.e. 𝑓 −1 ( 𝑝) = ⊔𝑟𝜆=1 𝑓𝜆−1 ( 𝑝). Then
Õ
𝑟
deg 𝑝 ( 𝑓 ) = deg 𝑝 ( 𝑓𝜆 ) .
𝜆=1
104
3.4. The degree of a continuous map
𝑈
Example 3.42
Assume that 𝑓 −1 ( 𝑝) is a finite set, i.e.
b
Example 3.43. Consider the map 𝑓 𝑘 : 𝑆1 → 𝑆1 with 𝑓 (𝑧) = 𝑧 𝑘 , 𝑘 > 0, and set 𝑝 = 1. We write
and find that 𝑓 𝑘 | small neighborhood of 𝜉𝜆 is a homeomorphism. Hence deg( 𝑓 𝑘,𝜆 ) = ±1. Since 𝑘 > 0,
the restriction of 𝑓 𝑘 to a small neighborhood of 𝜉𝜆 is homotopic to an embedding of a (𝑘 times
larger) neighborhood of 𝜉𝜆 into 𝑆1 . Hence deg1 ( 𝑓 𝑘,𝜆 ) = 1. We conclude deg1 ( 𝑓 𝑘 ) = 𝑘.
Proof of Proposition 3.41. Choose open neighborhoods 𝑉𝜆 such that 𝑓𝜆−1 ( 𝑝) ⊂ 𝑉𝜆 ⊂ 𝑈𝜆 with
𝑉𝜆 ∩ 𝑉𝜇 = ∅ for 𝜆 ≠ 𝜇. Now put 𝑉 = ∪𝑟𝜆=1𝑉𝜆 . Proposition 3.36 tells us deg 𝑝 ( 𝑓 ) = deg 𝑝 ( 𝑓 | 𝑉 ).
The commutative diagram
deg 𝑝 ( 𝑓 | 𝑉 )
𝐻𝑛 ( 𝑓 | 𝑉 ) ,
𝐻𝑛 (𝑆 𝑛 ) / 𝐻𝑛 (𝑉, 𝑉 \ 𝑓 −1 ( 𝑝))
O
/ 𝐻𝑛 (𝑆 𝑛 , 𝑆 𝑛 \ {𝑝}) o
O
𝐻𝑛 (𝑆
O
𝑛)
1
© ª
.®
.®
.® (1,...,1) (1,...,1)
®
«1¬
É𝑟 É𝑟 ⊕𝜆 𝐻𝑛 ( 𝑓𝜆 ) É𝑟 É𝑟
𝜆=1 𝐻𝑛 (𝑆 𝑛 ) /
𝜆=1 𝐻𝑛 (𝑉𝜆 , 𝑉𝜆 \ 𝑓𝜆−1 ( 𝑝)) /
𝜆=1 𝐻𝑛 (𝑆 𝑛 , 𝑆 𝑛 \ {𝑝}) o 2 𝜆=1 𝐻𝑛 (𝑆 𝑛 )
©deg 𝑝 ( 𝑓1 ) ª
®
®
.. ®
. ®
®
«
deg 𝑝 ( 𝑓𝑟 ) ¬
105
3. Homology Theory
yields
© ª ©.ª
deg 𝑝 ( 𝑓1 ) 1
deg 𝑝 ( 𝑓 | 𝑉 ) = (1, . . . , 1) · ..
. ® ®
® · .. ® = deg 𝑝 ( 𝑓1 ) + . . . + deg 𝑝 ( 𝑓𝑟 ).
« deg 𝑝 ( 𝑓𝑟 ) ¬ «1¬
Before continuing with topological considerations we clarify some of the underlying algebra.
Throughout this section let 𝑅 be a commutative ring with unit element.
𝜕𝑛+1 𝜕𝑛
... / 𝐾𝑛+1 / 𝐾𝑛 / 𝐾𝑛−1 / ...
𝜕𝑛 ◦ 𝜕𝑛+1 = 0 (3.9)
𝑍 𝑛 𝐾∗ := {𝑥 ∈ 𝐾𝑛 | 𝜕𝑛 𝑥 = 0} = ker(𝜕𝑛 ) ,
106
3.5. Homological algebra
Definition 3.46. Given two complexes 𝐾∗ and 𝐾∗′ a chain map 𝜑∗ : 𝐾∗ → 𝐾∗′ is a sequence of
homomorphisms 𝜑 𝑛 : 𝐾𝑛 → 𝐾𝑛′ such that
𝜑𝑛+1
𝐾𝑛+1 / 𝐾′
𝑛+1
′
𝜕𝑛+1
𝜕𝑛+1
𝜑𝑛
𝐾𝑛 / 𝐾′
𝑛
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∗ ) = id 𝐻𝑛 𝐾∗ .
𝜑∗ 𝜓∗
Definition 3.48. A sequence · · · −→ 𝐾∗′ −→ 𝐾∗ −→ 𝐾∗′′ −→ · · · of chain maps is called
𝜑𝑛 𝜓𝑛
exact iff the sequence · · · −→ 𝐾𝑛′ −→ 𝐾𝑛 −→ 𝐾𝑛′′ −→ · · · is exact for every 𝑛 ∈ Z.
𝑖∗ 𝑝∗
Proposition 3.49. If 0 → 𝐾∗′ −→ 𝐾∗ −→ 𝐾∗′′ → 0 is an exact sequence of complexes then
𝐻𝑛 (𝑖∗ ) 𝐻𝑛 ( 𝑝∗ )
the sequence 𝐻𝑛 𝐾∗′ −→ 𝐻𝑛 𝐾∗ −→ 𝐻𝑛 𝐾∗′′ is exact for every 𝑛 ∈ Z.
0 = 𝐻𝑛 ( 𝑝 ∗ ◦ 𝑖 ∗ ) = 𝐻𝑛 ( 𝑝 ∗ ) ◦ 𝐻𝑛 (𝑖 ∗ ).
Therefore im 𝐻𝑛 (𝑖 ∗ ) ⊂ ker 𝐻𝑛 ( 𝑝 ∗ ).
𝑝 𝑛 (𝑧 − 𝜕𝑥) = 𝜕 ′′ 𝑥 ′′ − 𝜕 ′′ 𝑝 𝑛+1 𝑥 = 𝜕 ′′ 𝑥 ′′ − 𝜕 ′′ 𝑥 ′′ = 0 .
107
3. Homology Theory
By exactness of the complex there exists 𝑦 ′ ∈ 𝐾𝑛′ such that 𝑧 − 𝜕𝑥 = 𝑖 𝑛 𝑦 ′ . Now we get
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
as follows: Let 𝑧′′ ∈ 𝑍 𝑛 𝐾∗′′ represent an element [𝑧′′ ] ∈ 𝐻𝑛 𝐾∗′′ . Since 𝑝 𝑛 is surjective we can
choose 𝑥 ∈ 𝐾𝑛 with 𝑝 𝑛 𝑥 = 𝑧′′ . For 𝜕𝑥 ∈ 𝐾𝑛−1 we observe
𝑝 𝑛−1 𝜕𝑥 = 𝜕 ′′ 𝑝 𝑛 𝑥 = 𝜕 ′′ 𝑧′′ = 0.
𝜕∗ [𝑧′′ ] := [𝑦 ′ ].
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 𝑝 𝑛 𝑥 = 𝑧′′ and that of the representing cycle 𝑧′′ itself.
As to the choice of 𝑥, let 𝑝 𝑛 𝑥 = 𝑝 𝑛 𝑥¯ = 𝑧′′ . For the corresponding elements 𝑦 ′ , 𝑦¯ ′ ∈ 𝐾𝑛−1
′ we
′ ′
have and 𝑖 𝑛−1 𝑦 = 𝜕𝑥 and 𝑖 𝑛−1 𝑦¯ = 𝜕 𝑥. ¯ Since 𝑝 𝑛 (𝑥 − 𝑥)¯ = 0 there exists an 𝜔 ∈ 𝐾𝑛′ with
′
𝑥 − 𝑥¯ = 𝑖 𝑛 (𝜔′ ). We compute
𝑖 𝑛−1 (𝑦 ′ − 𝑦¯ ′ ) = 𝜕 (𝑥 − 𝑥)
¯ = 𝜕 (𝑖 𝑛 𝜔′ ) = 𝑖 𝑛−1 𝜕𝜔′ ,
108
3.5. Homological algebra
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 𝑝 𝑛+1 is onto we can choose 𝜉 ∈ 𝐾𝑛+1 with
𝑝 𝑛+1 𝜉 = 𝜁 ′′ . Then 𝑥 = 𝜕𝜉 is an admissible choice because
𝑝 𝑛 𝑥 = 𝑝 𝑛 𝜕𝜉 = 𝜕 ′′ 𝑝 𝑛+1 𝜉 = 𝜕 ′′ 𝜁 ′′ = 𝑧′′ .
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 (𝑖∗ )
··· / 𝐻𝑛+1 𝐾 ′′ 𝜕∗
/ 𝐻𝑛 𝐾 ′ / 𝐻𝑛 𝐾∗ 𝐻𝑛 ( 𝑝∗ )/ 𝐻𝑛 𝐾 ′′ 𝜕∗
/ 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
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 𝜕∗ .
c) im 𝐻 ( 𝑝) ⊂ ker 𝜕∗ :
This follows from
𝜕∗ 𝐻 ( 𝑝) [𝑧] = 𝜕∗ [ 𝑝𝑧] = [𝑖 −1 𝜕𝑧 ] = 0 .
|{z}
=0
d) ker 𝜕∗ ⊂ im 𝐻 ( 𝑝):
Let 𝜕∗ [𝑧′′ ] = 0. We write 𝑧′′ = 𝑝𝑥 and compute
109
3. Homology Theory
𝜕 (𝑥 − 𝑖𝑥 ′ ) = 𝜕𝑥 − 𝑖𝜕 ′ 𝑥 ′ = 0 .
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) / ...
