Algebraic Topology by Tim Perutz
Algebraic Topology by Tim Perutz
Algebraic Topology by Tim Perutz
TIM PERUTZ
Contents
I. The fundamental group
1. Introduction
1.1. Homotopy equivalence
1.2. The fundamental groupoid
2. The fundamental group of the circle
2.1. Trivial loops
2.2. Computing 1 (S 1 )
2.3. Applications
3. Van Kampen in theory
3.1. Group presentations
3.2. Push-outs
3.3. Van Kampens theorem
4. Van Kampen in practice
4.1. Fundamental groups of spheres
4.2. A useful lemma
4.3. Fundamental groups of compact surfaces
4.4. The complement of a trefoil knot
5. Covering spaces
5.1. Deck transformations
5.2. Examples
5.3. Unique path lifting
6. Classifying covering spaces
6.1. An equivalence of categories
6.2. Existence of a simply connected covering space
II. Singular homology theory
7. Singular homology
7.1. The definition
7.2. The zeroth homology group
7.3. The first homology group
8. Simplicial complexes and singular homology
8.1. -complexes
8.2. The Hurewicz map revisited
9. Homological algebra
9.1. Exact sequences
9.2. Chain complexes
10. Homotopy invariance of singular homology
11. The locality property of singular chains
12. MayerVietoris and the homology of spheres
1
3
3
3
4
6
6
6
7
9
9
9
11
13
13
13
13
15
17
18
18
19
21
21
23
25
25
25
26
26
28
28
29
30
30
31
34
37
39
TIM PERUTZ
39
41
42
42
43
44
45
46
46
49
49
49
49
50
50
53
53
53
55
57
57
59
59
59
61
64
64
67
70
70
71
74
74
74
75
77
78
78
80
80
80
81
83
84
TIM PERUTZ
Convex subsets of Rn are contractible for a particular reason: their points are
deformation retracts. In general, if X is a space and i : A X the inclusion of a
subspace, we say that A is a deformation retract of X if there is a map r : X A
such that r i = idA and i r ' idX by a homotopy {ht } so that (in addition to
h0 = i r and h1 = idX ) one has ht (a) = a for all t and a A. Such a map r,
called a deformation retraction, is obviously a homotopy equivalence.
Exercise 1.4: Show carefully that the letter A, considered as a union of closed line
segments in R2 , is homotopy equivalent but not homeomorphic to the letter O. Show
briefly that all but one of the capital letters of the alphabet is either contractible or
deformation-retracts to a subspace homeomorphic to O. Show that the letters fall into
exactly three homotopy types. How many homeomorphism types are there? (View a
letter as a finite union of the images of paths [0, 1] R2 . Choose a typeface!)
Exercise 1.5: Let {X }A be a collection of spaces indexed
by a set A. Let x X
W
be basepoints. Define `
the wedge sum (or 1-point union) A X as the quotient space
of the disjoint union X by the equivalence relation x x for all , A.
Show carefully that, for n 1, the complement of p distinct points in Rn is homotopyequivalent to the wedge sum of p copies of the sphere S n1 = {x Rn : |x| = 1}.
Remark. Lets look ahead. Theorems of Hurewicz and J. H. C. Whitehead imply
that, among all spaces which are cell complexes, the sphere S n = {x Rn+1 :
|x| = 1}, with n > 1, is characterized up to homotopy equivalence by its homology
/ {0, 1} and its trivial fundamental
groups H0 (S n )
= Z, Hi (S n ) = 0 for i
= Hn (S n )
group. In general, distinct homotopy types can have trivial fundamental groups and
isomorphic homology groups (e.g. S 2 S 2 , CP 2 #CP 2 ). Another invariant, the
cohomology ring, distinguishes these two examples. When it fails to distinguish
spaces, one localizes the problem and works over Q and mod primes p. Over Q, a
certain commutative differential graded algebra gives a new invariant [D. Sullivan,
Infinitesimal computations in topology, Publ. Math. I.H.E.S. (1977)]. Mod p, one
considers the Steenrod operations on cohomology. There is an algebraic structure
which captures all this at once, and gives a complete invariant for the homotopy
type of cell complexes with trivial fundamental group [M. Mandell, Cochains and
homotopy type, Publ. Math. I.H.E.S. (2006)].
