Algebraic Topology: Or257/teaching/notes/at PDF
Algebraic Topology: Or257/teaching/notes/at PDF
Oscar Randal-Williams
https://www.dpmms.cam.ac.uk/∼or257/teaching/notes/at.pdf
1 Introduction 1
1.1 Some recollections and conventions . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Cell complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
3 Covering spaces 15
3.1 Covering spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.2 The fundamental group of the circle and applications . . . . . . . . . . . . 21
3.3 The construction of universal covers . . . . . . . . . . . . . . . . . . . . . 22
3.4 The Galois correspondence . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
6 Simplicial complexes 45
6.1 Simplicial complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
6.2 Simplicial approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
7 Homology 54
7.1 Simplicial homology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
7.2 Some homological algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
7.3 Elementary calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
7.4 The Mayer–Vietoris sequence . . . . . . . . . . . . . . . . . . . . . . . . . 61
7.5 Continuous maps and homotopy invariance . . . . . . . . . . . . . . . . . 64
7.6 Homology of spheres and applications . . . . . . . . . . . . . . . . . . . . 68
7.7 Homology of surfaces, and their classification . . . . . . . . . . . . . . . . 70
7.8 Rational homology, Euler and Lefschetz numbers . . . . . . . . . . . . . . 75
Last updated December 13, 2016. Corrections to [email protected].
Chapter 1
Introduction
Algebraic Topology is the art of turning existence questions in topology into existence
questions in algebra, and then showing that the algebraic object cannot exist: this then
implies that the original topological object cannot exist. This procedure usually has a
loss of information, and if the algebraic object does exist it does not typically allow us
to show that the original topological object does.
Many interesting topological problems can be expressed in the following form.
The most basic instance of this problem arises when taking X = Dn to be the n-
dimensional disc, A = Y = S n−1 to be the (n − 1)-dimensional sphere, and f = IdS n−1 :
S n−1 → S n−1 to be the identity. The problem then asks whether there is a continuous
function from the disc to its boundary which fixes each point on the boundary. In the
course we shall prove the following theorem.
Theorem. There is no continuous function F : Dn → S n−1 such that the composition
S n−1
incl. / Dn F / S n−1
is the identity.
You may find it intuitively quite clear that there is no such F , but the difficulty in
proving this is apparent: while it is obvious that various naïve choices of F : Dn →
S n−1 do not work, we must show that no choice can work, and in principle there are a
great many potential choices with very little mathematical structure to work with. The
machinery of Algebraic Topology translates proving the theorem above to proving:
Theorem. There is no group homomorphism F : {0} → Z such that the composition
Z
0 / {0} F /Z
is the identity.
I hope you will agree that this algebraic theorem is considerably easier to prove. There
are several other classical results in mathematics which we shall be able to rephrase as
an extension problem—or something like it—and hence solve using the machinery of
Algebraic Topology. One is that the notion of dimension is well-defined:
Theorem. If there is a homeomorphism Rn ∼ = Rm , then n = m.
Another is the fundamental theorem of algebra:
1
2 Chapter 1 Introduction
We will need to repeatedly justify why various formulas that we write down do indeed
defione maps, i.e. are continuous. Almost all of these situations will be dealt with using
the following convenient lemma.
Lemma 1.1.3 (Lesbegue number lemma). Let (X, d) be a metric space which is compact.
For any open cover U = {Uα }α∈I of X there exists a δ > 0 such that each ball Bδ (x) ⊂ X
is contained entirely in some Uα .
Proof. Suppose no such δ exists. Then for each n ∈ N there exist an xn ∈ X such that
B1/n (xn ) is not contained entirely in some Uα . As X is compact the set {xn }n∈N has a
limit point, y. This lies in some Uα , so there exists an r > 0 such that Br (y) also lies in
Uα . But now let N > 0 be such that
1
(i) N < 2r , and
Then B1/N (xN ) is contained inside Br (y), and hence inside Uα , a contradiction.
1.2 Cell complexes 3
A convenient way to construct topological spaces, which we shall see interacts well with
the tools of Algebraic Topology, is the notion of attaching cells.
Definition 1.2.1. For a space X and a map f : S n−1 → X, the space obtained by
attaching an n-cell to X along f is the quotient space
X ∪f Dn := (X t Dn )/ ∼
(i) starting with a finite set X 0 with the discrete topology, called the 0-skeleton,
(ii) having defined the (n − 1)-skeleton X n−1 , form the n-skeleton X n by attaching a
collection of n-cells along finitely many maps {fα : S n−1 → X n−1 }α∈I , i.e.
!
G
n n−1
X = X t Dα /x ∈ Sαn−1 ⊂ Dαn ∼ fα (x) ∈ X n−1 ,
n
α∈I
2.1 Homotopy
Proposition 2.1.2. For spaces X and Y and a subspace A ⊂ X, the relation “homotopic
relative to A" on the set of maps from X to Y is an equivalence relation.
4
2.1 Homotopy 5
r : R2 \ {0} −→ S 1
x
x 7−→ |x| .
Then r ◦ i = IdS 1 , so we may take the constant homotopy. On the other hand the
x
composition i ◦ r : R2 \ {0} → R2 \ {0} is x →
7 |x| . Define a homotopy
t
which is well-defined as if t+(1−t)·|x| = 0 then |x| = t−1 ≤ 0, which is impossible. This
is a homotopy from i ◦ r to IdR2 \{0} , and so i is a homotopy equivalence with homotopy
inverse r; S 1 ' R2 \ {0}.
This property is central in the subject, and gets its own name.
We wish to say that the relation of homotopy equivalence between spaces is an equiv-
alence relation, but first need a tool.
f
i /X ) h /Z
T 5Y
g
such that f ' g, then h ◦ f ' h ◦ g and f ◦ i ' g ◦ i. This is the form in which we will
generally use this lemma.
Proof.
f f0
) )
Xi Y i Z
g g0
such that
and
(f 0 ◦ f ) ◦ (g ◦ g 0 ) = f 0 ◦ (f ◦ g) ◦ g 0 ' f 0 ◦ IdY ◦ g 0 = f 0 ◦ g 0 ' IdZ .
2.2 Paths
Definition 2.2.1 (Paths and loops). For a space X and points x0 , x1 ∈ X, a path from
x0 to x1 is a map γ : I → X such that γ(0) = x0 and γ(1) = x1 . We write γ : x0 x1 .
If x0 = x1 then we call γ a loop based at x0 .
2
If we insist that i ◦ r ' IdX rel A then this is called a strong deformation retraction. Some
authors call this stronger notion “deformation retraction", so beware.
8 Chapter 2 Homotopy and the fundamental group
(This is continuous by the Gluing Lemma.) This is a path from x0 to x2 . We also define
a path γ −1 : T → X, the inverse of γ, by the formula
which gives a path from x1 to x0 . Finally, we define a path cx0 : I → X, the constant
path at the point x0 , by cx0 (t) = x0 .
then the above three constructions (concatenation, inverse, constant paths) show that
this is an equivalence relation.
Definition 2.2.2 (Path components and π0 ). The equivalence classes of ∼ are called
path components, and the set of equivalence classes is denoted π0 (X). If there is a
single equivalence class then we say X is path connected.
π0 (f ) : π0 (X) −→ π0 (Y )
Proof. We must show that the proposed function is well-defined, so let [x] = [x0 ] ∈ π0 (X):
this means that there is a path γ : I → X from x to x0 . Then f ◦ γ : I → Y is a path
from f (x) to f (x0 ), and so [f (x)] = [f (x0 )] ∈ π0 (Y ), as required.
Properties (ii) and (iii) are clear from the definition. For (i), let H : X × I → Y be
a homotopy from f to g. For x ∈ X the map H(x, −) : I → Y is a path from f (x) to
g(x), and so [f (x)] = [g(x)] ∈ π0 (Y ). Thus π0 (f )([x]) = π0 (g)([x]), and this holds for
any x so π0 (f ) = π0 (g) as required.
and similarly
π0 (g) ◦ π0 (f ) = π0 (g ◦ f ) = π0 (IdX ) = Idπ0 (X)
so π0 (g) is inverse to π0 (f ).
Example 2.2.5. The space {−1, 1} with the discrete topology is not contractible, i.e. is
not homotopy equivalent to the one-point space {∗}. This is because continuous paths in
{−1, 1} must be constant, so π0 ({−1, 1}) = {−1, 1} has cardinality 2. But π0 ({∗}) = {∗}
has cardinality 1, so these are not in bijection.
Example 2.2.6. The space R2 does not retract (cf. Definition 2.1.10) on to the subspace
{(−1, 0), (1, 0)}. If we write i : {(−1, 0), (1, 0)} → R2 for the inclusion, and suppose
r : R2 → {(−1, 0), (1, 0)} is a retraction, so r ◦ i = Id{(−1,0),(1,0)} , then we obtain a
factorisation
π0 (i) π0 (r)
π0 ({(−1, 0), (1, 0)}) −→ π0 (R2 ) −→ π0 ({(−1, 0), (1, 0)})
of the identity map. By the previous example π0 ({(−1, 0), (1, 0)}) has cardinality 2, but
R2 is path-connected so π0 (R2 ) has cardinality 1. In particular, π0 (i) cannot be injective,
so π0 (r) ◦ π0 (i) cannot be the identity map.
Definition 2.2.7. Two paths γ, γ 0 : I → X from x0 to x1 are homotopic as paths
if they are homotopic relative to {0, 1} ⊂ I in the sense of Definition 2.1.1. We write
γ ' γ 0 as paths from x0 to x1 .
Lemma 2.2.8. If γ0 ' γ1 as paths from x0 to x1 , and γ00 ' γ10 as paths from x1 to x2 ,
then γ0 · γ00 ' γ1 · γ10 as paths from x0 to x2 .
Proof. Let H be a relative homotopy from γ0 to γ1 , and H 0 be a relative homotopy from
γ00 to γ10 . As H(1, t) = x1 = H 0 (0, t) for all t, these homotopies fit together to a function
H 00 : I × I → X given by
(
H(2s, t) 0 ≤ s ≤ 1/2
H 00 (s, t) = 0
H (2s − 1, t) 1/2 ≤ s ≤ 1
This satisfies
H 00 (−, 0) = (γ0 · γ00 )(−) H 00 (−, 1) = (γ1 · γ10 )(−) H 00 (0, t) = x0 H 00 (1, t) = x2
as required.
(iii) γ0 · γ0−1 ' cx0 as paths from x0 to x0 , and γ0−1 · γ0 ' cx1 as paths from x1 to x1 .
Theorem 2.3.1. Let X be a space and x0 ∈ X be a point. Let π1 (X, x0 ) denote the set
of homotopy classes of loops in X based at x0 . Then the rule [γ] · [γ 0 ] := [γ · γ 0 ] and the
element e := [cx0 ] define a group structure on π1 (X, x0 ).
Proof. Lemma 2.2.8 shows that the proposed composition law is well-defined, and Propo-
sition 2.2.9 shows that the group axioms are satisfied.
(i) Note that f ◦cx0 = cy0 , so π1 (f ) preserves the unit. Also f ◦(γ ·γ 0 ) = (f ◦γ)·(f ◦γ 0 ),
so π1 (f ) preserves the composition law.