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
∗ ∗
110
3.5. Homological algebra
𝐻𝑛+1 ( 𝑝∗ ) 𝐻𝑛 (𝑖∗ )
... / 𝐻𝑛+1 𝐾 ′′ 𝜕∗
/ 𝐻𝑛 𝐾 ′ / 𝐻𝑛 𝐾∗ 𝐻𝑛 ( 𝑝∗ )/ 𝐻𝑛 𝐾 ′′ 𝜕∗
/ ...
∗ ∗ ∗
is commutative as well.
𝜕∗
𝐻𝑛 𝐾 ′′ / 𝐻𝑛−1 𝐾 ′
𝐻𝑛 ( 𝜑 ′′ ) 𝐻𝑛−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] ,
Example 3.56. Let 𝑓 ∈ 𝐶 ((𝑋, 𝐴), (𝑌 , 𝐵)). For 𝐾𝑛′ = 𝑆𝑛 ( 𝐴), 𝐾𝑛 = 𝑆𝑛 (𝑋), 𝐾𝑛′′ = 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.
111
3. Homology Theory
Proposition 3.58. (i) The relation “≃” is an equivalence relation on the set of all chain
maps 𝐾 → 𝐾 . ′
𝜑′ 𝜑 − 𝜑′ 𝜓 = 𝜑′ (𝜕 ′ ℎ + ℎ𝜕) = 𝜕𝜑′ ℎ + 𝜑′ ℎ𝜕 ,
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
ℎ 𝜓
𝜑′ 𝜑 ≃ 𝜑′ 𝜓 ≃ 𝜓 ′ 𝜓 .
112
3.6. Proof of the homotopy axiom
b b b
(𝑒0 , 1)
𝑒0 𝑒1
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)
113
3. Homology Theory
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 ◦ 𝐹𝑛+1
𝑖
(1 ≤ 𝑖 ≤ 𝑛),
𝑇𝑛0 ◦ 𝐹𝑛+1
0
= 𝑖1 ,
𝑇𝑛𝑛 ◦ 𝐹𝑛+1
𝑛+1
= 𝑖0 ,
If 𝑖 ≤ 𝑘 ≤ 𝑗 then
𝑗+1 𝑖 𝑗+1 𝑗
𝑇𝑛 ◦ 𝐹𝑛+1 (𝑒 𝑘 ) = 𝑇𝑛 (𝑒 𝑘+1 ) = (𝑒 𝑘+1 , 0) = (𝐹𝑛𝑖 × id) (𝑒 𝑘 , 0) = (𝐹𝑛𝑖 × id) ◦ 𝑇𝑛−1 (𝑒 𝑘 ).
If 𝑘 > 𝑗 then
𝑗+1 𝑖 𝑗+1 𝑗
𝑇𝑛 ◦ 𝐹𝑛+1 (𝑒 𝑘 ) = 𝑇𝑛 (𝑒 𝑘+1 ) = (𝑒 𝑘 , 1) = (𝐹𝑛𝑖 × id) (𝑒 𝑘−1 , 1) = (𝐹𝑛𝑖 × id) ◦ 𝑇𝑛−1 (𝑒 𝑘 ).
The proofs of the other formulas are similar exercises in index shifting.
𝑗
Now let 𝜎 : Δ𝑛 → 𝑋 be a singular 𝑛-simplex. Note that (𝜎 × id) ◦ 𝑇𝑛 ∈ 𝐶 (Δ𝑛+1 , 𝑋 × 𝐼). We
define 𝑃𝜎 ∈ 𝑆𝑛+1 (𝑋 × 𝐼) by
Õ
𝑛
𝑗
𝑃𝜎 := (−1) 𝑗 (𝜎 × id) ◦ 𝑇𝑛 .
𝑗=0
114
3.6. Proof of the homotopy axiom
Lemma 3.63. Let (𝑋, 𝐴) be a pair of spaces. The operator 𝑃 is a chain homotopy for
Õ
0≤ 𝑗<𝑖≤𝑛
𝑗+1
− (−1) 𝑖+ 𝑗+1 (𝜎 × id) ◦ 𝑇𝑛 𝑖
◦ 𝐹𝑛+1 .
0≤𝑖≤ 𝑗 ≤𝑛−1
Õ
𝑗=0 𝑖=0
𝑗
= (−1) 𝑖+ 𝑗 (𝜎 × id) ◦ 𝑇𝑛 ◦ 𝐹𝑛+1
𝑖
(𝑖 < 𝑗)
0≤𝑖< 𝑗 ≤𝑛
Õ𝑛
+ (𝜎 × id) ◦ 𝑇𝑛𝑖 ◦ 𝐹𝑛+1
𝑖
(𝑖 = 𝑗)
𝑖=0
Õ
𝑛+1
− (𝜎 × id) ◦ 𝑇𝑛𝑖−1 ◦ 𝐹𝑛+1
𝑖
(𝑖 = 𝑗 + 1)
𝑖=1
Õ
𝑗
+ (−1) 𝑖+ 𝑗 (𝜎 × id) ◦ 𝑇𝑛 ◦ 𝐹𝑛+1
𝑖
(𝑖 > 𝑗 + 1)
1≤ 𝑗+1<𝑖≤𝑛+1
115
3. Homology Theory
Õ ′ +1 𝑗 ′ +1
= (−1) 𝑖+ 𝑗 (𝜎 × id) ◦ 𝑇𝑛 𝑖
◦ 𝐹𝑛+1
0≤𝑖≤ 𝑗 ′ ≤𝑛−1
+(𝜎 Õ
× id) ◦ 𝑖1 − (𝜎 × id) ◦ 𝑖0
′ 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.
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 ◦ 𝑖 0𝑋 .
Then by (3.10) we get
𝑆( 𝑓 ) − 𝑆(𝑔) = 𝑆(𝐹)𝑆(𝑖1𝑋 ) − 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).
𝐴 𝑈
116
3.7. Proof of the excision axiom
Í
Let 𝑋 be a topological space and let U = {𝑈𝑖 }𝑖∈ 𝐼 be an open cover of 𝑋. A chain
𝜎 = 𝑗 𝛼 𝑗 𝜎 𝑗 ∈ 𝑆𝑛 (𝑋) is called U-small iff for each 𝑗 there exists an 𝑖 such that 𝜎 𝑗 (Δ𝑛 ) ⊂ 𝑈𝑖 .
We denote
𝑆𝑛U (𝑋) := {𝜎 ∈ 𝑆𝑛 (𝑋) | 𝜎 is U-small}
!
Ê ⊕𝑖 𝑆𝑛 ( 𝑗𝑖 )
= im 𝑆𝑛 (𝑈𝑖 ) −−−−−−−−−→ 𝑆𝑛 (𝑋)
𝑖∈ 𝐼
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.
˚ Now set 𝑈1 := 𝐴˚
To this extent let (𝑋, 𝐴) be a pair of spaces and let 𝑈 ⊂ 𝐴 be such that 𝑈¯ ⊂ 𝐴.
¯
and 𝑈2 := 𝑋 \ 𝑈. Then U = {𝑈1 , 𝑈2 } forms an open cover of the space 𝑋. We compute
𝑆𝑛U (𝑋)
𝑆𝑛U (𝑋, 𝐴) =
𝑆𝑛U ( 𝐴)
˚ + 𝑆𝑛 (𝑋 \ 𝑈)
𝑆𝑛 ( 𝐴) ¯
=
˚ + 𝑆𝑛 ( 𝐴 \ 𝑈)
𝑆𝑛 ( 𝐴) ¯
¯
𝑆𝑛 (𝑋 \ 𝑈)
=
˚ + 𝑆𝑛 ( 𝐴 \ 𝑈))
(𝑆𝑛 ( 𝐴) ¯ ∩ 𝑆𝑛 (𝑋 \ 𝑈)
¯
¯
𝑆𝑛 (𝑋 \ 𝑈)
= .
¯
𝑆𝑛 ( 𝐴 \ 𝑈)
Similarly we get
𝑆𝑛 ( 𝐴˚ \ 𝑈) + 𝑆𝑛 (𝑋 \ 𝑈)
¯ ¯
𝑆𝑛 (𝑋 \ 𝑈)
𝑆𝑛U (𝑋 \ 𝑈, 𝐴 \ 𝑈) = = .
𝑆𝑛 ( 𝐴˚ \ 𝑈) + 𝑆𝑛 ( 𝐴 \ 𝑈)
¯ ¯
𝑆𝑛 ( 𝐴 \ 𝑈)
Thus
𝑆𝑛U (𝑋, 𝐴) = 𝑆𝑛U (𝑋 \ 𝑈, 𝐴 \ 𝑈).
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) / 𝑆 𝑛 (𝑋, 𝐴)
117
3. Homology Theory
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 Δ𝑛 by
1 Õ
𝑛
𝐵𝑛 := 𝑒𝑗.
𝑛 + 1 𝑗=0
𝐵1 𝐵2
𝐵0 = 𝑒 0 ,
1 𝑒0 𝑒1 𝑒0 𝑒1
𝐵1 = (𝑒0 + 𝑒1 ),
2
1 Figure 78. Barycenters in 1 and 2 di-
𝐵2 = (𝑒0 + 𝑒1 + 𝑒2 ). mensions
3
For an affine map 𝜎 : Δ𝑛 → Δ𝑛+1 we define the affine map 𝐶𝑛 𝜎 : Δ𝑛+1 → Δ𝑛+1 by
(
𝐵𝑛+1 , 𝑘 = 0,
(𝐶𝑛 𝜎) (𝑒 𝑘 ) =
𝜎 (𝑒 𝑘−1 ), 𝑘 ≥ 1.
Now we set
Õ
𝑛 𝑛
𝑆aff
𝑘 (Δ ) := 𝜎 ∈ 𝑆 𝑘 (Δ ) 𝜎 = 𝛼 𝑗 𝜎 𝑗 and each 𝜎 𝑗 is affine .