1.2. The fundamental groupoid.
1.2.1. Our first invariants of homotopy type are the fundamental groupoid and the
isomorphism class of the fundamental group.
A path in a space X is a map f : I X, where I = [0, 1]. Two paths f0 and f1
are homotopic rel endpoints if there is a homotopy {ft }t[0,1] between them such
that ft (0) and ft (1) are both independent of t. Write for the equivalence relation
of homotopy rel endpoints.
Two paths f and g are composable if f (1) = g(0). In this case, their composite
f g is the result of traversing first f , then g, both at double speed: (f g)(t) = f (2t)
for t [0, 1/2] and (f g)(t) = g(2t 1) for t [1/2, 1].
The composition operation is not associative: (f g) h 6= f (g h). What is
true, however, is that (f g) h f (g h). (Proof by picture.)
If f is a path, let f 1 denote the reversed path: f 1 (t) = f (1 t). One has
f f 1 cf (0) and f 1 f cf (1) , where cx denotes the constant path at x.
(Picture.) Moreover, cf (0) f ' f and f cf (1) ' f .
[] 7 [f ] [] [f 1 ]
is an isomorphism.
If F : X Y is a map, there is an induced homomorphism
F : 1 (X, x) 1 (Y, F (x)),
[f ] 7 [F f ].
TIM PERUTZ
p1
[t + ) Uf (t) Vf0(t) .
This does indeed define an extension of f to [0, t + ). So T is an open set.
Now suppose f exists and is unique on [0, t). Since f (s) f (t) as s t,
when 0 < t s 1 the lifts f(s) must lie in one of the open sets V projecting
TIM PERUTZ
D2n
= ha, b | an , b2 , (ba)2 i,
Z2
= ha, b | aba1 b1 i.
3.2. Push-outs.
Definition 3.3. Consider three groups, G1 , G2 and H, and a pair of homomorphisms
f1
f2
G1 H G2 .
10
TIM PERUTZ
f2 y
G1
p1
y
p2
G2 P
and satisfying a universal property: given any other such square (a group K and
homomorphisms k1 : G1 K and k2 : G2 K such that k1 f1 = k2 f2 ), there
is a unique homomorphism h : P K such that k1 = h p1 and k2 = h p2 .
Exercise 3.1: Prove that the universal property determines P up to isomorphism. In
what sense is the isomorphism unique?
We can understand push-outs concretely using group presentations. Suppose
G1 = ha1 , . . . , an | r1 , . . . rm i and that G2 = hb1 , . . . , bp | s1 , . . . , sq i. Also suppose
that H has generators h1 , . . . , ho . In the push-out square above, the group P is then
a group called the free product of G1 and G2 amalgamated along H and notated
G1 H G2 . It has the presentation
G1 H G2 = ha1 , . . . , an , b1 , . . . , bp | r1 , . . . rm , s1 , . . . , sq , c1 , . . . , co i,
where ci = f1 (hi )f2 (hi )1 .
Exercise 3.2: Check that K = G1 H G2 fits into a push-out square for f1 and f2 .
Note that there was no need for our group presentations to be finite, except for
notational convenience: we can allow infinite sets of generators and relations.
Exercise 3.3: The free product G1 G2 of groups G1 and G2 is the push-out of the
diagram G1 {1} G2 . Define D as the subgroup of the group of affine transformations R R generated by x 7 x and x 7 x+1. Prove that D
= (Z/2)(Z/2).