(ii) if f ' f 0 rel {x0 } then f ◦ γ ' f 0 ◦ γ rel {0, 1}, as γ({0, 1}) = {x0 }.
(iii) π1 (k ◦h)([γ]) = [k ◦h◦γ] = π1 (k)([h◦γ]) = π1 (k)(π1 (h)([γ])), for any [γ] ∈ π1 (A, a).
After having introduced the notation π1 (f ) for the group homomorphism induced by
a based map f , will will now discard it in favour of the shorter notation f∗ .
u# : π1 (X, x0 ) −→ π1 (X, x1 )
[γ] 7−→ [u−1 · γ · u]
satisfying
u# (f ◦u)#
f∗
π1 (X, x1 ) / π1 (Y, y1 )
commutes: going either way around it from the top left corner to the bottom right
corner yields the same homomorphism.
(v) if x1 = x0 then u# is the automorphism of π1 (X, x0 ) given by conjugation by
[u] ∈ π1 (X, x0 ).
Proof. That u# is a group homomorphism is a consequence of invoking Proposition 2.2.9
several times. Properties (i) – (iii) and (v) are by now standard arguments. For (iv) we
calculate
((f ◦ u)# ◦ f∗ )([γ]) = (f ◦ u)# ([f ◦ γ]) = [(f ◦ u)−1 · (f ◦ γ) · (f ◦ u)]
= [f ◦ (u−1 · γ · u)] = f∗ ([u−1 · γ · u]) = f∗ (u# ([γ])).
Warning 2.3.5. By this proposition the fundamental groups based at different points
in a path connected space X are isomorphic, but they are not canonically isomorphic!
One has to choose a path between their basepoints to obtain an isomorphism.
Thus it makes sense to say “the fundamental group is trivial / abelian / has group-
theoretic property P" without referring to a basepoint. But it does not make sense
to have an element a0 ∈ π1 (X, x0 ) and say “let a1 ∈ π1 (X, x1 ) be the corresponding
element".
If H : X × I → Y is a homotopy from f to g, and x0 ∈ X is a basepoint, then
u := H(x0 , −) : I → Y is a path from f (x0 ) to g(x0 ) in Y . We can ask about the
relationship between the three homomorphisms
f∗ : π1 (X, x0 ) −→ π1 (Y, f (x0 ))
g∗ : π1 (X, x0 ) −→ π1 (Y, g(x0 ))
u# : π1 (Y, f (x0 )) −→ π1 (Y, g(x0 )),
which fit into the diagram
π1 (Y, f (x0 ))
7
f∗
∼
= u# (2.3.1)
g∗
π1 (X, x0 ) / π1 (Y, g(x0 )).
2.3 The fundamental group 13
Consider the path `+ : I → I × I from (0, 1) to (1, 1) given by s 7→ (s, 1), and the path
`− : I → I × I from (0, 1) to (1, 1) given by concatenating the paths
These are paths between the same points, and we can define a homotopy
between them, using the vector space structure on R2 ⊃ I × I and the fact that I × I is
convex. This shows that `+ ' `− as paths from (0, 1) to (1, 1).
Hence F ◦ `+ ' F ◦ `− as paths from F (0, 1) = H(x0 , 1) = g(x0 ) to F (1, 1) =
H(x0 , 1) = g(x0 ). But [F ◦ `+ ] = [g ◦ γ] and [F ◦ `− ] = [u−1 · (f ◦ γ) · u], as required.
is an isomorphism.
which is an isomorphism: thus the map f∗ we are studying is injective, and the map
is surjective. If can show that g∗ is also injective then we are done, as then it is an
isomorphism and so f∗ = (g∗ )−1 ◦ u0# is also an isomorphism.
To show that g∗ is injective we consider the path u : I → Y given by u(t) =
H(f (x0 ), 1 − t), going from f (x0 ) to f ◦ g ◦ f (x0 ). Applying Lemma 2.3.6 again gives
g∗
u# = (f ◦ g)∗ = f∗ ◦ g∗ : π1 (Y, f (x0 )) −→ π1 (X, g ◦ f (x0 )) −→ π1 (Y, f ◦ g ◦ f (x0 )),
which is an isomorphism. (Note the second map here is induced by f but is not the map
f∗ we are studying, as the basepoints involved are different.) Thus the map (2.3.2) is
also injective, as required.
Lemma 2.3.10. A space X is simply connected if and only if for each pair of points
x0 , x1 ∈ X there is a unique homotopy class of path between them.
Covering spaces
For such an open neighbourhood U , we shall usually write p−1 (U ) = α∈I Vα where
`
each p|Vα : Vα → U is a homeomorphism. We call these open sets U evenly covered.
Example 3.1.4. Let S 1 ⊂ C be the set of complex numbers of modulus 1, and consider
the map p : R → S 1 given by p(t) = e2πit . For the set Uy>0 := {x + iy ∈ S 1 | y > 0} we
have a
p−1 (Uy>0 ) = (j, j + 12 ),
j∈Z
p| 1 : (j, j + 12 ) −→ Uy>0
(j,j+ 2 )
Uy>0 −→ (j, j + 12 )
x + iy 7−→ j + arccos(x)/2π.
15
16 Chapter 3 Covering spaces
Similarly with Uy<0 , Ux>0 and Ux<0 , and these four open sets cover S 1 . Thus p is a
covering map.
Example 3.1.6. Let S 2 ⊂ R3 be the unit sphere, and ∼ be the equivalence relation on
S 2 generated by x ∼ −x, with quotient space RP2 := S 2 / ∼ (the real projective plane)
and quotient map p : S 2 → RP2 . Let V := {(x, y, z) ∈ s2 | z 6= 0}, an open set, and
U = p(V ). Then p−1 (U ) = V`, and so U ⊂ RP2 is open.
Furthermore, V = Vz<0 Vz>0 , and we claim that p|Vz>0 : Vz>0 → U is a homeo-
morphism (similarly for Vz<0 of course). To see this, we shall construct an inverse. By
the definition of the quotient topology, a continuous map U → Y for some space Y is
the same as a continuous map V → Y which in constant on ∼-equivalence classes. Thus
define the map
t : V −→ Vz>0
(
(x, y, z) if z > 0,
(x, y, z) 7−→
(−x, −y, −z) if z < 0.
This is clearly continuous and constant on ∼-equivalence classes, and so defines a con-
tinuous function t̄ : U → Vz>0 . It is inverse to p|Vz>0 , and so p is a covering map.
Our interest in covering spaces comes from the following important property they
have with respect to paths and homotopies.
3.1 Covering spaces 17
(i) H(−,
e 0) = f˜0 (−), and
(ii) p ◦ H
e = H.
{Uα }α∈I be an open cover of X by sets which are evenly covered, and write
Proof. Let `
∼
p−1 (Uα ) = β∈Iα Vβ with p|Vβ → Uα .
Then {H −1 (Uα )}α∈I is an open cover of Y × I. For each y0 ∈ Y this restricts to an
open cover of the compact space {y0 } × I, and by the Lesbegue number lemma there is
a N = N (y0 ) such that each path
H|{y0 }×[i/N,(i+1)/N ] : {y0 } × [ Ni , i+1
N ] −→ X
lies entirely inside some Uα . In fact, as {y0 }×I is compact there is an open neighbourhood
Wy0 of y0 such that H(Wy0 × [ Ni , i+1 N ]) lies entirely inside some Uα for each i. We are
now in the situation depicted in Figure 3.3.
We obtain a lift H|
e Wy ×I of H|Wy ×I as follows:
0 0
H|
e 1 := (p|Vβ )−1 ◦ H| 1 : Wy0 × [0, N1 ] −→ Vβ ⊂ X.
e
Wy0 ×[0, N ] Wy0 ×[0, N ]
18 Chapter 3 Covering spaces
Figure 3.3
(ii) Proceed in the same way, lifting the (short) homotopy H| 1 2 starting at
Wy0 ×[ N , N ]
H|
e 1 , and so on.
Wy0 ×{ N }
At the end of this process, we obtain a map H| e Wy ×I lifting H|Wy ×I and extending
0 0
f˜0 |Wy0 .
We may do this for every point y0 ∈ Y , so it is now enough to check that on (Wy0 ×
I) ∩ (Wy1 × I) = (Wy0 ∩ Wy1 ) × I the two lifts we have constructed agree. The two
lifts do agree on (Wy0 ∩ Wy1 ) × {0}, as here they both restrict to f0 |Wy0 ∩Wy1 , and so
by Lemma 3.1.8 they agree on an open and closed subset of (Wy0 ∩ Wy1 ) × I containing
(Wy0 ∩ Wy1 ) × {0}. Restricting to each {y} × I we must obtain an open and closed subset
containing {y} × {0}, and so the whole of {y} × I (as I is connected): thus the two maps
agree on the whole of (Wy0 ∩ Wy1 ) × I.
The following corollary is obtained by taking Y = {∗} in the homotopy lifting lemma.
(ii) p ◦ γ̃ = γ.
(i) H(−,
e 1) is a lift of γ 0 starting at x̃0 , so is equal to γ̃ 0 ,
(ii) H(0, −) and H(1, −) are constant paths, so their lifts H(0,
e −) and H(1,
e 0) are too,
Proof. Let γ : I → X be a path from x0 to x1 , and for each y0 ∈ p−1 (x0 ) let γ̃y0 be the
unique lift of γ starting at y0 . Define the function
Similarly, the inverse path γ −1 defines a function (γ −1 )∗ : p−1 (x1 ) → p−1 (x0 ). Now for
any y0 ∈ p−1 (x0 ) we calculate
so (γ −1 )∗ is a left inverse for γ. The analogous calculation with the other composition
shows it is also a right inverse.
e → X is an n-sheeted cover (for n ∈
Definition 3.1.13. Say a covering map p : X
−1
N ∪ {∞}) if each p (x) has cardinality n.
We now come to the first connection between covering spaces and the fundamental
group.
Lemma 3.1.14. Let p : X e → X be a covering map, x0 ∈ X, and x̃0 ∈ p−1 (x0 ). Then
e x̃0 ) → π1 (X, x0 ) is injective.
the map p∗ : π1 (X,
Proof. Let γ : I → X e be a loop based at x̃0 such that p∗ ([γ]) = [cx ], so p ◦ γ ' cx as
0 0
paths. Let H be a homotopy of paths from p ◦ γ to cx0 , and lift it starting at the lift γ
of p ◦ γ. This gives a homotopy of paths H e from γ to a lift of cx , which must be cx̃ ,
0 0
and hence [γ] = [cx̃0 ].
In the proof of Corollary 3.1.12 we constructed, for a covering map p : X e → X and
−1 −1
path γ : I → X from x0 to x1 , a function γ∗ : p (x0 ) → p (x1 ). By Corollary 3.1.11
this function only depends on the homotopy class of γ as a path. In particular, this
construction defines a (right) action • of the group π1 (X, x0 ) on the set p−1 (x0 ).
e → X be a covering map, and X be path connected.
Lemma 3.1.15. Let p : X
(i) π1 (X, x0 ) acts transitively on p−1 (x0 ) if and only if X
e is path connected.