𝑗
𝑛+1 𝑛+1
𝐶𝑛 : 𝑆aff
𝑛 (Δ ) → 𝑆aff
𝑛+1 (Δ ).
118
3.7. Proof of the excision axiom
𝑒2
b
𝐺
b
b
𝑒0 𝑒1
𝑒0 𝑒1 𝐶1 𝐺
𝑒2
b
b b
𝑒0 𝑒1
Figure 79. 𝐶1
Proof. (i) It suffices to show the assertion for an affine simplex 𝑐. We then find
and hence
𝜕𝐶0 (𝑐) (𝑒0 ) = (𝐶0 (𝑐)) (𝑒1 ) − (𝐶0 (𝑐)) (𝑒0 ) = 𝑐(𝑒0 ) − 𝐵0 = (𝑐 − 1𝐵0 ) (𝑒0 )
as desired.
Lemma 3.70. To each topological space 𝑋 and each 𝑛 ∈ N0 we can associate homomorphisms
such that
119
3. Homology Theory
(iii) Sd∗ and 𝑄 ∗ are natural, i.e., for every 𝑓 ∈ 𝐶 (𝑋,𝑌 ) the following diagrams commute:
Sd𝑛 𝑄𝑛
𝑆𝑛 (𝑋) / 𝑆 𝑛 (𝑋) 𝑆𝑛 (𝑋) / 𝑆 𝑛+1 (𝑋)
𝑆𝑛 ( 𝑓 ) 𝑆𝑛 ( 𝑓 ) 𝑆𝑛 ( 𝑓 ) 𝑆𝑛+1 ( 𝑓 )
Sd𝑛 𝑄𝑛
𝑆𝑛 (𝑌 ) / 𝑆 𝑛 (𝑌 ) 𝑆𝑛 (𝑌 ) / 𝑆 𝑛+1 (𝑌 )
(iv) If the map 𝜎 : Δ𝑛 → Δ𝑛 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 𝑄 0 := 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 : Δ𝑛 → 𝑋:
| {z }
∈𝑆𝑛aff (Δ𝑛 )
aff (Δ𝑛 )
∈𝑆𝑛−1
| {z }
aff (Δ𝑛 )
∈𝑆𝑛−1
| {z }
∈𝑆𝑛aff (Δ𝑛 )
and
𝑄 𝑛 (𝜎) = 𝑆𝑛+1 (𝜎) (𝐶𝑛 (idΔ𝑛 −Sd𝑛 (idΔ𝑛 ) − 𝑄 𝑛−1 𝜕 idΔ𝑛 ) ).
| {z }
| {z }
∈𝑆𝑛aff (Δ𝑛 )
aff (Δ𝑛 )
∈𝑆𝑛+1
120
3.7. Proof of the excision axiom
Next we check (i): First we consider the case that 𝑋 = Δ𝑛 and 𝜎 = idΔ𝑛 .
𝜕 (Sd𝑛 (idΔ𝑛 )) = 𝜕 (𝐶𝑛−1 (Sd𝑛−1 (𝜕 idΔ𝑛 )))
L. 3.69
= Sd𝑛−1 (𝜕 idΔ𝑛 ) − 𝐶𝑛−2 𝜕 (Sd𝑛−1 (𝜕 idΔ𝑛 ))
ind. hyp.
|{z}
= Sd𝑛−1 (𝜕 idΔ𝑛 ) − 𝐶𝑛−2 Sd𝑛−2 𝜕𝜕 idΔ𝑛
=0
= Sd𝑛−1 (𝜕 idΔ𝑛 ).
For a general simplex 𝜎 : Δ𝑛 → 𝑋 we then find
𝜕 (Sd𝑛 (𝜎)) = 𝜕 (𝑆𝑛 (𝜎) (Sd𝑛 (idΔ𝑛 )))
= 𝑆𝑛 (𝜎) (𝜕 (Sd𝑛 (idΔ𝑛 )))
= 𝑆𝑛 (𝜎) (Sd𝑛−1 (𝜕 idΔ𝑛 ))
(𝑖𝑖𝑖)
= Sd𝑛−1 (𝑆𝑛 (𝜎) (𝜕 (idΔ𝑛 )))
= Sd𝑛−1 (𝜕𝑆𝑛 (𝜎) (idΔ𝑛 ))
= Sd𝑛−1 (𝜕𝜎).
Now we check (ii): Again we first consider the case 𝑋 = Δ𝑛 and 𝜎 = idΔ𝑛 .
𝜕𝑄 (idΔ𝑛 ) = 𝜕𝐶𝑛 (idΔ𝑛 −Sd𝑛 (idΔ𝑛 ) − 𝑄 𝑛−1 𝜕 idΔ𝑛 )
L. 3.69
= idΔ𝑛 −Sd𝑛 (idΔ𝑛 ) − 𝑄 𝑛−1 𝜕 idΔ𝑛
−𝐶𝑛−1 𝜕 idΔ𝑛 −Sd𝑛 (idΔ𝑛 ) − 𝑄 𝑛−1 𝜕 idΔ𝑛
ind. hyp.
= idΔ𝑛 −Sd𝑛 (idΔ𝑛 ) − 𝑄 𝑛−1 𝜕 idΔ𝑛
−𝐶𝑛−1 𝜕 idΔ𝑛 −𝜕Sd𝑛 (idΔ𝑛 ) − (𝜕 idΔ𝑛 −Sd𝑛−1 (𝜕 idΔ𝑛 )
|{z}
+ 𝑄 𝑛−2 ( 𝜕𝜕 idΔ𝑛 ))
=0
(𝑖)
= idΔ𝑛 −Sd𝑛 (idΔ𝑛 ) − 𝑄 𝑛−1 𝜕 idΔ𝑛 .
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
Figure 80. Vertices of maximal distance
121
3. Homology Theory
𝑛−1 𝑛
𝑑 ( 𝑝, 𝑞) ≤ diam(face of 𝜎) < diam(𝜎) .
𝑛 𝑛+1
2. One vertex is 𝐵𝑛 :
𝐵𝑛
𝑝𝑖
Figure 81. Barycenter is a vertex
Then we find
1 Õ
𝑛
𝑑 ( 𝑝 𝑖 , 𝐵𝑛 ) = 𝑝 𝑖 − 𝑝𝑗
𝑛 + 1 𝑗=0
1 Õ
𝑛
= ( 𝑝𝑖 − 𝑝 𝑗 )
𝑛 + 1 𝑗=0
1 Õ
𝑛
≤ | 𝑝𝑖 − 𝑝 𝑗 |
𝑛 + 1 𝑗=0 | {z }
≤diam( 𝜎) and =0 for 𝑗=𝑖
𝑛
≤ diam(𝜎) .
𝑛+1
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
122
3.7. Proof of the excision axiom
Lemma 3.72. For 𝜎 : Δ𝑛 → 𝑋 continuous there exists a 𝜀 > 0 such that all 𝜀-balls ∩Δ𝑛 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 𝑝 𝑘 ∈ Δ𝑛 such
that
𝐵 1 ( 𝑝 𝑘 ) = {𝑥 ∈ Δ𝑛 | |𝑥 − 𝑝 𝑘 | < 𝜀 𝑘 }
𝑘
is not contained in any of the 𝜎 −1 (𝑈𝑖 ). After passing to a subsequence we have that 𝑝 𝑘 → 𝑝 ∈ Δ𝑛
by compactness of Δ𝑛 . Choose 𝑖0 with 𝑝 ∈ 𝜎 −1 (𝑈𝑖0 ). Since 𝜎 −1 (𝑈𝑖0 ) is open there exists a 𝛿 > 0
such that 𝐵 𝛿 ( 𝑝) ⊂ 𝜎 −1 (𝑈𝑖0 ). Now choose 𝑘 so large, that | 𝑝 𝑘 − 𝑝| < 𝛿/2 and 𝜀 𝑘 = 𝑘1 < 𝛿/2.
We then find
𝐵 1 ( 𝑝 𝑘 ) ⊂ 𝐵 𝛿 ( 𝑝) ⊂ 𝜎 −1 (𝑈𝑖0 ),
𝑘
a contradiction.
Corollary 3.73. Assume that 𝜎 : Δ𝑛 → 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).
Sd𝑟 𝑥 = Sd𝑟 𝜕𝑥
|{z}
𝜕
U (𝑋) for large 𝑟
∈𝑆𝑛+1
= Sd𝑟 𝑧
= 𝑧 − 𝜕𝑄 (𝑟 ) 𝑧 − 𝑄 (𝑟 ) 𝜕𝑧 .
|{z}
=0
Hence
𝑧 = 𝜕 ( Sd𝑟 𝑥 + 𝑄 (𝑟 ) 𝑧 )
| {z }
∈𝑆𝑛U (𝑋) for large 𝑟
123
3. Homology Theory
ii) We show surjectivity: Let [𝑧] ∈ 𝐻𝑛 (𝑋) be given. We know that Sd𝑟 𝑧 ∈ 𝑆𝑛U (𝑋) for 𝑟 large
enough. We compute
b) Now we pass to general (𝑋, 𝐴). Consider the commutative diagram with exact rows:
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.
with the inclusion maps 𝑖 𝜈 : 𝑈1 ∩ 𝑈2 → 𝑈𝜈 and 𝑗 𝜈 : 𝑈𝜈 → 𝑋. We then get the following long
exact homology sequence:
𝐻𝑛 (𝑖1 )
· · · → 𝐻𝑛 (𝑈1 ∩ 𝑈2 )
−𝐻𝑛 (𝑖2 )
/ 𝐻𝑛 (𝑈1 ) ⊕ 𝐻𝑛 (𝑈2 ) (𝐻𝑛 ( 𝑗1 ),𝐻𝑛 ( 𝑗2 ) ) / 𝐻 U (𝑋) →
𝜕
𝐻𝑛−1 (𝑈1 ∩ 𝑈2 ) → · · ·
𝑛
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).