Exercise 3.4: In this exercise we show that the modular group, P SL2 (Z) = SL2 (Z)/{I},
is the free product (Z/2) (Z/3). Define three elements of SL2 (Z),
0 1
1 1
0 1
S=
, T =
, U = ST =
.
1 0
0
1
1 1
(a) Verify that S 2 = U 3 = I.
(b)
Show that,
for any A SL2 (Z), there is an n Z such that the matrix
a b
= AT n has c = 0 or |d| |c|/2.
c d
(c) Explain how to find an integer l 0 and a sequence of integers n1 , . . . , nl
such that either AT n1 ST n2 S . . . ST l or AT n1 ST n2 S . . . ST l S has 0 as its
lower-left entry.
(d) Show that S and T generate SL2 (Z).
(e)* Define : ha, b | a2 , b3 i = (Z/2) (Z/3) P SL2 (Z) to be the unique
homomorphism such that (a) = S and (b) = U . Remind yourself
how P SL2 (R) acts on the upper half-plane H C by Mobius maps. Take
1 6= w (Z/2) (Z/3). Prove that the Mobius map w corresponding to
(w) P SL2 (R) has the property that w (D) D = , where
D = {z H : 0 < Re z < 1/2, |z 1| > 1}.
[Hint: consider A := {z H : Re z > 0} and B := {z H : |z 1| >
max(1, |z|)}.] Deduce that is an isomorphism.
11
y
y
1 (V, x) 1 (X, x)
of maps induced by the inclusions is a push-out square.
Example 3.5. Let Cn be the complementWof n points in the plane. Observe that
n
Cn deformation-retracts
to the wedge sum i=1 S 1 . We have 1 (Cn )
= Fn . Indeed,
Wn
1
when n > 0, i=1 S is the union of a subspace U which deformation-retracts to
Wn1 1
1
i=1 S , and a subspace V which deformation-retracts to S , where the subspace
U V is contractible. By induction, 1 (U )
F
.
We
know
1 (V )
= n1
= Z = F1 .
The push-out of Fn1 and Z along the trivial group H is Fn1 F1
F
= n . Thus the
result follows from van Kampens theorem.
Lemma 3.6. For any loop : (I, I) (X, x), there exists a finite, strictly increasing sequence 0 = s0 < s1 < s2 < < sn = 1 such that maps each interval
[si , si+1 ] into U or into V .
Proof. Every x I has a connected open neighbourhood whose closure maps to U
or to V . Since I is compact, finitely many of these intervals cover I. The endpoints
of the intervals in the finite cover form a finite subset of I, which we may enumerate
in ascending order as (s0 , . . . , sn ).
Let us call the sequence (si ) a subdivison for .
Lemma 3.7. Suppose = {t }t[0,1] is a homotopy of paths (I, I) (X, x).
Then there are increasing sequences 0 = t0 < t1 < < tm = 1, and 0 = s0 <
< sn = 1, such that maps each rectangle [ti , ti+1 ] [sj , sj+1 ] into U or into
V . Moreover, we can take the sequence (si ) to refine given subdivisions of 0 and
1 .
Exercise 3.5: Prove the lemma.
Proof of Van Kampens theorem. Suppose we are given a group G and homomorphisms f : 1 (U ) G, g : 1 (V ) G which agree on the images of 1 (U V ). We
construct a map : 1 (X) G so that f = (iU ) and g = (iV ) , where
iU : U X and iV : V X are the inclusions.
Take : (I, I) (X, x), and choose a subdivision s0 < < sn . Label the
intervals [si , si+1 ] as red or blue, in such a way that maps red intervals to U
and blue intervals to V . For 0 < i < n, connect (si ) to x by a path i inside
U (if both adjacent intervals [si1 , si ] and [si , si+1 ] are red), inside V (if both
adjacent intervals are blue), or inside U V (if the adjacent intervals are different
colours). Then i := i1 |[si ,si+1 ] i is a loop in either U or V . Define [] =
1 [0 ] n1 [n1 ], where i is either f or g according to whether [si , si+1 ] is
red or blue.