(iii) If X
e is path connected then there is a bijection
π1 (X, x0 )
−→ p−1 (x0 )
p∗ π1 (X,
e y0 )
Conversely, if [γ 0 ] ∈ π1 (X,
e y0 ) then γ 0 is a (and so the) lift of p ◦ γ 0 starting at y0 . It ends
at y0 too, and so y0 • [p ◦ γ 0 ] = y0 .
For (iii), we simply apply the Orbit-Stabiliser theorem from the theory of group
actions.
e → X is a universal cover if X
Definition 3.1.16. We say that a covering map p : X e
is simply connected.
The following connection between universal covers and the fundamental group is
immediate from this definition and Lemma 3.1.15.
e → X is a universal cover, then each x̃0 ∈
Corollary 3.1.17. If a covering map p : X
−1
p (x0 ) determines a bijection
∼
` : π1 (X, x0 ) −→ p−1 (x0 )
[γ] 7−→ x̃0 • [γ].
We wish to use this bijection to understand the fundamental group, but it must be
used carefully, as π1 (X, x0 ) is a group but p−1 (x0 ) does not have a group structure.
Unravelling the definitions shows that the multiplication law on p−1 (x0 ) induced by this
bijection can be described as follows: for y0 , z0 ∈ p−1 (x0 ), the product `(`−1 (y0 )·`−1 (z0 ))
is obtained by
Proof. Recall from Example 3.1.4 that we constructed a covering space p : R → S 1 via
p(t) = e2πit . The space R is contractible, and so simply connected, so this is a universal
cover. As a set we have p−1 (1) = Z ⊂ R, and we may choose x̃0 = 0 ∈ Z = p−1 (1) as a
basepoint. Acting on this basepoint then gives a bijection
` : π1 (S 1 , 1) −→ Z.
To compute `−1 (m) we can choose the path ũm : I → R given by ũn (t) = mt, which
goes from 0 to m, so that `−1 (m) = [p ◦ ũm ]. Then we find that
Thus the multiplication law on Z = p−1 (1) induced by ` is the usual addition of integers.
Furthermore, `([u]) = 1.
of the identity map. But D2 is contractible, and so π1 (D2 , 1) = {[c1 ]} is the trivial
group. This is a contradiction, as the identity map of Z cannot factor through a trivial
group.
Corollary 3.2.3 (Brouwer’s fixed point theorem, 1909). Every continuous map f : D2 →
D2 has a fixed point.
Proof. Suppose not, and let p(z) = z n +a1 z n−1 +· · ·+an be a (monic) polynomial having
no root. Fix (
|a1 | + |a2 | + · · · + |an |
r≥
1
and note that on the circle |z| = r we have the estimate
which is well-defined as the loop s 7→ pt (r · e2πis ) is never 0. When t = 0 this gives the
loop s 7→ e2πins , which represents n ∈ π1 (S 1 , 1). When t = 1 this gives the loop
It practice we can often directly construct a universal cover of a given space X, but for
theoretical work we need a general construction.
semilocally simply connected, and we have just argued that it is necessary in order
for a space to have a universal cover.
We may use this bijection in order to construct a universal cover, when we don’t yet have
one.
Theorem 3.3.1. Let X be path connected, locally path connected1 , and semilocally simply
e → X with X
connected. Then there exists a covering map p : X e simply connected.
(iii) The map π1 (U, x) → π1 (X, x) factors through the trivial map π1 (U 0 , x) → π1 (X, x),
so is trivial.
1
Recall from Example Sheet 1 that X is locally path connected if for every point x ∈ X and every
neighbourhood U 3 x, there exists a smaller neighbourhood x ∈ V ⊂ U such that V is path connected.
24 Chapter 3 Covering spaces
u# u#
trivial /
π1 (U, x) π1 (X, x)
where the vertical maps are isomorphisms, so the top map is trivial too.
Note that if x, y ∈ U ∈ U then there exists a path from x to y in U , and all such
paths are homotopic in X. For a [α] ∈ Xe and a U ∈ U such that α(1) ∈ U , define
Proof of claim. We must check that [β] ∈ (α0 , U0 ) ∩ (α1 , U1 ) has a neighbourhood of this
form.
Let W ⊂ U0 ∩ U1 be a neighbourhood of β(1) ∈ U0 ∩ U1 in the collection U, which
exists as U is a basis for the topology on X. Then (β, W ) is a neighbourhood of [β],
and it is enough to show that (β, W ) ⊂ (α0 , U0 ) ∩ (α1 , U1 ). If [γ] ∈ (β, W ) then it is a
concatenation of β and a path in W , but then it is a concatenation of α0 and a path in
U0 too; similarly, it is a concatenation of α1 and a path in U1 .
p|(α,U ) : (α, U ) → U
and so [δ] ∈ (β, U ). Thus (α, U ) ⊂ (β, U ), and the reverse inclusion follows by the same
argument.
Finally, we must show that X e is simply connected. The fundamental observation is
e starting at [cx ] ∈ X
that if γ : I → X is a path starting at x0 , then its lift to X 0
e is the
path
s 7−→ [t 7→ γ(st)] : I −→ X,e
which ends at the point [γ] ∈ X. e Thus if a loop γ in X based at x0 lifts to a loop in X e
e x̃0 )) = {e} ≤ π1 (X, x0 ). But
based at [cx0 ] then [γ] = [cx0 ], which shows that p∗ (π1 (X,
then by Lemma 3.1.14 the group π1 (X, x̃0 ) must be trivial.
e
and
e → X −→ conjugacy classes of
n o
covering maps p : X .
subgroups of π1 (X, x0 )
We wish to show that dividing out by an appropriate equivalence relation on the left
hand side turn both of these functions into bijections.
Proposition 3.4.1 (Surjectivity). Suppose that X is path connected, locally path con-
nected, and semilocally simply connected. Then for any subgroup H ≤ π1 (X, x0 ) there is
e x̃0 ) → (X, x0 ) such that p∗ (π1 (X,
a based covering map p : (X, e x̃0 )) = H.
Note that taking H to be the trivial group, this proposition implies the existence of
a universal cover. Thus it requires all the technical hypotheses of Theorem 3.3.1.
Thus ∼H is an equivalence relation. (We could have defined ∼H for any subset H of
π1 (X, x0 ), but this argument shows that such a ∼H is an equivalence relation if and only
if H is a subgroup.)
Let us define X eH := X/ ∼H to be the quotient space, and pH : X eH → X to be the
induced map. Note that if [γ] ∈ (α, U ) and [γ ] ∈ (β, U ) have [γ] ∼H [γ 0 ] then (α, U ) and
0
r : (X, [cx0 ]) −→ (X
e1 , x̃1 )
given by sending the point [γ] to the end point of the lift γ̃ of γ to X
e1 starting at x̃1 .
Now
of covering spaces over (X, x0 ). This map is also open, as both covering maps to X are
local homeomorphisms, and so q is a homeomorphism.
3.4 The Galois correspondence 27
for some [γ] ∈ π − 1(X, x0 ). If we lift γ starting at x̃1 then it ends at some x̃01 ∈ X
e1 , and
e1 , x̃0 ) → (X
Then the previous proposition gives a based homeomorphism h : (X e2 , x̃2 ).
1
Chapter 4
Let S = {sα }α∈I be a set, called the alphabet, and let S −1 = {s−1
α }α∈I (we suppose
that S ∩ S −1 = ∅). A word in the alphabet S is a (possibly empty) finite sequence
(x1 , x2 , . . . , xn )
(sα , s−1
α ) or (s−1
α , sα ).
Definition 4.1.1 (Free group). The free group on the alphabet S, F (S), is the set of
reduced (possibly empty) words in S. The group operation is given by concatenating
words and then applying elementary reductions until the word is reduced.
It is not really clear that this group operation is well-defined, or that it is associative.
It is clear though that it is unital
In Section 4.2 we give an alternative description of F (S) which is obviously a group, and
show it agrees with that of Definition 4.1.1.
By construction there is a function ι : S → F (S) given by sending the element sα to
the reduced word (sα ).
Lemma 4.1.2 (Universal property of free groups). For any group H, the function
group homomorphisms functions
−→ ,
ϕ : F (S) → H φ:S→H
28
4.1 Free groups and presentations 29
Note that if (sα11 , . . . , sαnn ) is not reduced, so contains for example (sα , s−1
α ), then the
product φ(sα1 )1 · · · φ(sαn )n contains φ(sα )·φ(sα )−1 = 1 and so we may reduce the word
(sα11 , . . . , sαnn ) without changing the value of ϕ on it. As the group operation in F (S) is
given by concatenation and reduction of words, this shows that ϕ is a homomorphism.
Definition 4.1.3 (Presentation). Let S be a set and R ⊂ F (S) be a subset. The group
hS | Ri is defined to be the quotient F (S)/hhRii of (S) by the smallest normal subgroup
containing R. More concretely,
We call hS | Ri the group with generators S and relations R, and call the data (S, R) a
presentation of this group. If S and R are both finite, we call it a finite presentation.
Lemma 4.1.4 (Universal property of group presentations). For any group H, the func-
tion
group homomorphisms functions φ : S → H such
−→ ,
ψ : hS | Ri → H that ϕ(r) = 1 for each r ∈ R
ι quot. ψ
given by sending ψ to the composition φ : S → F (S) → hS | Ri → H, is a bijection
Proof. If ψ and ψ 0 give the same functions φ = φ0 : S → H, then by Lemma 4.1.2 the
two homomorphisms
φ
quot.
/ hS | Ri *
F (S) 4H
φ0
ψ : hG | Ri −→ G
By this example every group may be given a presentation, though of course this is
quite useless for practical work.
1
or “stupid"
30 Chapter 4 Some group theory
Example 4.1.6. Let G = ha, b | ai and H = ht | i, and let us show that these groups are
isomorphic. First consider the function
φ : {a, b} −→ H
a 7−→ e
b 7−→ [t]
φ0 : {t} −→ G
t 7−→ [b]
in the group G, as a is a relation. As [a] and [b] generate G, it follows that ψ 0 ◦ ψ = IdG .
Example 4.1.7. Let G = ha, b | ab−3 , ba−2 i. In this group we have [a] = [b]3 and
[b] = [a]2 , so putting these together [a] = [a]6 , and multiplying by [a]−1 we get [a]5 = e.
Thus using [b] = [a]2 we can write every word in a and b just using the letter a, and
using [a]5 = e we may ensure that the largest power of a appearing is 4. Thus the group
G has at most 5 elements, e, [a], [a]2 , [a]3 , [a]4 . Lets guess that it might be Z/5, and
prove this.
Consider the function
φ : {a, b} −→ Z/5
a 7−→ 1
b 7−→ 2.
This homomorphism sends [a] to 1 ∈ Z/5 so is surjective, but we already argued that G
had at most 5 elements, so it is also injective: thus ψ is an isomorphism.
Fix a set S and let W be the set of reduced words in the alphabet S, and P (W ) be the
group of permutations of the set W .
4.2 Another view of free groups 31
Note that in the second case x2 6= sα , otherwise (x1 , x2 , . . . , xn ) would not be reduced.
so (x1 , x2 , . . . , xn ) = (y1 , y2 , . . . , ym ).
We can now give an alternative definition of the free group, which is manifestly a
group.
Definition 4.2.3 (Free group). The free group F (S) is the subgroup of P (W ) generated
by the elements {Lα }α∈I .