(𝐻𝑛 ( 𝑗1 ),𝐻𝑛 ( 𝑗2 ) ) 𝜕
· · · → 𝐻𝑛 (𝑈1 ) ⊕ 𝐻❱𝑛 (𝑈2 ) / 𝐻 U (𝑋) / 𝐻𝑛−1 (𝑈1 ∩ 𝑈2 ) → · · ·
𝑛
❱❱❱❱ ✐✐4
❱❱❱❱
❱❱❱❱ ✐✐✐ ✐✐✐✐
✐✐
(𝐻𝑛 ( 𝑗1 ),𝐻𝑛 ( 𝑗2 ) )❱❱❱❱❱❱ ✐✐✐✐ 𝜕 𝑀𝑉
❱* ✐✐✐✐
𝐻𝑛 (𝑋)
124
3.8. The Mayer-Vietoris sequence
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
𝑖 𝜈 :(𝑈1 ∩ 𝑈2 , 𝐴1 ∩ 𝐴2 ) → (𝑈𝜈 , 𝐴𝜈 ),
𝑗 𝜈 :(𝑈𝜈 , 𝐴𝜈 ) → (𝑋, 𝐴),
be the inclusion maps, 𝜈 = 1, 2. Then the following sequence is exact and natural
𝐻𝑛 (𝑖1 )
𝜕 𝑀𝑉 / 𝐻𝑛 (𝑈1 ∩ 𝑈2 , 𝐴1 ∩ 𝐴2 ) −𝐻𝑛 (𝑖2 )
/ 𝐻𝑛 (𝑈1 , 𝐴1 ) ⊕ 𝐻𝑛 (𝑈2 , 𝐴2 )
···
❝ ❝ ❝ ❝ ❝❝❝❝
❝❝❝❝❝❝❝❝
❝❝❝❝❝❝❝ ❝❝❝❝❝❝❝❝(𝐻𝑛 ( 𝑗1 ),𝐻𝑛 ( 𝑗2 ) )
❝❝❝
q ❝❝❝❝❝❝❝
❝
𝐻𝑛 (𝑋, 𝐴) 𝑀𝑉 / 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
and
𝑈1 ∩ 𝑈2 ≈ (0, 1) ⊔ (0, 1) ≃ {𝑝 1 , 𝑝 2 }
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
11
0 / 𝐻1 (𝑆 1 ) / 𝑅2 / 𝑅2
125
3. Homology Theory
𝑈1 𝑈2
0 / 𝐻𝑛 (𝐺 2 ) /0
𝐻0 ({𝑝}) 𝑅 /𝐻 (𝑆 1) ⊕: 𝐻0 (𝑆1 ) 𝑅 2
0
1
1
𝐻1 (𝐺 2 ) 𝐻1 (𝑈1 ) ⊕ 𝐻1 (𝑈2 ) 𝑅 2 .
126
3.9. Generalized Jordan curve theorem
Lemma 3.77. Let 𝑋 be a topological space and let 𝑈𝑖 ⊂ 𝑋 be open with 𝑈𝑖 ⊂ 𝑈𝑖+1 and
∪𝑖∈N𝑈𝑖 = 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 ∈ 𝐻 𝑘 (𝑋; 𝑅) there exists an 𝑛0 such that 𝛼 ∈ im(𝐻 𝑘 (𝜄𝑛 )) for all 𝑛 ≥ 𝑛0 .
(ii) For each 𝛼𝑛 ∈ 𝐻 𝑘 (𝑈𝑛 ; 𝑅) with 𝐻 𝑘 (𝜄𝑛 ) (𝛼𝑛 ) = 0 there exists an 𝑚 0 such that
𝐻 𝑘 (𝜄𝑛,𝑚 ) (𝛼𝑛 ) = 0 for all 𝑚 ≥ 𝑚 0 .
Õ
𝑙
𝛼 𝑗 𝜎 𝑗 ∈ 𝑍 𝑘 (𝑋; 𝑅), 𝛼 𝑗 ∈ 𝑅, 𝜎 𝑗 ∈ 𝐶 (Δ𝑘 , 𝑋).
𝑗=1
Ð
Note that 𝜎 𝑗 (Δ𝑘 ) ⊂ 𝑋 is a compact subset and therefore 𝐶 := 𝑙𝑗=1 𝜎 𝑗 (Δ𝑘 ) ⊂ 𝑋 is
compact. Hence there exists an 𝑛0 such that for all 𝑛 ≥ 𝑛0 we have 𝐶 ⊂ 𝑈𝑛 . We conclude
that
Õ 𝑙
𝛼 𝑗 𝜎 𝑗 ∈ 𝑍 𝑘 (𝑈𝑛 ; 𝑅)
𝑗=1
Í
(ii) Again represent 𝛼𝑛 ∈ 𝐻 𝑘 (𝑈𝑛 ) by 𝑙𝑗=1 𝛼 𝑗 𝜎 𝑗 . From 𝐻𝑛 (𝜄𝑛 ) (𝛼𝑛 ) = 0 we know that there
Í
exists a 𝛽 ∈ 𝐶 𝑘+1 (𝑋; 𝑅) such that 𝑙𝑗=1 𝛼 𝑗 𝜎 𝑗 = 𝜕 𝛽. As before there exists a compact
subset 𝐶 ′ ⊂ 𝑋 such that 𝛽 ∈ 𝐶 𝑘+1 (𝐶 ′ ; 𝑅). Since there exists an 𝑚 0 with 𝐶 ′ ⊂ 𝑈𝑚 for all
𝑚 ≥ 𝑚 0 and thus 𝛽 ∈ 𝐶 𝑘+1 (𝑈𝑚 ; 𝑅) we have that 𝐻 𝑘 (𝜄𝑛,𝑚 ) (𝛼𝑛 ) = 0.
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) .
Then the space [0, 1] × 𝑌 also has this property.
127
3. Homology Theory
𝑈0 := 𝑆 𝑛 \ 𝑓 ([0, 21 ] × 𝑌 ) and 𝑈1 := 𝑆 𝑛 \ 𝑓 ([ 12 , 1] × 𝑌 ) .
| {z } | {z }
compact compact
𝑈0 ∩ 𝑈1 = 𝑆 𝑛 \ 𝑓 ([0, 1] × 𝑌 ) and 𝑈0 ∪ 𝑈1 = 𝑆 𝑛 \ 𝑓 ({ 12 } × 𝑌 ) .
| {z }
≈𝑌
0 = 𝐻𝑖+1 (𝑆 𝑛 \ 𝑓 ({ 12 } × 𝑌 )) / 𝐻𝑖 (𝑆 𝑛 \ 𝑓 ([0, 1] × 𝑌 ))
𝐻𝑖 (𝑆 𝑛 \ 𝑓 ([0, 12 ] × 𝑌 )) ⊕ 𝐻𝑖 (𝑆 𝑛 \ 𝑓 ([ 21 , 1] × 𝑌 ))
𝐻𝑖 (𝑆 𝑛 \ 𝑓 ([0, 1] × 𝑌 )) → 𝐻𝑖 (𝑆 𝑛 \ 𝑓 ([0, 21 ] × 𝑌 ) ⊕ 𝐻𝑖 (𝑆 𝑛 \ 𝑓 ([ 12 , 1] × 𝑌 )) .
Hence
and
0 ≠ 𝐻𝑖 (𝜄0,𝑘 ) (𝛼) ∈ 𝐻𝑖 (𝑆 𝑛 \ 𝑓 (𝐼 𝑘 × 𝑌 ))
for all 𝑘. Here 𝑉𝑘 := 𝑆 𝑛 \ 𝑓 (𝐼 𝑘 × 𝑌 ) is open in 𝑆 𝑛 and 𝜄 𝑘,𝑙 : 𝑉𝑘 ↩→ 𝑉𝑙 for 𝑙 ≥ 𝑘 denotes the
inclusion map. We find
∩𝑘 ∈N 𝐼 𝑘 = {𝑡}
∪𝑘 ∈N𝑉𝑘 = 𝑆 𝑛 \ 𝑓 ({𝑡} × 𝑌 ) =: 𝑋
Now we apply the previous Lemma 3.77 for the inclusion 𝜄 : 𝑉0 ↩→ 𝑋. Hence, for 𝑖 ≥ 1,
giving a contradiction. The proof for 𝑖 = 0 is similar (or in fact the same if one uses augmented
homology).
128
3.9. Generalized Jordan curve theorem
𝐻∗ (𝑆 𝑛 \ 𝑓 (𝐷 𝑟 )) 𝐻∗ ({point})
𝑆 𝑛 \ 𝑓 (𝐷 0 ) ≈ R𝑛 ≃ {point}
𝐻∗ (𝑆 𝑛 \ 𝑓 (𝑆𝑟 )) 𝐻∗ (𝑆 𝑛−𝑟 −1 ).
𝐻𝑖 (𝑈+ ) ⊕ 𝐻𝑖 (𝑈 − )
| {z }
𝐻𝑖 (point) ⊕𝐻𝑖 (point) =0
if 𝑖 ≥1
129
3. Homology Theory
Proof. a) We know that 𝐻0 (𝑆 𝑛 \ 𝑓 (𝑆 𝑛−1 )) 𝐻0 (𝑆0 ) 𝑅 2 , 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 ).
𝑆 𝑛 = R𝑛 ∪ {∞}, 𝑓 : 𝑆 𝑛−1 → R𝑛 ⊂ 𝑆 𝑛 .
130
3.10. CW-complexes
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 ≈ 𝐷˚ 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.
Ý
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:
Ý
(i) 𝑋 = 𝜎 ∈ X 𝜎.
𝜑 𝜎 : 𝐷𝑛 → 𝜎
¯ ⊂𝑋
131
3. Homology Theory
Definition 3.85. An element 𝜎 ∈ X𝑛 is called an n-cell. The map 𝜑 𝜎 is called the characteris-
tic map of 𝜎 and 𝑋 𝑛 is called the n-skeleton of (𝑋, X). The map 𝜑 𝜎 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
𝜑 𝜎𝑛
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 = . . .