We need to see that is well-defined, and does not depend on the choices of
path, subdivision and colouring. Observe that for a fixed and fixed subdivision,
12
TIM PERUTZ
changing the colouring does not affect , because f and g agree on the image of
1 (U V ). Moreover, refining a subdivision for given does not affect the definition
of . Nor does changing the choice of a path i (instead of trying to replace i by a
rival path i0 , insert an extra point into the subdivision, and use both paths i and
i0 ).
Hence we are left with considering homotopic paths 0 and 1 with a common
subdivision s0 < < sn .
Given a homotopy = {t }, we can subdivide [0, 1] [0, 1] into rectangles
Rij = [ti , ti+1 ] [sj , sj+1 ] and color the Rij as red or blue in such a way so that
maps the red rectangles to U and the blue ones to V . It will suffice to show that
0 and t1 give the same definition for .
This last part of the argument requires pictures, which I will draw in class.
(Consult Hatcher if you need to.) The idea is this: rather than going along the
bottom edge of R0j we can go around the other three sides. We have 0 ' 0 n ,
but by going round these three sides we can replace i by a new loop i0 , and this
will not affect . By eliminating backtracking we can get from 0 n to t1 ,
again without affecting .
Knowing it is well-defined, one can check that is a homomorphism making the
two triangles commute (do so!). Note also that it is the unique such homomorphism:
since is homotopic to the composite of the i |[si ,si+1 ] i1 , we have no choice
but to define this way. This concludes the proof.
13
p
y
{1} P
is a pushout square. Then p is surjective, and its kernel is the normalizer of im f .
Proof. Put P 0 = G/N , where N is the normalizer of im f , and define p0 : G P 0
to be the quotient map. It is easy to check that P 0 and p0 fit into a push-out square
for the homomorphisms f : H G and H {1}. Thus P is isomorphic to P 0 so
that p is identified with p0 .
In conjunction with van Kampens theorem, this lemma has the following consequence.
Proposition 4.3. Suppose that X is the union of a path-connected open set U and
a simply connected open set V , with U V path-connected. Let x U V . Then
1 (X, x) is generated by loops in U . A based loop in U becomes trivial in 1 (X) iff
it lies in the normal subgroup of 1 (U, x) generated by loops in U V .
4.3. Fundamental groups of compact surfaces.
Proposition 4.4. Let T 2 be the 2-torus, RP 2 the real projective plane, K 2 the
Klein bottle. Then
1 (T 2 )
= Z2 ; 1 (RP 2 )
= Z/2; 1 (K 2 )
= ha, b | aba1 bi.
No two of these spaces are homotopy-equivalent.
Proof. These spaces X are all quotient spaces q : I 2 X of the square I 2 R2 ,
obtained by gluing together its sides in pairs. Take p in int(I 2 ).
Let U = q(I 2 \ {p}), and V = q(D) with D a small open disc containing p.
Thus U V deformation-retracts to a circle and V is simply connected. By the last
proposition, 1 (X) is generated by loops in the subspace U , which deformationretracts to q(I 2 ).
Going anticlockwise round I 2 , we label the sides as s1 , s2 , s3 , s4 (as directed
paths).
14
TIM PERUTZ
1
In T 2 , q(s1 ) = q(s1
3 ) and q(s2 ) = q(s4 ). Thus U deformation-retracts to
a wedge of two circles a = q(s1 ) and b = q(s2 ), and U ' s1 s2 s3 s4 '
a b a1 b1 . To apply van Kampen, note that 1 (U V ) = Z and 1 (U ) = ha, bi.
The homomorphism Z F2 induced by U V , U sends 1 to aba1 b1 . Thus,
by the last proposition,
1 (T 2 )
= ha, b | aba1 b1 i
= Z2 .
n y
y n1
Gn
gn
Gn1 ,
15
[ag , bg ] = 01 n+1
from U and V . It is easy to check that this system of generators and relations are
equivalent to those given.