This identifies F (S) with the set of reduced words W in the alphabet S, and shows
that the group operation is given by concatenation of words followed by word reduction.
Thus the definition given in Definition 4.1.1 is indeed a group.
and so φ is surjective.
As the {Lα }α∈I generate F (S), any element σ may be represented by a concatenation
σ = Lα11 · · · Lαnn ∈ P (W ).
As Lα · L−1 −1 1 n
α = IdW and Lα · Lα = IdW , if the word (sα1 , . . . , sαn ) is not reduced then we
can simplify Lα11 · · · Lαnn while giving the same element σ ∈ P (W ). Thus we may suppose
that any σ is represented by Lα11 · · · Lαnn such that the associated word (sα11 , . . . , sαnn ) is
reduced. But then
φ(σ) = σ · () = (sα11 , . . . , sαnn ),
from which we can recover σ = Lα11 · · · Lαnn , which shows that φ is injective.
32 Chapter 4 Some group theory
G1 ∗ G2 := hS1 ∪ S2 | R1 ∪ R2 i.
Note that the free product is the free product with amalgamation for H = {e}. By
construction, the square
i1
H / G1
i2 j1
j2
G2 / G1 ∗H G2
commutes, that is j1 ◦ i1 = j2 ◦ i2 : H → G1 ∗H G2 .
Lemma 4.3.1 (Universal property of amalgamated free products). For any group K the
function
group homomorphisms group homomorphisms φ1 : G1 → K
−→ ,
φ : G1 ∗H G2 → K and φ2 : G2 → K such that φ1 ◦ i1 = φ2 ◦ i2
π1 (B, x0 ) / π1 (X, x0 )
where all the homomorphisms are those induced on the fundamental group by the natural
inclusions. By the universal property of amalgamated free products (Lemma 4.3.1) there
is a corresponding homomorphism
The Seifert–van Kampen theorem concerns conditions under which this homomorphism
is an isomorphism.
Theorem 5.1.1 (Seifert–van Kampen). Let X be a space, and A, B ⊂ X be open subsets
which cover X and such that A ∩ B is path connected. Then for any x0 ∈ A ∩ B the
homomorphism
φ : π1 (A, x0 ) ∗π1 (A∩B,x0 ) π1 (B, x0 ) −→ π1 (X, x0 )
is an isomorphism.
Before giving the proof of this theorem, we give some important examples.
Example 5.1.2 (Spheres). Consider the sphere S n for n ≥ 2. This may be covered by
the open sets
U := {(x1 , x2 , . . . , xn+1 ) ∈ S n | xn+1 > −1}
V := {(x1 , x2 , . . . , xn+1 ) ∈ S n | xn+1 < 1}
the complements of the north and south poles. Each of these is open, and stereographic
projection shows that they are homeomorphic to Rn and so contractible. Finally, cylin-
drical projection gives a homeomorphism
U ∩V ∼
= S n−1 × (−1, 1)
and as n ≥ 2 this is path connected. Thus by the Seifert–van Kampen theorem we have
π1 (S n , x0 ) ∼
= {e} ∗π1 (U ∩V,x0 ) {e}
and so π1 (S n , x0 ) ∼
= {e}. Thus the n-sphere is simply connected for n ≥ 2.
33
34 Chapter 5 The Seifert–van Kampen theorem
Example 5.1.3 (Real projective space). Recall that we have shown that the quotient
map p : S n → RPn is a covering map, with two sheets. If n ≥ 2 then S n is simply
connected, by the previous example, and so this is the universal cover of RPn . By Lemma
3.1.15 (iii), there is then a bijection between the group π1 (RPn , x0 ) and the set p−1 (x0 )
with two elements: there is only one group of cardinality two, so π1 (RPn , x0 ) = Z/2.
For based spaces (X, x0 ) and (Y, y0 ), the wedge product X ∨ Y is the quotient
space of X t Y by the relation generated by x0 ∼ y0 . This common point is taken to be
the basepoint of X ∨ Y .
Example 5.1.4 (Wedge of two circles). Let the circle S 1 ⊂ C have the basepoint 1 ∈ S 1 ,
and consider the space S 1 ∨ S 1 with basepoint x0 the wedge point. Let us take the open
cover
U := S 1 ∨ (S 1 \ {−1}) V := (S 1 \ {−1}) ∨ S 1 .
As S 1 \ {−1} deformation retracts to the basepoint 1 ∈ S 1 , notice that U deformation
retracts to S 1 ∨ {1} ∼
= S 1 , V deformation retracts to {1} ∨ S 1 , and U ∩ V = (S 1 \ {−1}) ∨
1
(S \ {−1}) deformation retracts to {x0 }. Thus by the Seifert–van Kampen theorem we
have
π1 (S 1 ∨ S 1 , x0 ) ∼
= π1 (S 1 ∨ {1}, x0 ) ∗ π1 ({1} ∨ S 1 , x0 ).
If we let a be the standard loop around the first circle and b be the standard loop
around the second circle, as shown above, we thus have
π1 (S 1 ∨ S 1 , x0 ) ∼
= ha, b | i,
This is our first example of a space with a really complicated fundamental group. Let
us apply the theory of Section 3.4 to it, to investigate some of its covering spaces.
φ : {a, b} −→ Z/3
a 7−→ 1
b 7−→ 1
By Lemma 3.1.15 (iii), the group π1 (S 1 ∨ S 1 , ∗) acting on the point x̃0 ∈ p−1 (∗)
induces a bijection
π1 (S 1 ∨ S 1 , ∗)
−→ p−1 (∗)
p∗ (π1 (X, x˜0 ))
e
where the labels on the edges show to which edge in S 1 ∨ S 1 they map to.
φ : {a, b} −→ Z/3
a 7−→ 1
b 7−→ 0
Example 5.1.7. The universal cover of S 1 ∨ S 1 may be described as the infinite 4-valent
tree (i.e. graph with no loops), where each vertex has a copy of the edge a coming in, a
copy of the edge a going out, a copy of the edge b coming in, and a copy of the edge b
36 Chapter 5 The Seifert–van Kampen theorem
going out. This certainly describes a covering space of S 1 ∨ S 1 ; it is easy to see that this
tree is simply connected, by showing that it is in fact contractible.
Proof of Theorem 5.1.1. First observe that we may assume that A and B are also path
connected: if not, there is a path component of A or B which does not intersect A ∩ B,
and so this lies in a different path component of X to x0 . Hence loops based at x0 cannot
reach it, so removing it does not change any of π1 (A, x0 ), π1 (B, x0 ), π1 (A ∩ B, x0 ) or
π1 (A ∪ B, x0 ).
The map φ is surjective. Let γ : I → X be a loop based at x0 , so {γ −1 (A), γ −1 (B)} is
an open cover of I; let n1 be a Lesbegue number for this cover. Then each path γ| i i+1
[n, n ]
has image entirely in A or entirely in B (or both). If subsequent path pieces γ| i i+1
[n, n ]
and γ| i+1 i+2 both lie in A or both lie in B then we may concatenate them. Proceeding
[ n , n ]
in this way, we may express γ as a concatenation γ1 · γ2 · · · γk where each γi (0) and each
γi (1) lie in A ∩ B.
Now for each point γ1 (1), γ2 (1), . . . , γk−1 (1) ∈ A ∩ B choose a path ui from γi (1) to
x0 in A ∩ B. Then γ is homotopic to
(γ1 · u1 ) · (u−1 −1
1 · γ2 · u2 ) · · · (uk−1 · γk )
j
This homotopy can be taken to lie in A if either the square above or below [ i−1 i
N , N]× N
is labelled A, and similarly for B, as the paths ui−1,j and ui,j then lie in the required
subspace. Call this homotopy hi,j , and write h̄i,j for the reverse homotopy.
If i − 1 = 0 or i = N the homotopy can be suitably modified, as then the path
H| i−1 i j already starts or ends at x0 .
[ N , N ]× N
This shows that each path γi in Si may be homotoped (in Si ) to be the composition of
N/n
N/n loops γi1 · γi2 · · · γi given by
N/n
and so [γi ]Si = [γi1 ]Si · · · [γi ]Si in the amalgamated free product. Thus it is enough to
show that
N/n N/n
([γ11 ]S1 · · · [γ1 ]S1 ) · ([γ21 ]S2 · · · [γ2 ]S2 ) · · · ([γn1 ]Sn · · · [γnN/n ]Sn )
Note that each column of the above maps entirely into A or entirely into B, and the
points i/N along the top or bottom all map to x0 .
Step 4. As each column of the homotopy Gj lies entirely in A or B, and has its four
vertices mapping to x0 , it follows that the loop given by the top of each column in
homotopic to the loop given by the bottom of each column by a homotopy which lies
entirely in A or entirely in B. Thus the sequence of composable loops along the bottom
and the sequence of composable loops along the top represent the same element of the
amalgamated free product. As the sequence
N/n N/n
([γ11 ]S1 · · · [γ1 ]S1 ) · ([γ21 ]S2 · · · [γ2 ]S2 ) · · · ([γn1 ]Sn · · · [γnN/n ]Sn )
5.2 The effect on the fundamental group of attaching cells 39
arises as the bottom of G0 , this argument shows that it is equivalent in the amalgamated
free product to the sequence of composable loops arising from the top of GN −1 . But by
reversing the argument of Step 2, the sequence of composable loops arising from the top
of GN −1 is a refinement of the loop H(−, t), which was constant.
i∗ : π1 (X, x0 ) −→ π1 (Y, y0 ).
U := int(Dn )
V := X ∪f (Dn \ {0}).
These satisfy U ∩ V ∼
= S n−1 × (0, 1), which is path connected as long as n ≥ 2.
Theorem 5.2.1.
(ii) If n = 2 then the map i∗ is surjective with kernel the normal subgroup generated by
[f ] ∈ π1 (X, x0 ).
is an isomorphism, but both π1 (U ∩ V, y1 ) and π1 (U, y1 ) are trivial and so the map
π1 (V, y1 ) → π1 (Y, y1 ) is an isomorphism. Hence, the change-of-basepoint isomorphism
given by u shows that π1 (V, y0 ) → π1 (Y, y0 ) is an isomorphism. Finally, the inclusion
(X, x0 ) → (V, y0 ) is a based homotopy equivalence (by pulling Dn \ {0} on to its bound-
ary), and the theorem follows.
If n = 2 then U ∩ V ' S 1 has fundamental group Z. The map
Z∼
= π1 (U ∩ V, y1 ) −→ π1 (V, y1 )
is seen to send the generator to the loop u# ([f ]). Thus, as π1 (U, y1 ) = {e}, the Seifert–
van Kampen theorem shows that
The requirement in Theorem 5.1.1 that A and B be a pair of open sets does not fit with
how one typically wants to use the theorem. In this section we shall describe weaker
hypotheses that are usually more convenient.
Theorem 5.3.2. Let X be a space, and A, B ⊂ X be closed subsets which cover X and
such that A ∩ B is both path connected and a neighbourhood deformation retract in both
A and B. Then for any x0 ∈ A ∩ B the homomorphism
is an isomorphism.