132
3.10. CW-complexes
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 }}, 𝑚 = 0
X𝑚 =
∅, 𝑚 > 0.
𝑛−1
X𝑆 ,
𝑚
𝑚 ≤ 𝑛−1
𝑛
X𝑚𝑆 := { 𝐷˚ +𝑛 , 𝐷˚ 𝑛− },
𝑚=𝑛
∅,
𝑚>𝑛
𝑆𝑛
◦+
𝐷𝑛
𝑆 𝑛−1
◦−
𝐷𝑛
𝑋 0 = 𝑆0 , 𝑋 1 = 𝑆1 , . . . , 𝑋 𝑛 = 𝑆 𝑛 .
Example 3.88. Real projective space RP𝑛 . We define the real projective space as
where
𝑥 = (𝑥0 , . . . , 𝑥 𝑛 ) ∼ 𝑦 = (𝑦 0 , . . . , 𝑦 𝑛 )
iff there exists a 𝑡 ≠ 0 such that 𝑥 = 𝑡 𝑦. We consider the canonical projection map
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
133
3. Homology Theory
Then
¯ 𝑘 = {[𝑥0 , . . . , 𝑥 𝑛 ] ∈ RP𝑛 | 𝑥 𝑘+1 = . . . = 𝑥 𝑛 = 0} ≈ RP𝑘 .
𝜎
For the characteristic map 𝜑 𝑘 : 𝐷 𝑘 → 𝜎 ¯ 𝑘 we can take
q
𝜉 ↦→ [𝜉1 , . . . , 𝜉 𝑘 , 1 − |𝜉 | 2 , 0, . . . , 0] .
It is clear that the map 𝜑 𝑘 is continuous. Now we check that it is also surjective. Let
[𝑥 ′ , 𝑥 𝑘 , 0] ∈ 𝜎
¯ 𝑘 where 𝑥 ′ = (𝑥0 , . . . , 𝑥 𝑘−1 ). Without loss of generality we assume that 𝑥 𝑘 ≥ 0.
′
Set 𝜉 := √ ′ 𝑥2 2
∈ 𝐷 𝑘 . We compute
| 𝑥 | +| 𝑥 𝑘 |
" s #
𝑥′ |𝑥 ′ | 2
𝜑 𝑘 (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 𝜑 𝑘 | 𝐷˚ 𝑘 is injective. Let 𝜑 𝑘 (𝜉) = 𝜑 𝑘 (𝜂) for 𝜉, 𝜂 ∈ 𝐷˚ 𝑘 . This leads to the two
equations: q q
𝜉 = 𝑡𝜂 and 1 − |𝜉 | = 𝑡 1 − |𝜂| 2
2
|𝜉 | 2 + 1 − |𝜉 | 2 = 𝑡 2 |𝜂| 2 + 𝑡 2 (1 − |𝜂| 2 )
p
p
leading 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 𝜑 𝑘 : 𝐷 𝑘 → 𝜎 ¯ 𝑘 is closed, therefore the restriction
𝜑𝑘 ˚𝑘
𝐷
: 𝐷˚ → 𝜎𝑘
𝑋 0 = {point} ⊂ 𝑋 1 ⊂ 𝑋 2 ⊂ · · · ⊂ 𝑋 𝑛 .
|{z} |{z} |{z}
≈RP1 ≈RP2 =RP𝑛
Finally let us discuss the gluing map. For 𝜉 ∈ 𝜕𝐷 𝑘 = 𝑆 𝑘−1 , i.e. |𝜉 | = 1, the gluing map is given
by 𝜑 𝑘 (𝜉) = [𝜉, 0, 0] and hence 𝜑 𝑘 = 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,
134
3.11. Homology of CW-complexes
and
𝑋 0 = {point} = 𝑋 1 ⊂ 𝑋 2 = 𝑋 3 ⊂ 𝑋 4 = 𝑋 5 ⊂ · · · ⊂ 𝑋 2𝑛−1 = 𝑋 2𝑛 .
|{z} |{z} |{z}
≈CP1 ≈CP2 =CP𝑛
Example 3.90. Quaternionic projective space HP𝑛 . Similarly, for the quaternionic projective
space
and
𝑋 0 = {point} = 𝑋 1 = 𝑋 2 = 𝑋 3 ⊂ 𝑋 4 = . . . ⊂ 𝑋 4𝑛−3 = · · · = 𝑋 4𝑛 .
|{z} |{z}
≈HP1 =HP𝑛
Remark 3.91. Every compact differentiable manifold can be triangulated and is consequently a
finite CW-complex.
É ⊕ 𝜎∈X𝑛 𝐻𝑖 ( 𝜑 𝜎 )
𝐻𝑖 (𝐷 𝑛 , 𝑆 𝑛−1 ) / 𝐻𝑖 (𝑋 𝑛 , 𝑋 𝑛−1 )
𝜎 ∈ X𝑛
is an isomorphism.
135
3. Homology Theory
𝑥
The inclusion 𝑆 𝑛−1 ↩→ 𝐷¤ 𝑛 is a homotopy equivalence with homotopy inverse 𝑥 ↦→ |𝑥| .
𝐷𝑛
We define Þ Þ Þ
𝑌 𝑛 := 𝐷𝑛, 𝑌 𝑛−1 := 𝑆 𝑛−1 , 𝑌¤ 𝑛 := 𝐷¤ 𝑛
𝜎 ∈ X𝑛 𝜎 ∈ X𝑛 𝜎 ∈ X𝑛
𝑋1
Now consider:
Φ:=∪ 𝜎∈X𝑛 𝜑 𝜎
(𝑌 𝑛 , 𝑌 𝑛−1 ) / (𝑌 𝑛 , 𝑌¤ 𝑛 ) / (𝑋 𝑛 , 𝑋¤ 𝑛 ) o ? _ (𝑋 𝑛 , 𝑋 𝑛−1 )
136
3.11. Homology of CW-complexes
= =
𝐻𝑖 (𝑌¤ 𝑛 ) / 𝐻𝑖 (𝑌 𝑛 ) / 𝐻𝑖 (𝑌 𝑛 , 𝑌¤ 𝑛 ) / 𝐻𝑖−1 (𝑌¤ 𝑛 ) / 𝐻𝑖−1 (𝑌 𝑛 )
and the Five Lemma the map 𝐻𝑖 (𝑌 𝑛 , 𝑌 𝑛−1 ) → 𝐻𝑖 (𝑌 𝑛 , 𝑌¤ 𝑛 ) is also an isomorphism. Similarly,
we get that the inclusion (𝑋 𝑛 , 𝑋 𝑛−1 ) ↩→ (𝑋 𝑛 , 𝑋¤ 𝑛 ) induces an isomorphism 𝐻𝑖 (𝑋 𝑛 , 𝑋 𝑛−1 ) →
𝐻𝑖 (𝑋 𝑛 , 𝑋¤ 𝑛 ). The inclusions
(𝑌 𝑛 \ 𝑌 𝑛−1 , 𝑌¤ 𝑛 \ 𝑌 𝑛−1 ) ↩→ (𝑌 𝑛 , 𝑌¤ 𝑛 ),
(𝑋 𝑛 \ 𝑋 𝑛−1 , 𝑋¤ 𝑛 \ 𝑋 𝑛−1 ) ↩→ (𝑋 𝑛 , 𝑋¤ 𝑛 ),
Hence we find
Ê
𝐻𝑖 (𝐷 𝑛 , 𝑆 𝑛−1 ) 𝐻𝑖 (𝑌 𝑛 , 𝑌 𝑛−1 )
𝜎 ∈ X𝑛
𝐻𝑖 (𝑌 𝑛 , 𝑌¤ 𝑛 )
𝐻𝑖 (𝑌 𝑛 \ 𝑌 𝑛−1 , 𝑌¤ 𝑛 \ 𝑌 𝑛−1 )
𝐻𝑖 (𝑋 𝑛 \ 𝑋 𝑛−1 , 𝑋¤ 𝑛 \ 𝑋 𝑛−1 )
𝐻𝑖 (𝑋 𝑛 , 𝑋¤ 𝑛 )
𝐻𝑖 (𝑋 𝑛 , 𝑋 𝑛−1 ) .
Lemma 3.94. The sequence of groups 𝐾𝑛 (𝑋, X) together with 𝜕𝑛 forms a complex.
The pair (𝐾∗ (𝑋, X), 𝜕∗ ) is called the cellular complex of (𝑋, X).
137
3. Homology Theory
= 𝜕𝑛
*
𝜕
𝐻𝑛−1 (𝑋 𝑛−1 ) / 𝐻𝑛−1 (𝑋 𝑛−1 , 𝑋 𝑛−2 ) /
2 𝐻𝑛−1 (𝑋
𝑛−2 )
=0 = 𝜕𝑛−1
*
𝐻𝑛−1 (𝑋 𝑛−2 ) / 𝐻𝑛−2 (𝑋 𝑛−2 , 𝑋 𝑛−3 )
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.
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.
𝐻𝑛 (𝑋) 𝐻𝑛 (𝑋 𝑟 ) .
138
3.11. Homology of CW-complexes
Proof. The assertion follows from Corollary 3.97 and the exact sequence
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 𝐾∗ (𝑋, X) 𝐻𝑛 (𝑋)
Proof. Consider the commutative diagram with exact columns and rows
𝐻𝑛+1 (𝑋 𝑛+1 , 𝑋
❘
𝑛) 𝐻𝑛−1 (𝑋 𝑛−2 ) = 0
❘❘❘
❘❘❘𝜕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 )
139
3. Homology Theory
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)
𝑅 ←− 0 ←− · · · ←− 0 ←− 𝑅 ←− 0 ←− · · ·
Since all arrows are 0 the homology coincides with the complex, hence
(
𝑛 𝑅, 𝑗 = 0 or 𝑛,
𝐻 𝑗 (𝑆 )
0, otherwise,
Example 3.102. Now look at 𝑋 = CP𝑛 . Then we have for the CW-decomposition from Exam-
ple 3.89
𝛼0 = 𝛼2 = . . . = 𝛼2𝑛 = 1, 𝛼 𝑗 = 0 otherwise.