Another standard way to prove this is to think of g as an identification-space
of the 4g-gon.
4.4. The complement of a trefoil knot. The left-handed trefoil knot K is the
image of the embedding f : S 1 S 3 = {(z, w) C2 : |z|2 + |w|2 = 1} given by
1
1
f (e2it ) = ( e4it , e6it ).
2
2
3
2 3
Z Z Z,
and this gives the presentation claimed.
16
TIM PERUTZ
Exercise 4.3: The braid group on 3 strings. In this extended exercise (based on one in
Serres book Trees) well see that the following five groups are isomorphic:
1 (S 3 \ K), where K is a (left-handed) trefoil knot.
The group ha, b | a2 = b3 i.
The algebraic braid group on 3 strings, hs, t | sts = tsti.
The geometric braid group on 3 strings B3 , defined as the fundamental group
of the configuration space C3 of 3-element subsets of C.
1 (C2 \ C), where C C2 is the cuspidal cubic {(X, Y ) : X 2 = Y 3 }.
(a) We already know that 1 (S 3 \ K)
= ha, b | a2 = b3 i. Show that a 7 sts,
b 7 ts defines an isomorphism
ha, b | a2 = b3 i hs, t | sts = tsti.
(b) Take as basepoint {2, 0, 2} C3 . Define loops and in X3 , (t) =
{1 eit , 1 + eit , 2} and (t) = {2, 1 eit , 1 + eit } for t [0, 1].
Let s = [] and t = [ ] in B3 . Check that sts = tst, so that one has a
homomorphism hs, t | sts = tsti B3 .
(b) C3 is the subspace of Sym3 (C) (the quotient of C3 by the action of the
symmetric group S3 permuting coordinates) where the three points are distinct. Let Sym30 (C) = {{a, b, c} Sym3 (C) : a + b + c = 0}. Show that
Sym3 (C)
= C Sym30 (C). Define a homeomorphism h : Sym30 (C) C2 by
sending {a, b, c} to the point (x, y) such that
(t a)(t b)(t c) t3 + xt + y.
Verify that the points a, b and c are distinct iff 4x3 + 27y 2 6= 0. Deduce that
C3
= C (C2 \ C), hence that B3
= 1 (C2 \ C).
2
(d) Show that C \ C is homotopy-equivalent to S 3 \ K, whence 1 (C2 \ C)
=
1 (S 3 \ K).
(e)* Show that going round the full circle of homomorphisms, the resulting homomorphism 1 (S 3 \ K) 1 (S 3 \ K) is an isomorphism.
17
5. Covering spaces
Another basic method of computing fundamental groups is to identify the space
18
TIM PERUTZ
19
(, ) 7
.
(0, 0).
x
(3) The map p : 1 (X,
) 1 (X, x) is injective.
0
x
x
(4) If x
also lies in p1 (x) then p (1 (X,
0 )) and p (1 (X,
)) are conjugate
subgroups of 1 (X, x).
x
(5) All conjugates of p (1 (X,
)) arise in this way.
Proof. (1) The proof is exactly the same as the proof of unique path lifting for
R S 1 that we gave in our proof that 1 (S 1 ) = Z. Similarly (2).
(3) If p (0 ) and p (1 ) are homotopic rel endpoints then the unique lift of the
defines a homotopy rel endpoints between 0 and 1 .
homotopy to X
joining x
(4) Choose a path in X
0 to x
. Then we have
0
x
p (1 (X, x
)) = (p ) p (1 (X,
)) (p )1 .
(5) Follows from (1).