π1 (A, x0 ) −→ π1 (A ∪ V, x0 )
π1 (B, x0 ) −→ π1 (U ∪ B, x0 )
π1 (A ∩ B, x0 ) −→ π1 (U ∪ V, x0 )
Recall that for based spaces (X, x0 ) and (Y, y0 ) the wedge product X ∨ Y is the
quotient space of X t Y by the relation generated by x0 ∼ y0 . If the subspaces {x0 } ⊂ X
and {y0 } ⊂ Y are neighbourhood deformation retracts2 then it follows from Theorem
5.3.2 that
π1 (X ∨ Y, [x0 ]) ∼
= π1 (X, x0 ) ∗ π1 (Y, y0 ).
1
That is, a homotopy from IdU to a map r : U → A, which is constant when restricted to A.
2
Some authors say that a based space (X, x0 ) is well-based if {x0 } ⊂ X is a neighbourhood
deformation retract.
42 Chapter 5 The Seifert–van Kampen theorem
π1 (S 1 ∨ S 1 , [1]) ∼
= π1 (S 1 , 1) ∗ π1 (S 1 , 1) ∼
=Z∗Z∼
= ha, b | i
is a free group on two generators. By induction on n, for any finite n we thus have
n
S 1 , [1]) ∼
_
π1 ( = hx1 , x2 , . . . , xn | i,
Example 5.4.1 (Orientable surfaces). Consider the based space (X, x0 ) obtained from
the torus by removing an open triangle, as shown
ι : (S 1 ∨ S 1 , [1]) −→ (X, x0 )
of the loops a and b is a based homotopy equivalence. Thus we have π1 (X, x0 ) ∼ = ha, b | i.
The loop r given by the boundary of the open disc we removed represents the class
[a, b] := aba−1 b−1 ∈ π1 (X, x0 ), if we orient it as shown in the figure above.
Gluing g copies of X together as shown above gives a based space (Fg , f0 ). As the
gluings are performed along intervals, which are contractible and are neighbourhood
deformation retracts in X, Theorem 5.3.2 shows that there is an isomorphism
π1 (Fg , f0 ) ∼
= ha1 , b1 , a2 , b2 , . . . , ag , bg | i
5.4 The fundamental group of a surface 43
where ai and bi are the loops a and b in the ith copy of X. The boundary of Fg is a
circle, and the loop it represents is given by r1 · r2 · · · rg ∈ π1 (Fg , f0 ), the product of the
boundary loops on theQ individual copies of X. In terms of the generators ai and bi , the
boundary loop is thus gi=1 [ai , bi ]. Attaching D2 to Fg gives a closed surface Σg , and by
Theorem 5.2.1 we have
g
π1 (Σg ) ∼
Y
= ha1 , b1 , a2 , b2 , . . . , ag , bg | [ai , bi ]i.
i=1
Example 5.4.2 (The projective plane again). First recall from Example 5.1.3 that we
have shown that π1 (RP2 , x0 ) is Z/2. Let us try to understand the generator of this group.
The description of RP2 as a quotient space of S 2 also shows that it may be obtained
from D2 (thought of as the upper hemisphere of S 2 ) by dividing out by the equivalence
relation
x ∈ S 1 ⊂ D2 ∼ −x ∈ S 1 ⊂ D2
given by identifying opposite points on the boundary of D2 . If we give D2 the cell
structure shown in the figure above then we see that there is an induced cell structure
on RP2 having a single 0-cell, a single 1-cell, and a single 2-cell. If the loop around the
1-cell is called a, we see that the fundamental group of RP2 has the presentation
π1 (RP2 , x0 ) ∼
= ha | a2 i.
Example 5.4.3 (Nonorientable surfaces). Consider the based space (Y, y0 ) obtained
from RP2 by removing an open triangle, as shown
Simplicial complexes
Lemma 6.1.2. The set a0 , a1 , . . . , an ∈ Rm are affinely independent if and only if the
vectors a1 − a0 , a2 − a0 , . . . , an − a0 are linearly independent.
given by the convex hull of the points ai . The ai are called the vertices of σ, and are
said to span σ.
If x ∈ ha0 , a1 , . . . , an i then x can be written as x = ni=0 ti ai for unique real num-
P
bers ti (the uniqueness follows from the affine independence). The ti are called the
barycentric coordinates of x.
A face of a simplex σ = ha0 , a1 , . . . , an i is a simplex τ spanned by a subset of
{a0 , a1 , . . . , an }; we write τ ≤ σ, and τ < σ if τ is a proper face of σ.
The boundary of a simplex σ, written ∂σ, is the union of all of its proper faces. Its
interior, written σ̊, is σ \ ∂σ.
45
46 Chapter 6 Simplicial complexes
h : σ −→ τ
p
X p
X
ti ai 7−→ ti bi
i=0 i=0
Example 6.1.7. The standard n-simplex ∆n ⊂ Rn+1 is the simplex spanned by the
basis vectors e1 , e2 , . . . , en+1 . It, along with all its faces, defines a simplicial complex.
Example 6.1.8. The simplicial (n − 1)-sphere is the simplicial complex given by the
proper faces of ∆n , and all their faces. Its polyhedron is ∂∆n ⊂ Rn+1 .
Example 6.1.9. In Rn+1 consider the 2n+1 n-simplices given by h±e1 , ±e2 , . . . , ±en+1 i,
and let K be the simplicial complex given by these simplices and their faces. Define a
function
h : |K| ⊂ Rn+1 −→ S n
x
x 7−→
|x|
which is a bijection and continuous (it is the restriction of the map Rn+1 \ {0} → S n
given by the same formula) so is a homeomorphism (as the source is compact and the
target is Hausdorff). This defines a triangulation of S n .
Lemma 6.1.10. Any point of the polyhedron |K| lies in the interior of a unique simplex.
Proof. If x ∈ |K| = ∪σ∈KP σ then certainly x lies in some simplex σ, say with σ =
, ap i. Thus x = pi=0 ti ai . If τ ≤ σ is the face given by {ai | ti > 0} then we
ha0 , a1 , . . .P
have x = ni=0 si bi with τ = hb0 , b1 , . . . , bn i and si > 0 for all i, and so x is in τ̊ .
If x ∈ τ̊ 0 too, then τ̊ ∩ τ̊ 0 6= ∅. Thus τ ∩ τ 0 is not empty, so is a face of both τ and τ 0 .
But it contains an interior point of either simplex, so is not a proper face of either: thus
τ = τ ∩ τ 0 = τ 0.
Definition 6.1.11. For a simplicial complex K, let VK denote the set of 0-simplices of
K, which we call the vertices of K.
The reason for this perhaps surprising formulation is that the list f (a0 ), f (a1 ), . . . , f (ap )
is allowed to have repeats. Thus these elements to not need to be affinely independent,
they just need to become so when we throw away repeated elements.
f : ∆1 ⊂ R2 −→ ∆2 ⊂ R3
(1, 0) 7−→ (1, 0, 0)
(0, 1) 7−→ (0, 0, 1)
g : ∆2 ⊂ R3 −→ ∆1 ⊂ R2
(1, 0, 0) 7−→ (1, 0)
(0, 1, 0) 7−→ (1, 0)
(0, 0, 1) 7−→ (0, 1).
fσ : σ −→ |L|
p
X p
X
ti ai 7−→ ti f (ai ),
i=0 i=0
which is continuous by the gluing lemma. The formula for fσ shows that |f | behaves as
claimed under composition.
6.2 Simplicial approximation 49
Note that we can recover the simplicial map f from the continuous map |f | and the
subsets VK ⊂ |K| and VL ⊂ |L|. Thus a simplicial map from K to L could also be
defined as a continuous map |K| → |L| which sends vertices to vertices and is linear on
each simplex.
Definition 6.1.15. For a point x ∈ |K|,
(i) The (open) star of x is the union of the interiors of those simplices in K which
contain x, [
StK (x) = σ̊ ⊂ Rm .
x∈σ
The complement of StK (x) is the union of those simplices of K which do not contain
x: this is a polyhedron, so closed in |K|, so StK (x) is indeed open.
(ii) The link of x, LkK (x), is the union of those simplices of K which do not contain x,
but are a face of a simplex which does contain x. This is a subpolyhedron of |K|,
so is closed.
Figure 6.3 Three examples of points x with their open stars (in grey) and links (dotted).
for every v ∈ VK .
Lemma 6.2.2. If g is a simplicial approximation to a map f , then g defines a simplicial
map and f is homotopic to |g|. Furthermore, this homotopy may be supposed to be relative
to the subspace {x ∈ |K| |f (x) = |g|(x)} where the two maps already agree.
Proof. To show that g defines a simplicial map, for each σ ∈ K we must show that
the images under g of the vertices of σ span a simplex in L. For x ∈ σ̊ we have x ∈
∩v∈Vσ StK (v), and so
\ \
f (x) ∈ f (StK (v)) ⊂ StL (g(v)).
v∈Vσ v∈Vσ
50 Chapter 6 Simplicial complexes
Thus if τ ∈ L is the unique simplex such that f (x) ∈ τ̊ , then every g(v) is a vertex of τ ,
and so the collection {g(v) | v ∈ Vσ } span a face of τ , so span a simplex of L.
We now show that f is homotopic to |g|. If |L| ⊂ Rm , we try to define a homotopy
by the formula
H : |K| × I −→ |L|
(x, t) 7−→ t · f (x) + (1 − t) · |g|(x),
where the linear interpolation takes place in Rm . This certainly defines a continuous
map to Rm , and we must check that it lands in |L| ⊂ Rm .
Let x ∈ σ̊ ⊂ |K| and suppose that f (x) ∈ τ̊ ⊂ |L|. If σ = ha0 , a1 , . . . , ap i then by the
argument above each g(ai ) is a vetex of τ . Hence
p p
!
X X
|g|(x) = |g| ti ai = ti g(ai )
i=0 i=0
is a linear combination of vertices of τ , so lies in τ too. But then f (x) and |g|(x) both
lie in τ , so the straight line between them does too.
Figure 6.4 The barycentric subdivision of the 2-simplex ha0 , a1 , a2 i, with the simplex
hha
d \ \
2 i, ha0 , a2 i, ha0 , a1 , a2 ii indicated in grey.
It is not obvious that the collection of simplices K 0 described in this definition actually
forms a simplicial complex. To see that it does is rather involved, and we collect the
necessary arguments in the following proposition.
Proof.
If σ0 < σ1 < · · · < σp then the σ̂i are affinely independent. Suppose that
p
X p
X
ti σ̂i = 0 and ti = 0.
i=0 i=0
(i) σ 00 and τ 00 contain δ̂: let σ̄ 00 be the face of σ 00 obtained by removing δ̂, and similarly
τ̄ 00 . Then σ 00 ∩ τ 00 is the simplex spanned by σ̄ 00 ∩ τ̄ 00 and δ̂, and σ̄ 00 ∩ τ̄ 00 lies in the
boundary ∂δ. This is smaller dimensional than δ, so by induction we may assume
that σ̄ 00 ∩ τ̄ 00 is a simplex of K 0 , but then σ 00 ∩ τ 00 is too.
|K0 | = |K|. Note that hσ̂0 , σ̂1 , . . . , σ̂p i ⊆Pσp , so |K 0 | ⊆ |K|. On the other hand, for
x ∈ σ = ha0 , a1 , . . . , ap i written as x = pi=0 ti ai , we may reorder the ai so that t0 ≥
t1 ≥ · · · ≥ tp . Then
a0 + a1 a0 + a1 + a2
x = (t0 − t1 )a0 + 2(t1 − t2 ) + 3(t2 − t3 ) + ···
2 3
= (t0 − t1 )ha
d \ \
0 i + 2(t1 − t2 )ha0 , a1 i + 3(t2 − t3 )ha0 , a1 , a2 i + · · ·
D E
∈ had \ \
0 i, ha0 , a1 i, . . . , ha0 , a1 , . . . , ap i
which is a simplex of K 0 .