Again all arrows in the complex
𝑅 ←− 0 ←− · · · ←− 0 ←− 𝑅 ←− 0 ←− · · ·
Example 3.103. Consider 𝑋 = HP𝑛 . For the CW-decomposition from Example 3.90 we have
𝛼0 = 𝛼4 = 𝛼8 = . . . = 𝛼4𝑛 = 1, 𝛼 𝑗 = 0 otherwise
0 ←− 𝑅 ←− 0 ←− 0 ←− 0 ←− 𝑅 ←− · · · ←− 0 ←− 𝑅 ←− 0 ←− · · ·
140
3.12. Betti numbers and the Euler number
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) < ∞.
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:
Õ
∞
𝜒(𝑋, X) = (−1) 𝑖 𝑏𝑖 (𝑋; 𝑅).
𝑖=0
141
3. Homology Theory
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 𝛼𝑖 ’s don’t. We will henceforth write 𝜒(𝑋) instead of 𝜒(𝑋, X).
In order to prove the proposition we use the following
Proof of Lemma 3.107. To show the lemma we compute, using the dimension formula from
linear algebra,
Õ Õ
(−1) 𝑗 dim 𝑉 𝑗 = (−1) 𝑗 (dim(𝑑 𝑗 𝑉 𝑗 ) + dim ker(𝑑 𝑗 ))
Õ
𝑗 𝑗
= (−1) 𝑗 dim 𝐻 𝑗 𝑉∗ .
𝑗
Proof of Proposition 3.106. This follows from Lemma 3.107 with 𝑉 𝑗 = 𝐾 𝑗 (𝑋, X) 𝑅 𝛼 𝑗 .
𝑏0 = 𝑏 𝑛 = 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:
142
3.13. Incidence numbers
𝛼0 𝛼1 𝛼2
Tetrahedron 4 6 4
Cube 8 12 6
Octahedron 6 12 8
Dodecahedron 20 30 12
Icosahedron 12 30 20
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
𝛼0 = . . . = 𝛼𝑛 = 1, 𝛼 𝑗 = 0 otherwise.
Hence (
1, 𝑛 even,
𝜒(RP𝑛 ) =
0, 𝑛 odd.
𝜕
𝐻1 (𝑋 1 , 𝑋 0 ) / 𝐻0 (𝑋 0 )
É É
𝐻1 (𝜑 𝜏 ) (𝐻1 (𝐷 1 , 𝑆0 )) 𝑅
𝜏 ∈ X1 𝜎 ∈ X0
143
3. Homology Theory
Hence 𝜕1 is given by the (𝛼0 × 𝛼1 )-matrix (𝜕𝜎𝜏 ) where 𝜏 ∈ X1 and 𝜎 ∈ X0 . The entries 𝜕𝜎𝜏 ∈ 𝑅
of this matrix are easily computed. Namely, recall that a generator of 𝐻1 (𝐷 1 , 𝑆0 ) is represented
by 𝑐 : Δ1 = [0, 1] → 𝐷 1 = [−1, 1] with 𝑐(𝑡) = 2𝑡 − 1. Then
Thus if 𝜑 𝜏 (−1) = 𝜑 𝜏 (1) then 𝜕𝜎𝜏 = 0 for all 𝜎. If 𝜑 𝜏 (−1) ≠ 𝜑 𝜏 (1) then
1, for 𝜎 = 𝜑 𝜏 (1),
𝜏
𝜕𝜎 = −1, for 𝜎 = 𝜑 𝜏 (−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
Figure 91. A cell decomposition of the figure 8
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,
or possibly different signs which will not affect the homology however. Thus the cellular complex
is
©0 1 1ª
®
0 −1 −1¬
··· o 0o 𝑅 2 «o 𝑅3 o 0o ···
The image of the matrix is {(𝑥, −𝑥) | 𝑥 ∈ 𝑅} = 𝑅 · (1, −1) and hence
We summarize
𝑅 if 𝑗 = 0,
𝐻 𝑗 (𝑋; 𝑅) 𝑅 2
if 𝑗 = 1,
0
else.
144
3.13. Incidence numbers
𝜕𝑛+1
𝐾𝑛+1 (𝑋, X) / 𝐾𝑛 (𝑋, X)
É É
𝐻𝑛+1 (𝜑 𝜏 ) (𝐻𝑛+1 (𝐷 𝑛+1 , 𝑆 𝑛 )) 𝐻𝑛 (𝜑 𝜎 ) (𝐻𝑛 (𝐷 𝑛 , 𝑆 𝑛−1 ))
𝜏 ∈ X𝑛+1 𝜎 ∈ X𝑛
𝜏 )
(𝜕𝜎
𝑅 𝛼𝑛+1 / 𝑅 𝛼𝑛
Hence 𝜕𝑛+1 is given by the (𝛼𝑛 × 𝛼𝑛+1 )-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 ∈ 𝐷˚ 𝑛 . From the commutative diagram
𝜕𝑛+1
* pr 𝜎
𝜕
𝐻𝑛+1 (𝑋 𝑛+1
O
, 𝑋 𝑛) / 𝐻𝑛 (𝑋 𝑛 )
O
/ 𝐻𝑛 (𝑋 𝑛 , 𝑋 𝑛−1 ) / 𝐻𝑛 (𝜑 𝜎 )𝐻𝑛 (𝐷 𝑛 , 𝑆 𝑛−1 )
O
𝐻𝑛 ( 𝜑 𝜎 )
𝐻𝑛+1 ( 𝜑 𝜏 ) 𝐻𝑛 ( 𝜑 𝜏 | 𝑆 𝑛 ) 𝐻𝑛 (𝐷 𝑛O , 𝑆 𝑛−1 )
deg 𝑝 ( 𝜑 −1 −1 𝑛 𝑛
𝜕 𝜎 ◦𝜑 𝜏 :𝜑 𝜏 ( 𝜎)→𝐷 ⊂𝑆 )
𝐻𝑛+1 (𝐷 𝑛+1 , 𝑆 𝑛 )
/ 𝐻𝑛 (𝑆 𝑛 ) / 𝐻𝑛 (𝑆 𝑛 )
O
𝐻𝑛 ( 𝜑 −1
𝜎 ◦𝜑 𝜏 )
𝐻𝑛 (𝑆 𝑛 , 𝑆 𝑛 \ 𝜑 −1
𝜏 (𝜑 𝜎 ( 𝑝)))
o 𝐻𝑛 (𝜑 −1 −1 −1
𝜏 (𝜎), 𝜑 𝜏 (𝜎) \ 𝜑 𝜏 (𝜑 𝜎 ( 𝑝)))
we conclude that
𝜕𝜎𝜏 = deg 𝑝 (𝜑 −1 −1 𝑛 𝑛
𝜎 ◦ 𝜑 𝜏 | 𝜑 𝜏−1 ( 𝜎) : 𝜑 𝜏 (𝜎) → 𝐷 ⊂ 𝑆 ).
𝜕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 ∈ X 𝑗 and 𝜏 ∈ X 𝑗+1 . The gluing map
𝜑 𝜏 | 𝑆 𝑗 : 𝑆 𝑗 → 𝑋 𝑗 ≈ RP 𝑗 is given by the canonical projection. The image point 𝜑 𝜎 ( 𝑝) ∈ 𝑋 𝑗 has
145
3. Homology Theory
146
3.14. Homotopy versus homology
Proposition 3.112. Let (𝑋, X) be a finite CW-complex and let 𝑥0 ∈ 𝑋 0 . Then the inclusion
map 𝑗 : 𝑋 2 ↩→ 𝑋 induces an isomorphism
𝑗 # : 𝜋1 (𝑋 2 ; 𝑥0 ) → 𝜋1 (𝑋; 𝑥0 ) .
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 𝑥0 .
Let 𝑌 be a path-connected finite CW-complex and let 𝑌˜ be obtained from 𝑌 by attaching a 𝑘-cell.
More precisely, 𝑌˜ = 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
˜ 𝑥0 )
𝑗 # : 𝜋1 (𝑌 ; 𝑥0 ) → 𝜋(𝑌;
if 𝑘 ≥ 3. Let 𝐷 𝑘 ( 21 ) ⊂ 𝐷˚ 𝑘 be the closed 𝑘-dimensional subball of radius 12 . Cover 𝑌˜ by the two
open subsets 𝑈1 = 𝐷˚ 𝑘 and 𝑈2 = 𝑌˜ \ 𝐷 𝑘 ( 21 ) ≃ 𝑌 .
𝑌 𝑈2
147
3. Homology Theory
Example 3.113. Consider complex-projective space 𝑋 = CP𝑛 . We use the cell decomposition
from Example 3.89:
𝑋 0 = {point} = 𝑋 1 ⊂ 𝑋 2 = 𝑋 3 ⊂ 𝑋 4 = 𝑋 5 ⊂ . . . ⊂ 𝑋 2𝑛−1 ⊂ 𝑋 2𝑛
|{z} |{z} |{z}
≈CP1 ≈CP2 =CP𝑛
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 ⊂ 𝑋 4 = . . . ⊂ 𝑋 4𝑛−3 = . . . ⊂ 𝑋 4𝑛
|{z} |{z}
≈HP1 =HP𝑛
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# : 𝜋 𝑛 (𝑋 𝑛+1 ; 𝑥0 ) → 𝜋 𝑛 (𝑋; 𝑥0 )
where 𝑥0 ∈ 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 𝛼𝑛 ∈ 𝐻𝑛 (𝑊 𝑛 , 𝜕𝑊 𝑛 ; Z)
𝐻𝑛 (𝐷 𝑛 , 𝑆 𝑛−1 ; Z) Z. The elements of 𝜋 𝑛 (𝑋; 𝑥0 ) are homotopy classes relative to 𝜕𝑊 𝑛 of
maps 𝑓 : 𝑊 𝑛 → 𝑋 with 𝑓 (𝜕𝑊 𝑛 ) = {𝑥0 }. Then 𝐻𝑛 ( 𝑓 ) (𝛼𝑛 ) ∈ 𝐻𝑛 (𝑋, {𝑥0 }; Z). The long exact
homology sequence of the pair (𝑋, {𝑥0 }) yields for 𝑛 ≥ 1
148
3.14. Homotopy versus homology
Proof. Consider the map 𝑠1 : 𝑊1𝑛 = [0, 21 ] × [0, 1] 𝑛−1 → 𝑊 𝑛 given by (𝑥1 , . . . , 𝑥 𝑛 ) ↦→
(2𝑥1 , 𝑥2 , . . . , 𝑥 𝑛 ) and the map 𝑠2 : 𝑊2𝑛 = [ 21 , 1] × [0, 1] 𝑛−1 → 𝑊 𝑛 defined by (𝑥1 , . . . , 𝑥 𝑛 ) ↦→
(2𝑥1 − 1, 𝑥2 , . . . , 𝑥 𝑛 ).