Example 5.8. If X
=
2i/n
in
Z/n (the generator being multiplication by e
), while the image of 1 (X)
in 1 (X) is
1 (X) is nZ Z. We see in this example that the image of 1 (X)
a subgroup whose index is equal to the number of sheets of the covering. It is a
20
TIM PERUTZ
X
= Z and
21
Corollary 6.4. If p : X
acts freely and tran1
22
TIM PERUTZ
starting at x
defines an isomorphism Ix : G Aut(X/X)
where G = 1 (X, x).
x (H). It comes with a projection map
Take H 1 (X, x), and define G(H) = X/I
X, and this is certainly a covering. Its fibre over x
G(H) X, induced by p : X
is canonically identified with G/H, and 1 (G(X), [
x]) maps to H under the covering
map.
Every path-connected covering Y X is isomorphic to G(H) for some H.
Indeed, we take H to be the image of 1 (Y, y) in 1 (X, x) for some y lying over x,
cf. Corollary 6.2.
If K is another subgroup, and f : G/H G/K a map of G-sets, let f (H) = K.
Then, for all g G, we have f (gH) = gK. Notice that if h H then f (hH) =
hK = K, hence 1 H K; conversely, an element such that 1 H K
defines a map of G-sets.
x (H)
We shall define G(f ) via the lifting criterion. We are looking for a map X/I
X/Ix (K) covering the identity on X. Such a map will be unique once we specify its
x (H) such that the image
effect on a point. For existence, take a basepoint z X/I
x (H), z) in G is 1 H (cf. Lemma 5.7, (5)). By the lifting criterion,
of 1 (X/I
x (H) X/I
x (K) which sends z to
there is a unique map of covering spaces X/I
[
x]. This is G(f ). Its straightforward to check this gives a functor.
It remains to see that G gives a bijection between morphism sets. This is another
application of the lifting criterion, but we omit the details.
Let us spell out some aspects of this correspondence.
At one extreme, we can consider the trivial subgroup {1} G, which
corresponds to the universal cover. At the other extreme, G G gives the
trivial cover X X.
In general, the fibre of the covering G(H) corresponding to H G is G/H.
Thus finite index subgroups correspond to coverings with finite fibres.
23
the basepoint [
x] coming from x
X.)
The normal (or regular, or Galois) coverings of X are those coverings
q : Y X for which q 1 (Y ) is a normal subgroup of G. Equivalently,
Aut(Y /X) acts transitively on the fibre. A normal covering determines an
actual subgroup, not just a conjugacy class of subgroups.
The similarity of the classification theorem with the fundamental theorem of Galois
theory is not coincidental; the theory of etale maps in algebraic geometry unites
them. In particular, finite extensions of the function field K(X) of a variety X
correspond to finite (etale) coverings of X.
6.2. Existence of a simply connected covering space. Under very mild hypotheses, a simply connected covering exists. Assume X locally path connected.
X from a simply
Proposition 6.8. Suppose that X admits a covering map p : X
Then X is semi-locally simply connected, meaning that each
connected space X.
x X has a path-connected neighbourhood U such that im(1 (U ) 1 (X)) is
trivial.
Proof. Let U be a neighbourhood over which p is trivial. Then any loop in U
which is nullhomotopic (rel I). Projecting the nullhomotopy
lifts to a loop in X,
to X, we see that is nullhomotopic in X.
Exercise 6.1: Find a path connected, locally path connected space which is not semilocally simply connected.
as the set of homotopy classes [], where
Now fix a basepoint x X. Define X
X to
: I X with (0) = x and [] its homotopy class rel I. Define p : X
ought to be generated by
be the evaluation map [] 7 (1). The topology on X
the path components of the sets p1 (V ) with V X open. Path components do
not make sense a priori, but we can make sense of them, via path-lifting, when V
is path connected and im(1 (V ) 1 (X)) is trivial.
Proposition 6.9. If X is path-connected, locally path-connected and semi-locally
X is a covering map and X
is simply connected.
simply connected then p : X
Thus X admits a simply connected covering space.
more precise, then prove the proposition.
Exercise 6.2: Make the topology on X
Exercise 6.3: (From Mays book.) Identify all index 2 subgroups of the free group F2 .