We now consider the following construction, which compares K and K 0 . First note
that the vertices of K 0 are precisely the barycentres of simplices of K, so there is a
52 Chapter 6 Simplicial complexes
bijection VK 0 ∼
= K. Thus if we choose a function K → VK which sends each simplex σ
to a vertex vσ of σ, then we obtain a function
g : VK 0 −→ VK
σ̂ 7−→ vσ .
If hσ̂0 , σ̂1 , . . . , σ̂p i is a simplex of K 0 then σ0 < σ1 < · · · < σp , and so all vσi are vertices
of σp , and so hvσ0 , vσ1 , . . . , vσp i is a simplex of K. Thus g defines a simplicial map.
Furthermore, if the point σ̂ lies in a simplex τ 0 = hτ̂0 , τ̂1 , . . . , τ̂p i ∈ K 0 , then σ̂ ∈ τp
and so σ is a face of τp . In particular, vσ ∈ σ ⊂ τp . Thus τ̊ 0 ⊂ τ̊p ⊂ StK (vσ ), and hence
We will use the mesh as a measure of how fine a triangulation is, and will need to
know that sufficiently many barycentric subdivisions will make the mesh as small as we
like.
n r
Lemma 6.2.6. Suppose that K has dimension at most n. Then µ(K (r) ) ≤ ( n+1 ) ·µ(K),
(r)
and so µ(K ) → 0 as r → ∞.
Proof. This will follow by indutcion from the case r = 1. Let hτ̂ , σ̂i ∈ K 0 , so τ < σ as
simplices of K. Firstly we estimate
Proof. Consider the open cover {f −1 (StL (w))}w∈VL of |K|. This has a Lesbegue number
δ (with respect to the euclidean metric on |K|), and by Lemma 6.2.6 we may choose an
r such that µ(K (r) ) < δ. Then for each v ∈ VK (r) we have
Homology
Definition 7.1.1. Let K be a simplicial complex, and On (K) be the free abelian group
with basis given by the symbols
Note that the vi are considered to be ordered, and they could span a simplex of dimension
less than n (i.e. the list could have repeats).
Let Tn (K) ≤ On (K) be the subgroup generated by the elements
(i) [v0 , v1 , . . . , vn ] if the sequence contains repeated vertices,
(ii) [v0 , v1 , . . . , vn ] − sign(σ) · [vσ(0) , vσ(1) , . . . , vσ(n) ] for a permutation σ of {0, 1, . . . , n}.
Then we define Cn (K) := On (K)/Tn (K) to be the quotient group.
Lemma 7.1.2. There is a (non-canonical) isomorphism Cn (K) ∼
= Z# n-simplices of K .
Proof. Choose a total order ≺ on VK . Then each n-simplex σ ∈ K determines a canonical
ordered simplex [σ] given by ordering the vertices ai of σ so that a0 ≺ a1 ≺ · · · ≺ an .
This defines a homomorphism
φ : Zn-simplices of K −→ On (K)
σ 7−→ [σ]
and so a homomorphism
φ0 : Zn-simplices of K −→ Cn (K)
σ 7−→ [[σ]] .
ρ : On (K) −→ Zn-simplices of K
(
sign([a0 , a1 , . . . , an ]) · ha0 , a1 , . . . , an i if there are no repeats
[a0 , a1 , . . . , an ] 7−→
0 if there are repeats.
ρ0 : Cn (K) −→ Zn-simplices of K .
54
7.1 Simplicial homology 55
where τ is such that aτ (0) ≺ aτ (1) ≺ · · · ≺ aτ (n) , but in the quotient Cn (K) this is
equivalent to [a0 , a1 , . . . , an ]. As symbols [a0 , a1 , . . . , an ] with repeats are trivial in Cn (K),
it follows that φ0 ◦ ρ0 = Id.
We now define a homomorphism
[v0 , . . . , v̂i , . . . , vj−1 , vj+1 , vj , . . . , vn ] ≡ −[v0 , . . . , v̂i , . . . , vj−1 , vj , vj+1 , . . . , vn ] mod Tn−1 (K)
[v0 , . . . , vj−1 , vj+1 , vj , . . . , v̂i , . . . vn ] ≡ −[v0 , . . . , vj−1 , vj , vj+1 , . . . , v̂i , . . . vn ] mod Tn−1 (K)
as required. This establishes that the element (7.1.1) is trivial in On−1 (K)/Tn−1 (K)
whenever σ is a transposition: as any permutation σ is a product of transpositions of
the form (j, j + 1), it follows for arbitrary σ.
Now suppose that we have [v0 , v1 , . . . , vn ] with vj = vj+1 . Then
j−1
X
dn ([v0 , v1 , . . . , vn ]) = (−1)i [v0 , . . . , v̂i , . . . , vj , vj+1 , . . . , vn ]
i=0
+ (−1)j [v0 , . . . , . . . , vj−1 , vj+1 , . . . , vn ]
+ (−1)j+1 [v0 , . . . , . . . , vj , vj+2 , . . . , vn ]
Xn
+ (−1)i [v0 , . . . , vj , vj+1 , . . . , v̂i , . . . , vn ]
i=j+2
but the middle two terms cancel out, and the outer two terms contain repeated entries
so lie in Tn−1 (K). In general, if [v0 , v1 , . . . , vn ] with vj = vk then applying σ = (j, k + 1)
reduces to the case k = j + 1.
The consequence of this lemma is that dn : On (K) → On−1 (K) induces a homomor-
phism
dn : Cn (K) −→ Cn−1 (K).
The coefficient of [v0 , . . . , v̂a , . . . v̂b , . . . , vn ] is (−1)a (−1)b from k = a and i = b plus
(−1)a (−1)b−1 from i = a and k + 1 = b. These cancel out, so dn−1 ◦ dn ([v0 , . . . , vn ]) = 0.
the symbols [v0 , . . . , vn ] generate Cn (K), so dn−1 ◦ dn = 0.
The consequence of this lemma is that we have the containment Im(dn ) ⊂ Ker(dn−1 ).
Homology is the measurement of how different these two groups are.
Example 7.1.6 (Homology of the simplicial circle). Consider the simplicial complex
K inside R3 given by all proper faces of the standard 2-simplex ∆2 ⊂ R3 . If we write
{e1 , e2 , e3 } for the standard basis of R3 , then K has simplices
he1 i, he2 i, he3 i, he1 , e2 i, he1 , e3 i, he2 , e3 i.
Choosing the order e1 ≺ e2 ≺ e3 on the vertices, we obtain isomorphisms
C0 (K) ∼
= Z{[e1 ], [e2 ], [e3 ]}
∼
C1 (K) = Z{[e1 , e2 ], [e1 , e3 ], [e2 , e3 ]},
so both groups are free abelian of rank three, and have the bases shown. With respect
to these isomorphisms, we have
d1 ([ei , ej ]) = [ej ] − [ei ],
so in the given bases d1 is given by the matrix
−1 −1 0
1 0 −1 .
0 1 1
Hence
H0 (K) = Z{[e1 ], [e2 ], [e3 ]}/h[ei ] − [ej ] | i, j ∈ {1, 2, 3}i ∼
=Z
and
H1 (K) = Ker(d1 ) = Z{[e1 , e2 ] − [e1 , e3 ] + [e2 , e3 ]} ∼
= Z.
Also Hi (K) = 0 for i ≥ 2, as K has no simplices of dimension 2 or higher.
Example 7.1.7 (Homology of the 2-simplex). Now consider the simplicial complex L
in R3 given by the standard 2-simplex ∆2 in R3 along with all its faces. The simplicial
complex L is similar to K from the previous example, but has a 2-simplex as well. Thus
we have
d2 d1
C2 (L) / C1 (L) / C0 (L)
where C1 (L) = C1 (K) and C0 (L) = C0 (K), and the formula for d1 is the same as for K.
Now d2 is injective, so H2 (L) = 0. Also Im(d2 ) = Z{[e1 , e2 ] − [e1 , e3 ] + [e2 , e3 ]} =
Ker(d1 ) and so H1 (L) = 0. Finally, we obtain H0 (L) ∼
= Z just as in the previous example.
We write Zn (C• ) := Ker(dn ), and call this the abelian group of n-cycles in C• , and
Bn (C• ) := Im(dn+1 ), and call this the abelian group of n-boundaries in C• .
A chain map f• : C• → D• is a sequence of homomorphisms fn : Cn → Dn such
that fn ◦ dn+1 = dn+1 ◦ fn+1 ; in other words, such that the square
fn+1
Cn+1 / Dn+1
dn+1 dn+1
fn
Cn / Dn
commutes.
A chain homotopy from f• to g• (two chain maps from C• to D• ) is a sequence of
homomorphisms hn : Cn → Dn+1 such that
gn − fn = dn+1 ◦ hn + hn−1 ◦ dn : Cn → Dn .
Note that in the above definitions we use the same symbol, dn , for the differential in
any chain complex. Which one we mean in any given situation will be clear from context.
f∗ : Hn (C• ) −→ Hn (D• )
[x] 7−→ [fn (x)].
(ii) If [x] = [y], then x − y ∈ Bn (C• ), and so x − y = dn+1 (z) for some z ∈ Cn+1 . Then
fn (x) − fn (y) = fn (dn+1 (z)) = dn+1 (fn+1 (z)) is a boundary, so [fn (x)] = [fn (y)] ∈
Hn (D• ).
Now suppose that f• and g• are chain homotopic via a chain homotopy h• , and let
x ∈ Zn (C• ). We have
but x is a cycle so dn (x) = 0. Thus gn (x) − fn (x) is a boundary, but then [gn (x)] =
[fn (x)].
Just as we did with the notion of homotopy of maps between spaces, we check that
(i) being chain homotopic defines an equivalence relation on the set of chain maps from
C• to D• , written f• ' g• ,
7.3 Elementary calculations 59
Returning to simplicial complexes, note that for a simplicial complex K we have defined
in Section 7.1 a chain complex C• (K), which we call the simplicial chain complex
of K. We want to show that a simplicial map f : K → L induces a chain map f• :
C• (K) → C• (L).
fn : Cn (K) −→ Cn (L)
[a0 , a1 , . . . , an ] 7−→ [f (a0 ), f (a1 ), . . . , f (an )]
Proof. We first need that fn (Tn (K)) ⊂ Tn (L), but this is clear. Then we compute
n
!
X
i
fn−1 (dn ([a0 , a1 , . . . , an ])) = fn−1 (−1) · [a0 , . . . , âi , . . . , an ]
i=0
n
X
= (−1)i · [f (a0 ), . . . , f[
(ai ), . . . , f (an )]
i=0
= dn (fn ([a0 , a1 , . . . , an ])).
Proposition 7.3.3. If K is a cone with cone point v0 , then the inclusion i : {v0 } → K
induces a chain homotopy equivalence i• : C• ({v0 }) → C• (K). Hence
(
Z{[v0 ]} n=0
Hn (K) =
0 else.