𝑊𝑛 𝑊1𝑛 𝑊2𝑛 𝑊 𝑛
𝑠1
𝑠2
𝑛 = 1: 𝑛 = 2:
𝜎1 𝜎3 𝑐1 = 𝜎1 + 𝜎2
𝑐2 = 𝜎3 + 𝜎4
𝜎2 𝜎4
𝑐1 𝑐2
ℎ([ 𝑓1 ] · [ 𝑓2 ]) = ℎ([𝑔])
= 𝐻𝑛 ( 𝑗) −1 𝐻𝑛 (𝑔) (𝛼𝑛)
= 𝐻𝑛 ( 𝑗) −1 𝐻𝑛 (𝑔) ([𝑐1 + 𝑐2 ])
= 𝐻𝑛 ( 𝑗) −1 𝐻𝑛 (𝑔) ([𝑐1 ]) + 𝐻𝑛 ( 𝑗) −1 𝐻𝑛 (𝑔) ([𝑐2 ])
= 𝐻𝑛 ( 𝑗) −1 𝐻𝑛 ( 𝑓1 ◦ 𝑠1 ) ([𝑐1 ]) + 𝐻𝑛 ( 𝑗) −1 𝐻𝑛 ( 𝑓2 ◦ 𝑠2 ) ([𝑐2 ])
149
3. Homology Theory
Proposition 3.118. The Hurewicz homomorphism ℎ is natural, i.e., for every 𝑓 ∈ 𝐶 (𝑋, 𝑌 )
with 𝑓 (𝑥0 ) = 𝑦 0 the following diagram commutes:
ℎ / 𝐻𝑛 (𝑋; Z)
𝜋 𝑛 (𝑋; 𝑥0 )
𝑓# 𝐻𝑛 ( 𝑓 )
ℎ
𝜋 𝑛 (𝑌 ; 𝑦 0 ) / 𝐻𝑛 (𝑌 ; Z)
150
3.14. Homotopy versus homology
Then
ℎ : 𝜋 𝑛 (𝑋; Z) → 𝐻𝑛 (𝑋; Z)
is an isomorphism.
In particular, we conclude that 𝐻1 (𝑋; Z) = . . . = 𝐻𝑛−1 (𝑋; Z) = 0. 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; 𝑥0 ) is not in general. However, ℎ induces a homomorphism
Already Poincaré showed that if 𝜋0 (𝑋, 𝑥0 ) = {1} (i.e., 𝑋 is path-connected) then the map ℎ̄ is an
isomorphism.
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.
151
3. Homology Theory
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 )
Remark 3.125. Under the assumptions of the theorem of Hurewicz 3.119 not much can be
said about ℎ : 𝜋 𝑘 (𝑋, 𝑥0 ) → 𝐻 𝑘 (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 𝐻 𝑘 (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 𝜋 𝑘 (𝑆 𝑛 )
for 𝑘 > 𝑛 is a difficult problem and many of these groups are not known to date.
3.15. Exercises
3.1. Let 𝑋 be a topological space. Show:
a) If 𝑋 is path-connected then
𝐻0 (𝑋; 𝑅) 𝑅
3.2. Let 𝑌𝑘 = {1, ..., 𝑘} be equipped with the discrete topology. Compute 𝐻𝑛 (𝑌𝑘 ; 𝑅) without
using Exercise 3.1. Instead use the Eilenberg-Steenrod axioms.
Hint: Consider the pair (𝑌𝑘 , 𝑌𝑘−1 ).
𝜕 # : 𝑆0 (𝑋; 𝑅) → 𝑅,
Õ Õ
𝜕# 𝛼𝑖 𝜎𝑖 = 𝛼𝑖 ,
152
3.15. Exercises
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}.
3.6. Let 𝑝 be a complex polynomial without zeros on the unit circle 𝑆1 ⊂ C. Show: The degree
of the map
𝑝(𝑧)
𝑝ˆ : 𝑆1 → 𝑆1 , 𝑝(𝑧)ˆ = ,
| 𝑝(𝑧)|
coincides with the number of zeros of 𝑝 in the interior of the unit disk (counted with multiplicities).
3.7 (Homotopy invariance of the local mapping degree). Let 𝑉 ⊂ 𝑆 𝑛 be open and 𝐹 : 𝑉 ×
Ð
[0, 1] → 𝑆 𝑛 continuous. We put 𝑓𝑡 (𝑥) := 𝐹 (𝑥, 𝑡). Let 𝑝 ∈ 𝑆 𝑛 such that 𝑡 ∈ [0,1] 𝑓𝑡−1 ( 𝑝) is
compact. Show:
deg 𝑝 ( 𝑓1 ) = deg 𝑝 ( 𝑓0 ).
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
153
3. Homology Theory
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′ ′
𝑓𝑛
𝑛+1
· · · → 𝐺 ′𝑛+1 −−−−−−−−→ 𝐺 ′𝑛 −−−−−−→ 𝐺 ′𝑛−1 → · · ·
is a subsequence, i.e., 𝐺 ′𝑛 ⊂ 𝐺 𝑛 and 𝑓𝑛′ = 𝑓𝑛 | 𝐺𝑛′ . Prove that the subsequence is exact if and only
if the quotient sequence
is exact.
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)
b) Let 𝑝 1 , ..., 𝑝 𝑛 ∈ R2 be pairwise distinct. Compute 𝐻1 (R2 \{𝑝 1 , ..., 𝑝 𝑛 }) and sketch generators.
154
3.15. Exercises
Sketch generators.
3.21. Show:
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,
Ø
𝑘
𝑋= { 𝑗 } × 𝑆𝑛 ∼
𝑗=1
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
A. Appendix
We usually use a somewhat sloppy notation and will not distinuish between an element 𝑠 ∈ 𝑆
and the corresponding function 𝑓𝑠 : 𝑆 → 𝑅. Thus, instead of (A.1) we will write
Õ
𝑓 = 𝑓 (𝑠)𝑠.
𝑠∈𝑆
1 If you
are not familiar with modules over rings think of the special case that 𝑅 is a field. Then an 𝑅-module is just
an 𝑅-vector space.
157
A. Appendix
In Section 3.1 we consider the free 𝑅-module generated by 𝑆 = 𝐶 (Δ𝑛 , 𝑋) and write 𝑆𝑛 (𝑋; 𝑅)
instead of 𝑌 .