Show that they are all free groups and identify generators for them.
Exercise 6.4: (a) The universal cover of the torus T 2 is R2 . Identify all the deck
transformations and hence determine (once again) the fundamental group. Which
surfaces can cover T 2 ? (b) Show that the Klein bottle is also covered by R2 ; identify
the deck transformations and hence the fundamental group.
Exercise 6.5: Let p : Y X be a covering (with Y path connected and X locally path
24
TIM PERUTZ
Exercise 6.6: Define a regular tetrahedron as a set of four distinct, unordered, equidistant points on S 2 R3 . Let T be the space of regular tetrahedra. (a) Show that
1 (T ) has a central subgroup Z
= Z/2 such that 1 (T )/Z
= A4 . (b) Identify several
(at least 5) pairwise non-isomorphic, path connected covering spaces of T , describing
them geometrically. (c) Show that the fundamental group of the space P of regular
icosahedra (unordered collections of 20 distinct points on S 2 forming the vertices of a
regular icosahedron) has order 120, but that the abelianization 1 (P)ab has order at
most 2. (In fact it is trivial.) [Recall that the icosahedral group A5 is simple.]
Exercise 6.7: Rotation about a fixed axis, by angles increasing from 0 up to 2, determines a loop in SO(3). Show that is nullhomotopic.
25
xi = 1}.
In alternative notation, n =
Pn
i=0
j < i.
0j<in+1
=
=
0kln
=0.
(1)i+j i j
0ijn
(1)
i+j
j i1 +
0j<in+1
(1)i+j i j +
(1)k+l+1 k l +
(1)i+j i j
0ijn
X
0ijn
(1)i+j i j
26
TIM PERUTZ
It is convenient to let 0 : S0 (X) 0 be the zero-map. The nth singular homology of X is the abelian group
Hn (X) := ker n / im n+1 .
Elements of ker n are called n-cycles; elements of im n+1 are n-boundaries. By
the lemma, an n-boundary is an n-cycle, and the nth homology is the group of
n-cycles modulo n-boundaries.
In future lectures we will develop these groups systematically. Today we will
look only at the zeroth and first homology groups.
7.2. The zeroth homology group.
P
P
Proposition 7.2. The map : S0 (X) Z, ( ni i ) = ni induces a surjection
H0 (X) Z provided only that X is non-empty. When X is path-connected, this
map is an isomorphism.
Proof. We have to show that descends to H0 (X) = S0 (X)/ im 1 . If is a 1simplex then = 0 1 . Thus ( ) = 1 1 = 0. Hence (im 1 ) = 0,
and descends to H0 (X). For
P any 0-simplex , (n) = n, so is surjective. If X
is path-connected, take s = ni i ker . We may assume ni = 1 for all i. The
number of + and signs is equal, so we may partition the 0-simplices into pairs
(i , j ) with ni = 1 and nj = 1. But i j is the boundary of a 1-simplex (i.e.,
of a path), since X is path-connected. Hence s im 1 .
L
Exercise 7.1: Show that, in general, Hn (X) = Y 0 (X) Hn (Y ), where 0 (X) is the
Z0 (X) .
set of path-components of X. Thus H0 (X) =
So, whilst S0 (X) is typically very large (often uncountably generated), H0 (X)
is finitely generated for all compact spaces.
7.3. The first homology group. Theres a homeomorphism I 1 given by
t 7 tv1 +(1t)v0 . Thus a path : I X defines a 1-simplex . When (0) = (1),
is a 1-cycle.
Lemma 7.3. Fix a basepoint x X. The map 7 induces a homomorphism
h : 1 (X, x) H1 (X).
Proof. A constant loop is the boundary of a constant 2-simplex. Loops which are
homotopic rel endpoints give homologous 1-simplices (by subdividing a square into
two triangles and using the fact that constant loops are boundaries). Thus h is
well-defined. If f and g are composable paths, the composition f g maps under
define a 2-simplex = (f g) p : 2 X, where p is the projection
h to f + h:
[v0 , v1 , v2 ] [v0 , v2 ], t0 v0 + t1 v1 + t2 v2 7 t1 v1 + t2 v2 . We have = g f[
g + f.