60 Chapter 7 Homology
Proof. The (only) map r : Vk → {v0 } defines a simplicial map r : K → {v0 }. We will
show that r• is a chain homotopy inverse to i• . One direction is easy: the composition
r• ◦ i• : C• ({v0 }) → C• ({v0 }) is equal to the identity map.
For the other direction, define a homomorphism
hn : On (K) −→ On+1 (K)
[a0 , . . . , an ] 7−→ [v0 , a0 , . . . , an ]
and observe that it takes Tn (K) into Tn+1 (K), and hence gives a homomorphism hn :
Cn (K) → Cn+1 (K). If n > 0 then
n
X
(hn−1 ◦ dn + dn+1 ◦ hn )([a0 , . . . , an ]) = (−1)i [v0 , a0 , . . . , âi , . . . , an ]
i=0
+ [a0 , . . . , an ]
Xn
+ (−1)i+1 [v0 , a0 , . . . , âi , . . . , an ]
i=0
= [a0 , . . . , an ] = (Id − in ◦ rn )([a0 , . . . , an ]).
If n = 0 then
(h−1 ◦ d0 + d1 ◦ h0 )([a0 ]) = d1 ([v0 , a0 ]) = [a0 ] − [v0 ]
= (Id − i0 ◦ r0 )([a0 ]).
Thus h• is a chain homotopy from i• ◦ r• to IdC• (K) .
Corollary 7.3.4. Let ∆n be the standard simplex in Rn+1 , and L be the simplicial
complex given by ∆n and all its faces. Then
(
Z i=0
Hi (L) =
0 else.
Proof. Any vertex of L is a cone point.
Corollary 7.3.5. Let K be the simplicial complex given by all the proper faces of ∆n ⊂
Rn+1 . Then for n ≥ 2 we have
(
Z i = 0 or n − 1
Hi (K) =
0 else.
Proof. Note that K is the (n − 1)-skeleton of the simplicial complex L from the previous
corollary, so Ci (K) = Ci (L) for 0 ≤ i ≤ n − 1, and Ci (K) vanishes in higher degrees.
Thus the two chain complexes are
dL dL dL dL dL
C0 (L) o C1 (L) o ··· o Cn−2 (L) o Cn−1 (L) o
1 2 n−2 n−1 n
Cn (L)
dK dK dK dK
C0 (K) o C1 (K) o ··· o Cn−2 (K) o Cn−1 (K) o
1 2 n−2 n−1
0,
7.4 The Mayer–Vietoris sequence 61
We will develop a tool akin to the Seifert–van Kampen Theorem, which will allow us to
compute the homology of a simplicial complex by decomposing it into pieces, computing
their homology, and then assembling them again using an algebraic machine.
Definition 7.4.1. We say that a pair of homomorphisms of abelian groups
f g
A −→ B −→ C
are exact at B if Im(f ) = Ker(g). More generally, we say that a collection of homo-
morphisms
· · · −→ Ai −→ Ai+1 −→ Ai+2 −→ Ai+3 −→ Ai+4 −→ · · ·
is exact if it is exact at Ai whenever there is both a map entering and a map exiting Ai .
A short exact sequence of abelian groups is an exact sequence
f g
0 −→ A −→ B −→ C −→ 0,
that is, f is injective (exactness at A), g is surjective (exactness at C), and Im(f ) =
Ker(g) (exactness at B).
Chain maps i• : A• → B• and j• : B• → C• form a short exact sequence of chain
complexes if for each n
n i jn
0 −→ An −→ Bn −→ Cn −→ 0
is a short exact sequence of abelian groups.
62 Chapter 7 Homology
• i j•
Theorem 7.4.2. If 0 → A• → B• → C• → 0 is a short exact sequence of chain
complexes, then there are natural homomorphisms ∂∗ : Hn (C• ) → Hn−1 (A• ) such that
∂∗ i∗ j∗
··· / Hn (A• ) / Hn (B• ) / Hn (C• )
∂∗
i∗ j∗
/ Hn−1 (A• ) / Hn−1 (B• ) / Hn−1 (C• ) / ···
is exact.
Let us defer the proof of this theorem.
Theorem 7.4.3 (Mayer–Vietoris). Let K be a simplicial complex, M and N be subcom-
plexes, and L = N ∩ M . Suppose that M and N cover K (i.e. every simplex of K lies
in M or N ). Let us write
L _
j
/ N
_
i l
k
M /K
for the inclusion maps between these simplicial complexes. Then there are natural homo-
morphisms ∂∗ : Hn (K) → Hn−1 (L) such that
∂∗ i∗ +j∗ k∗ −l∗
··· / Hn (L) / Hn (M ) ⊕ Hn (N ) / Hn (K)
∂∗
/ Hn−1 (L) i∗ +j∗ / Hn−1 (M ) ⊕ Hn−1 (N ) k∗ −l∗ / Hn−1 (K) / ···
i∗ +j∗ k∗ −l∗
··· / H0 (M ) ⊕ H0 (N ) / H0 (K) /0
is exact.
Proof. Note that for each n the sequence of abelian groups
in +jn n n k −l
0 −→ Cn (L) −→ Cn (M ) ⊕ Cn (N ) −→ Cn (K) −→ 0
is exact:
(i) in and jn are both injective,
(iii) if (x, y) ∈ Ker(kn − ln ) then kn (x) = ln (y), so x and y are both sums of simplices
in both M and N , so lie in L and are equal.
Thus the chain maps i• + j• : C• (L) → C• (M ) ⊕ C• (N ) and k• − l• : C• (M ) ⊕ C• (N ) →
C• (K) form a short exact sequence of chain complexes, so Theorem 7.4.2 applies, giving
the result.
7.4 The Mayer–Vietoris sequence 63
Proof of Theorem 7.4.2. The proof consists of what is affectionately known as “diagram-
chasing", that is, following the logical consequences of an element being at a certain place
in a commutative diagram.
Constructing ∂∗ (The Snake Lemma). Consider the commutative diagram
in jn
0 / An / Bn / Cn /0
d d d
in−1 jn−1
0 / An−1 / Bn−1 / Cn−1 /0
as x was a cycle. By exactness of the bottom row at Bn−1 , we thus have d(y) = in−1 (z)
for some z ∈ An−1 .
Now we have in−2 (d(z)) = d(in−1 (z)) = d(d(y)) = 0, and in−2 is injective, so d(z) =
0. Thus z representes a homology class ∂∗ ([x]) := [z] ∈ Hn−1 (A• ).
∂∗ is well-defined. If [x] = [x0 ] ∈ Hn (C• ) then x − x0 = d(a) for some a ∈ Cn+1 .
Proceeding as above for x0 , we get a y 0 ∈ Bn . As jn+1 is surjective, let b ∈ Bn+1 be such
that jn+1 (b) = a, then
In this section we want to show that to each map f : |K| → |L| we may associate a
homomorphism f∗ : Hi (K) → Hi (L), even when f does not come from a simplical map,
and further that if f ' f 0 then f∗ = f∗0 .
Definition 7.5.1. Simplicial maps f, g : K → L are contiguous if for each σ ∈ K the
simplices f (σ) and g(σ) are faces of some simplex τ ∈ L.
Lemma 7.5.2. If f, g : K → L are both simplicial approximations to the same map,
then they are contiguous.
Proof. Let F : |K| → |L| be a continuous map which f and g are simplicial approxima-
tions to. If x ∈ σ̊ ⊂ |K| and F (x) ∈ τ̊ ⊂ |L|, then as in the proof of Lemma 6.2.2 we
have that f (σ) and g(σ) are both faces of τ .
Lemma 7.5.3. If f, g : K → L are contiguous, then f• ' g• : C• (K) → C• (L) and so
f∗ = g∗ .
Proof. Choose an ordering ≺ on VK , so we can represent basis elements for Cn (K) by
ordered simplices [a0 , . . . , an ] such that a0 ≺ a1 ≺ · · · ≺ an . Define a homomorphism
hn : Cn (K) → Cn+1 (L) on basis elements by
n
X
hn ([a0 , . . . , an ]) = (−1)i [f (a0 ), . . . , f (ai ), g(ai ), . . . , g(an )].
i=0
Proof. First consider the case where K consists of a single p-simplex, σ, and all its faces.
Note that K 0 is a cone with cone point σ̂, and we have seen that K is also a cone, with any
vertex as cone point. By Proposition 7.3.3 we can thus calculate the homology of either
of these simplicial complexes, and we find that it is 0 in strictly positive degrees, and
Z in degree zero, generated by any vertex. Thus by direct computation, any simplicial
approximation to the identity a : K 0 → K induces an isomorphism on homology in every
degree.
Now let us prove the proposition by double induction on (i) the dimension of K,
and (ii) the number of simplices of K of maximal dimension. If σ is a simplex of K of
maximal dimension, then it is not a proper face of any simplex and so L := K \ {σ} is
also a simplicial complex. Let S be the simplicial complex given by σ and all its faces,
and T = S ∩ L. Any simplicial approximation to the identity a : K 0 → K sends L0 into
L and S 0 into S (so also T 0 into T ). We thus obtain a map between Mayer–Vietoris
sequences
We will now use the isomorphisms of this proposition to identify Hn (K) with Hn (K 0 ),
and so by iteration to identify Hn (K) with Hn (K (r) ). Note that by Lemmas 7.5.2
and 7.5.3 all simplicial approximations to the identity a : K 0 → K induce the same
∼
isomorphism a∗ : H∗ (K 0 ) → H∗ (K). We call this isomorphism νK , and we write νK,r
for the composition νK (r−1) ◦ · · · ◦ νK 0 ◦ νK and, for r ≥ s, νK,r,s for the composition
νK (r−1) ◦ · · · ◦ νK (s+1) ◦ νK (s) .
Proposition 7.5.6. To each continuous map f : |K| → |L| there is an associated ho-
−1
momorphism f∗ : H∗ (K) → H∗ (L) given by f∗ = s∗ ◦ νK,r , where s : K (r) → L is a
simplicial approximation to f , which exists for some r by the Simplicial Approximation
Theorem.
and hence a homomorphism Hn (K) → Hn (K̄). Identifying these groups with Hn (X)
and Hn (X̄) respectively, we obtain a homomorphism
h∗ : Hn (X) → Hn (X̄).
One can check that this is independent of choice of h-triangulations of either space, and
that if h ' h0 : X → X̄ then h∗ = h0∗ .
As a first application, we can use homology to upgrade our proof of Brouwer’s fixed point
theorem (Corollary 3.2.3) to all dimensions.
Proof. If ∆n ⊂ Rn+1 is the standard n-simplex, then its boundary ∂∆n is homeomorphic
to S n−1 . This boundary is the polyhedron of the simplicial complex of Corollary 7.3.5,
where its homology was computed.
Proof. The argument we gave in Corollary 3.2.3 shows that (i) implies (ii). To prove (i),
suppose that n ≥ 2, let r : Dn → S n−1 be a proposed retraction, and i : S n−1 → Dn be
the inclusion, and consider
i r
Z∼
= Hn−1 (S n−1 ) −→ Hn−1 (Dn ) −→ Hn−1 (S n−1 ) ∼
∗ ∗
= Z.