158
List of Figures
1. (Non-) homeomorphic spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2. Hilbert’s curve2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3. Surface classification3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
4. Open subset . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
5. Hausdorff property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
6. Interval with endpoints identified . . . . . . . . . . . . . . . . . . . . . . . . . 11
7. Comb space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
8. 𝐴 is not a strong deformation retract of 𝑋 . . . . . . . . . . . . . . . . . . . . 19
9. Concatenation of paths . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
10. Concatenations of homotopic paths are homotopic. . . . . . . . . . . . . . . . 21
11. Concatenation with constant path is homotopic to original path . . . . . . . . . 22
12. Inverse modulo homotopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
13. Associativity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
14. Dependence on base point . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
15. Auxiliary homotopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
16. The deformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
17. Cutting a sandwich4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
18. The hyperplane function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
19. The volume function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
20. Monotonicity of the volume function . . . . . . . . . . . . . . . . . . . . . . . 34
21. The disk and its boundary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
22. Constructing a retraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
23. Connected but not path-connected . . . . . . . . . . . . . . . . . . . . . . . . 40
24. Subdividing 𝜔 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
25. The homotopy to start with . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
26. Subdividing the homotopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
27. Deforming the homotopy in each square . . . . . . . . . . . . . . . . . . . . . 48
28. Deforming along one row . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
29. Covering the figure 8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
30. Punctured manifold . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
31. Start with two manifolds... . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
32. ... and consider their connected sum. . . . . . . . . . . . . . . . . . . . . . . . 51
33. Surface of genus 𝑔 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
34. Building a surface from the disk . . . . . . . . . . . . . . . . . . . . . . . . . 52
35. Starting the induction with the torus . . . . . . . . . . . . . . . . . . . . . . . 53
36. Induction step by cutting off a segment . . . . . . . . . . . . . . . . . . . . . . 53
159
List of Figures
160
List of Figures
161
Bibliography
[1] Greenberg, M. J.; Harper, J. R.: Algebraic topology. A first course, Benjamin/Cummings
Publishing, 1981
[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
[6] Peano, G.: Sur une courbe, qui remplit toute une aire plane, Math. Ann. 36 (1890), 157–160
[8] Warner, F. W.: Foundations of differentiable manifolds and Lie groups, Springer-Verlag,
New York, 1983
163
Index
Symbols H, quaternions . . . . . . . . . . . . . . . . . . . . . . . . 99
ℎ, Hurewicz homomorphism . . . . . . . . . . 149
𝑎, antipodal map . . . . . . . . . . . . . . . . . . . . . . 96 𝐻𝑛 ( 𝑓 ), map induced on homology . . . . . 83
¯ closure of 𝐴 . . . . . . . . . . . . . . . . . . . . . . . 89
𝐴, 𝐻𝑛 𝐾∗ , 𝑛-th homology . . . . . . . . . . . . . . . . 106
˚ interior of 𝐴 . . . . . . . . . . . . . . . . . . . . . . . 89
𝐴, 𝐻𝑛 (𝜑∗), map induced on homology . . . 107
𝑏 𝑗 (𝑋; 𝑅), Betti number . . . . . . . . . . . . . . 141 𝐻𝑛 (𝑋, 𝐴; 𝑅), relative 𝑛-th singular
𝐵𝑛 , barycenter . . . . . . . . . . . . . . . . . . . . . . . 118 homology . . . . . . . . . . . . . . . . . . . 84
𝐵𝑛 𝐾∗ , 𝑛-boundaries . . . . . . . . . . . . . . . . . . 106 𝐻𝑛 (𝑋; 𝑅), 𝑛-th singular homology . . . . . . 82
𝐵𝑛 (𝑋, 𝐴; 𝑅), relative singular 𝑛-boundaries HP𝑛 , quaternionic projective space . . . . 135
84 𝐾𝑛 (𝑋, X), cellular complex . . . . . . . . . . . 137
𝐵𝑛 (𝑋; 𝑅), singular 𝑛-boundaries . . . . . . . 82 N (𝑆), normal subgroup generated by S . 41
𝐵(𝑥, 𝑟), open ball of radius 𝑟 centered at 𝑥 6 Ω (𝑋; 𝑥0 , 𝑥1 ), set of paths . . . . . . . . . . . . . . 20
𝜒(𝑋, X), Euler number. . . . . . . . . . . . . . . 141 Ω (𝑋; 𝑥0 ), set of loops . . . . . . . . . . . . . . . . . 20
𝐶𝑛 , barycentric subdivision . . . . . . . . . . . 118 𝑃, prism operator . . . . . . . . . . . . . . . . . . . . 114
CP𝑛 , complex projective space. . . . .77, 134 𝜑 𝜎 , attaching map . . . . . . . . . . . . . . . . . . . 132
𝐶 𝑋, cone over 𝑋 . . . . . . . . . . . . . . . . . . . . . . 14 𝜋1 (𝑋; 𝑥0 ), fundamental group . . . . . . . . . . 21
𝐶 ((𝑋, 𝐴), (𝑌 , 𝐵)), set of continuous maps 𝜋 𝑛 (𝑋; 𝑥0 ), 𝑛-th homotopy group . . . . . . . . 58
between pairs of spaces . . . . . . . 83 𝑄 𝑛 , chain homotopy for subdivision map
𝐶 (𝑋, 𝑌 ), set of continuous maps . . . . . . . 15 119
deg, degree . . . . . . . . . . . . . . . . . . . . . . . 29, 97 RP𝑛 , real projective space . . . . . 73, 77, 133
deg 𝑝 , local degree . . . . . . . . . . . . . . . . . . . 102 Sd𝑛 , subdivision chain map . . . . . . . . . . . 119
𝐷 𝑛 , closed 𝑛-dimensional unit ball . . . . . . . 6 𝑆aff 𝑛
𝑘 (Δ ), affine chains . . . . . . . . . . . . . . . . 118
Δ𝑛 , standard simplex . . . . . . . . . . . . . . . . . . 79 𝑆𝑛 ( 𝑓 ), map induced on singular chains. . 82
𝐷 𝑛− , lower hemisphere . . . . . . . . . . . . . . . . . 91 𝑆𝑛 (𝑋, 𝐴; 𝑅), relative singular 𝑛-chains . . 83
𝐷 +𝑛 , upper hemisphere . . . . . . . . . . . . . . . . . 91 𝑆𝑛 (𝑋; 𝑅), singular 𝑛-chains . . . . . . . . . . . . 80
𝜕, boundary of singular chain . . . . . . . . . . 81 𝑆𝑛U (𝑋, 𝐴), small singular 𝑛-chains. . . . .117
𝜕, connecting homomorphism for Σ𝑋, suspension of 𝑋 . . . . . . . . . . . . . . . . . . 14
homology groups . . . . . . . . 87, 108 𝜎 (𝑖) , 𝑖-th face of singular simplex . . . . . . 80
𝜕, connecting homomorphism for 𝑇𝑛𝑖 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
homotopy groups . . . . . . . . . . . . . 66 TR𝑛 , standard topology of 𝑅 𝑛 . . . . . . . . . . . . 6
¯
𝜕, boundary of relative singular chain . . . 84 T𝑋 , topology of 𝑋 . . . . . . . . . . . . . . . . . . . . . . 7
𝜀 𝑥0 , constant loop . . . . . . . . . . . . . . . . . . . . . 20 𝑈 ( 𝑓 , 𝑝), winding number . . . . . . . . . . . . . . 75
Exp, covering map of 𝑆1 . . . . . . . . . . . . . . . 27 𝑊 𝑛 , 𝑛-dimensional cube . . . . . . . . . . . . . . . 13
𝑓# , map induced on fundamental group . . 23 𝑋 𝑛 , 𝑛-skeleton . . . . . . . . . . . . . . . . . . . . . . 132
𝐹𝑔 , surface of genus 𝑔 . . . . . . . . . . . . . . . . . . 5 𝑍 𝑛 𝐾∗ , 𝑛-cycles . . . . . . . . . . . . . . . . . . . . . . 106
𝐹𝑛𝑖 , face map . . . . . . . . . . . . . . . . . . . . . . . . . 80 𝑍 𝑛 (𝑋, 𝐴; 𝑅), relative singular 𝑛-cycles . . 84
165
Index
B E
166
Index
HLP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 O
homeomorphic . . . . . . . . . . . . . . . . . . . . . . . . 8
open (for the product topology) . . . . . . . . 12
homeomorphism . . . . . . . . . . . . . . . . . . . . . . . 8
open subset of R𝑛 . . . . . . . . . . . . . . . . . . . . . . 6
homogeneous space . . . . . . . . . . . . . . . . . . . 65
open subset of a topological space . . . . . . . 7
homology . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
augmented . . . . . . . . . . . . . . . . . . . . . . 153 P
relative singular . . . . . . . . . . . . . . . . . . 84
singular . . . . . . . . . . . . . . . . . . . . . . . . . .82 pair of spaces. . . . . . . . . . . . . . . . . . . . . . . . . 83
homotopic chain maps . . . . . . . . . . . . . . . 111 path . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
homotopic relative to 𝐴 . . . . . . . . . . . . . . . 15 path-connected . . . . . . . . . . . . . . . . . . . . . . . 39
homotopy axiom . . . . . . . . . . . . . . . . . . . . . . 88 plane-filling curve . . . . . . . . . . . . . . . . . . . . . 4
homotopy equivalence . . . . . . . . . . . . 16, 112 pointed set . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
homotopy equivalent . . . . . . . . . . . . . . . . . . 16 presentation of a group . . . . . . . . . . . . . . . . 55
homotopy group . . . . . . . . . . . . . . . . . . . . . . 58 prism operator. . . . . . . . . . . . . . . . . . . . . . .114
homotopy lifting property . . . . . . . . . . . . . 61 projective space
homotopy relative to 𝐴 . . . . . . . . . . . . . . . . 15 complex . . . . . . . . . . . . . . . . . . . . . 77, 134
Hopf fibration . . . . . . . . . . . . . . . . . . . . . . . . 77 quaternionic . . . . . . . . . . . . . . . . . . . . 135
Hurewicz homomorphism . . . . . . . . . . . . 150 real . . . . . . . . . . . . . . . . . . . . . 73, 77, 133
Q
I
quaternionic projective space . . . . . . . . . 135
incidence numbers . . . . . . . . . . . . . . . . . . . 145 quaternions . . . . . . . . . . . . . . . . . . . . . . . . . . 99
induced topology of metric space . . . . . . . . 7
induced topology of subspace . . . . . . . . . . . 7 R
interior . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 real projective space . . . . . . . . . . 73, 77, 133
L relative singular homology . . . . . . . . . . . . . 84
retract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
lift . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 retraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
local degree . . . . . . . . . . . . . . . . . . . . . . . . . 102
S
long exact homology sequence of a pair of
spaces . . . . . . . . . . . . . . . . . . . . . . . 89 Seifert-van Kampen theorem . . . . . . . . . . . 46
long exact homotopy sequence of a Serre Serre fibration . . . . . . . . . . . . . . . . . . . . . . . . 62
fibration . . . . . . . . . . . . . . . . . . . . . 69 simply-connected . . . . . . . . . . . . . . . . . . . . . 39
loop . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 singular 𝑛-boundary . . . . . . . . . . . . . . . . . . . 82
singular 𝑛-chain . . . . . . . . . . . . . . . . . . . . . . 80
M singular 𝑛-cycle . . . . . . . . . . . . . . . . . . . . . . 81
singular 𝑛-simplex . . . . . . . . . . . . . . . . . . . . 79
Mayer-Vietoris sequence . . . . . . . . . . . . . 125
singular homology . . . . . . . . . . . . . . . . . . . . 82
Möbius strip . . . . . . . . . . . . . . . . . . . . . . . . . 76
skeleton . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
N small chain . . . . . . . . . . . . . . . . . . . . . . . . . 117
small chain theorem . . . . . . . . . . . . . . . . . 117
natural . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 split exact sequence . . . . . . . . . . . . . . . . . . . 78
normal. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .40 standard simplex . . . . . . . . . . . . . . . . . . . . . . 79
normal subgroup generated by a set . . . . . 41 standard topology of R𝑛 . . . . . . . . . . . . . . . . 6
167
Index
T vector field . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
total space . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 W
U
weak fibration . . . . . . . . . . . . . . . . . . . . . . . . 62
universal property winding number . . . . . . . . . . . . . . . . . . . . . . 75
168