Recall that in Example 6.1.9 we gave a different triangulation of S n , using the sim-
plicial complex K having simplices
in Rn+1 , and all their faces. By Lemma 7.5.12 we must also have
(
∼ Z i = 0, n
Hi (K) =
0 else,
2
See Example Sheet 1, Q1
70 Chapter 7 Homology
Definition 7.7.1. A surface is a Hausdorff topological space in which each point has an
open neighbourhood homeomorphic to an open subset of R2 . A triangulated surface
is a surface X equipped with a triangulation h : |K| → X.
Example 7.7.2 (Orientable surfaces). Let K be the simplicial complex shown in Figure
7.1, which is a triangulation of the torus with an open disc removed, X.
Its boundary is identified with the boundary of a 2-simplex. We can glue g copies
of K together along an edge to obtain a simplicial complex Kg which triangulates the
surface Fg from Example 5.4.1, with boundary given by a sub-simplicial complex L. We
can then glue on the cone CL on L to obtain a triangulation of the surface Σg from
Example 5.4.1.
is such that the inclusion |W | ,→ |K| is a homotopy equivalence, and hence the inclusion
induces an isomorphism on homology. By an application of Mayer–Vietoris, decomposing
W into two circles glued together at 1, we obtain H0 (K) = Z{[1]} and H1 (K) = Z{a, b}
where
a = [1, 2] + [2, 3] + [3, 1] b = [1, 4] + [4, 7] + [7, 1],
and H2 (K) = 0. Furthermore, the boundary cycle r = [6, 5]+[9, 6]+[5, 9] of K represents
the same homology class as a + b − a − b, so the zero homology class.
0 / H2 (Kg−1 ) / H2 (Kg )
∂∗
i∗ +j∗ k∗ −l∗ /
/ H1 (∆1 ) / H1 (K) ⊕ H1 (Kg−1 ) H1 (Kg )
∂∗
/ H0 (∆1 ) i∗ +j∗ / H0 (K) ⊕ H0 (Kg−1 ) k∗ −l∗ / H0 (Kg ) /0
∂∗
i∗ +j∗ k∗ −l∗ /
/0 / Z2 ⊕ H1 (Kg−1 ) H1 (Kg )
0
/ Z i∗ +j∗ / Z ⊕ H0 (Kg−1 ) k∗ −l∗ / H0 (Kg ) /0
and so by induction
0 / H2 (Kg ∪ CL)
∂∗
/ H1 (L) i∗ +j∗ / H1 (Kg ) ⊕ H1 (CL) k∗ −l∗ / H1 (Kg ∪ CL)
∂∗
/ H0 (L) i∗ +j∗ / H0 (Kg ) ⊕ H0 (CL) k∗ −l∗ / H0 (Kg ∪ CL) /0
0 / H2 (Kg ∪ CL)
∂∗
/Z i∗ / k∗ −l∗ /
Z2g ⊕0 H1 (Kg ∪ CL)
0
/ Z (1,1) / Z ⊕ Z k∗ −l∗ / H0 (Kg ∪ CL) /0
and the map i∗ : H1 (L) → H1 (Kg ) sends the generator to the boundary cycle r1 +· · ·+rg ,
which is the zero homology class, so this map is zero. We deduce that
The group H2 (Kg ∪ CL) is generated by any 2-cycle x such that ∂∗ (x) ∈ H1 (L) ∼
= Z is
a generator.
Figure 7.2 A triangulation of Y , the projective plane with an open disc removed.
7.7 Homology of surfaces, and their classification 73
Its boundary is identified with the boundary of a 2-simplex. We can glue n copies
of K together along an edge to obtain a simplicial complex Kn which triangulates the
surface En from Example 5.4.3, with boundary given by a sub-simplicial complex L. We
can then glue on the cone CL on L to obtain a triangulation of the surface Sn from
Example 5.4.3.
is such that the inclusion |W | ,→ |K| is a homotopy equivalence, and hence the inclusion
induces an isomorphism on homology. We obtain H0 (K) = Z{[1]} and H1 (K) = Z{u}
where
u = a + b = [1, 2] + [2, 3] + [3, 1],
and H2 (K) = 0. Furthermore, the boundary cycle r = [4, 5]+[5, 6]+[6, 4] of K represents
the same homology class as 2u.
0 / 0 ⊕ H2 (Kn−1 ) / H2 (Kn )
∂∗
/ H1 (∆1 ) i∗ +j∗ / H1 (K) ⊕ H1 (Kn−1 ) k∗ −l∗ / H1 (Kn )
∂∗
/ H0 (∆1 ) i∗ +j∗ / H0 (K) ⊕ H0 (Kn−1 ) k∗ −l∗ / H0 (Kn ) /0
The boundary cycle r1 +· · ·+rn represents the same homology class as 2u1 +2u2 +· · ·+2un .
74 Chapter 7 Homology
∂∗
/ H1 (L) i∗ +j∗ / H1 (Kn ) ⊕ H1 (CL) k∗ −l∗ / H1 (Kn ∪ CL)
∂∗
/ H0 (L) i∗ +j∗ / H0 (Kn ) ⊕ H0 (CL) k∗ −l∗ / H0 (Kn ∪ CL) /0
∂∗
/Z i∗ / Zn ⊕ 0 k∗ −l∗ / H1 (Kn ∪ CL)
0
/ Z (1,1) / Z ⊕ Z k∗ −l∗ / H0 (Kn ∪ CL) /0
and the map i∗ : H1 (L) → H1 (Kg ) sends the generator to the boundary cycle r1 +· · ·+rn ,
which is the 2u1 + 2u2 + · · · + 2un . Thus this map is 1 7→ (2, 2, . . . , 2), which is injective.
We deduce that
H0 (Kn ∪ CL) = Z generated by any vertex
H1 (Kn ∪ CL) = Zn /Z{(2, 2, . . . , 2)} generated by u1 , . . . , un subject to 2u1 + · · · + 2un = 0
H2 (Kn ∪ CL) = 0.
Note that the change of basis
e1 = u1 + u2 + · · · + un ei = ui 2 ≤ i ≤ n
gives an isomorphism of abelian groups Zn /Z{(2, 2, . . . , 2)} ∼
= Z/2 ⊕ Zn−1 .
To summarise these two examples, as well as Lemma 7.6.1 in dimension 2, the (tri-
angulable) surfaces S 2 , Σg for g ≥ 1, and Sn for n ≥ 1, all have non-isomorphic first
homology groups:
H1 (S 2 ) = 0 H1 (Σg ) ∼
= Z2g H1 (Sn ) ∼
= Z/2 ⊕ Zn−1 .
Thus no two of these surfaces are homeomorphic, or even homotopy equivalent. This
proves one half of the following theorem. The other half is much more elementary, but
quite long: it can be found for example in Maunder’s Algebraic Topology.
Theorem 7.7.4 (Classification of triangulable surfaces). Any triangulable surface is
homeomorphic to one of
S 2 , Σ1 , Σ2 , Σ3 , . . .
S1 , S2 , S3 , . . .
and no two of these are homeomorphic.
7.8 Rational homology, Euler and Lefschetz numbers 75
We have been using the terms “orientable surface” for one obtained by gluing together
copies of the torus, and “nonorientable surface” for one obtained by gluing together copies
of the projective plane. These are in fact intrinsic properties of surfaces, and not just a
feature of how we choose to construct them: this may be seen homologically by
This difference accounts for our partition in Theorem 7.7.4 of surfaces into two series,
the orientable ones (S 2 , Σ1 , Σ2 , Σ3 , . . .) having second homology Z, and the nonorientable
ones (S1 , S2 , S3 , . . .) having second homology 0.
Definition 7.8.1. For a simplicial complex K, define the rational n-chains, Cn (K; Q),
by letting On (K; Q) be the Q-vector space with basis the oriented simplices, and Tn (K; Q)
be the sub vector space spanned by the usual collection of “trivial” simplices. We define
dn : Cn (K; Q) → Cn−1 (K; Q) by the usual formula, and write
Everything we did for ordinary homology remains true for rational homology (induced
maps, subdivision isomorphism, homotopy invariance, Mayer–Vietoris sequence). The
following lemma gives some idea of what rational homology measures. Its proof is Q8 on
Example Sheet 4.
We thus know the homology of several spaces. For spheres, with n > 0,
(
Q i = 0, n
Hi (S n ; Q) ∼
=
0 else.
Example 7.8.4. Computing the Euler characteristic for the examples we have gives
(
n 2 n even
χ(S ) = χ(Σg ) = 2 − 2g χ(Sn ) = 2 − n.
0 n odd
where X is the matrix for B with respect to the basis e1 , e2 , . . ., er and Z is the matrix
for C with respect to the basis [er+1 ], . . . , [en ]. Now Tr X0 YZ = Tr(X) + Tr(Z), as
required.
and
0 −→ Zi (K; Q) −→ Ci (L; Q) −→ Bi−1 (K; Q) −→ 0.
7.8 Rational homology, Euler and Lefschetz numbers 77
Let fiH , fiB , fiZ , and fiC be the maps induced by f• on homology, boundaries, cycles,
and chains respectively. Then
X
L(|f |) = (−1)i Tr(fiH )
i≥0
X
(−1)i Tr(fiZ ) − Tr(fiB )
=
i≥0
X
(−1)i Tr(fiC ) − Tr(fi−1
B
) − Tr(fiB )
=
i≥0
X
= (−1)i Tr(fiC ).
i≥0
(r) (r)
If σ ∈ K is an i-simplex then si (σ) is a sum of simplices inside σ, and so gi (si (σ)) is
(r)
a sum of simplices disjoint from σ. Thus the matrix for gi ◦ si with respect to the basis
of simplices has zeroes on the diagonal, so has trace zero.
Example 7.8.10. If X is a contractible polyhedron then L(f ) = 1 for any f : X → X,
so any f has a fixed point. This recovers Brouwer’s fixed point theorem, but is far more
general.
Example 7.8.11. let h : |K| → G be a triangulation of a topological group G which is
connected and nontrivial. The for g 6= 1 ∈ G multiplication by g is a homeomorphism
with no fixed points, so 0 = L(− · g). On the other hand, choosing a path from g to
1 ∈ G gives a homotopy from − · g to the identity, so
0 = L(− · g) = L(− · 1) = L(IdG ) = χ(G).
78 Chapter 7 Homology
Among the surfaces, only Σ1 and S2 have Euler characteristic zero, so no other
surface admits a topological group structure. The torus Σ1 = S 1 × S 1 does admit a
group structure, as R2 /Z2 . The surface S2 is the Klein bottle, and does not admit a
group structure. One way to see this is to use Q13 on Example Sheet 1, where the
fundamental group of the Klein bottle is computed and shown to be non-abelian: it is
an easy exercise to show that the fundamental group of a topological group (based at
the identity element) is always abelian.
and f∗ : H0 (S3 ; Q) → H0 (S3 ; Q) is the identity map, as this group is generated by any
vertex of a triangulation of S3 . The linear map
f∗ : Q2 ∼
= H1 (S3 ; Q) −→ H1 (S3 ; Q) ∼
= Q2
satisfies (f∗ )2 = Id, so by elementary linear algebra its eigenvalues are all ±1. Thus its
trace is one of −2, 0, or 2, and so