The map h is sometimes called the Hurewicz map.
Proposition 7.4. The kernel of the Hurewicz map h : = 1 (X, x) H1 (X)
contains the commutator subgroup [, ], and hence h induces a homomorphism
ab := /[, ] H1 (X).
When X is path-connected, h is surjective.
27
28
TIM PERUTZ
simp
ker(n : Snsimp (X) Sn1
(X))
simp
im(n+1 : Sn+1
(X) Snsimp (X)
This comes with a natural homomorphism Hnsimp (X) Hn (X), induced by the
inclusion Snsimp (X) Sn (X).
Exercise 8.1: Think of S 2 as a tetrahedron, i.e., a -complex with four 2-simplices, six
1-simplices and four 0-simplices. Show that for this structure
H0simp (S 2 ) = Z,
H1simp (S 2 ) = 0,
H2simp (S 2 ) = Z,
simp
H>2
(S 2 ) = 0.
Exercise 8.2: Compute Hsimp for the spaces T 2 , RP 2 and K 2 , each thought of as a
-complex with two 2-simplices (and some 1- and 0-simplices).
Remark. You may like to keep in mind the following fact, even though its not
part of the logical development of this course: the map Hnsimp (X) Hn (X) an
isomorphism. So, for example, the homology of a -complex is finitely generated.
The following simple observation gives some geometric insight into singular homology.
Lemma 8.2. Let z be a singular n-cycle in X, so n z = 0. Write it as z =
PN
i=1 i i with i = 1. Then there is an -complex Z, with precisely N nsimplices (1 , P
. . . , N ) and no higher-dimensional simplices, and a map f : Z X,
such that (i)
i i represents a simplicial n-cycle for Z, and (ii) i = f i for
each i.
Proof. Since n z = 0, each face i j must cancel with another face i0 j 0 .
Thus, we can partition the set of faces of all i into pairs. We define a -complex
Z by gluing N n-simplices together along their faces, paired up in the way just
determined. This has the right properties.
More generally, if n z = y, we can build a -complex and a map from it into X
so that the summed boundary of the n-simplices in the complex maps to X as the
cycle y.
29
8.2. The Hurewicz map revisited. Last lecture, we introduced the Hurewicz
map h : 1 (X)ab H1 (X) and proved its surjectivity (assuming X path connected). We did not analyse its kernel. We now finish the job.
Theorem 8.3. When X is path-connected, the Hurewicz map h : 1 (X)ab H1 (X)
is an isomorphism.
Example 8.4.
Recall that 1 (S 1 )
= Z. Since this group is already abelian,
1
H1 (S ) = Z also.
Recall that 1 (T 2 ) = 1 (T 2 )ab
= Z2 , 1 (K 2 )ab
= Z Z/2 and 1 (RP 2 ) =
2 ab
2
2
1 (RP ) = Z/2, where T is the 2-torus, K the Klein bottle, and RP 2
the real projective plane.
Recall that the closed, oriented surface g of genus g has
1 (g )
= ha1 , . . . , ag ; b1 , . . . , bg | [a1 , b1 ] [ag , bg ]i.
Thus H1 (g )
= Z2g , generated by the classes of a1 , . . . , ag and b1 , . . . , bg .
Proof of the theorem. Take a loop ker h: say = 2 . We will show that [] is
a product of commutators in , hence lies in [, ]. We have already proved that
h is onto, so this will complete the proof. But we can build from the 2-chain a
2-dimensional -complex K, and a map f : K X, with the following properties:
if z is the sum of the 2-simplices, then 2 z = for a 1-simplex such that f = .
Moreover, the image of in K is a loop K. It suffices, then, to show that K
is in the commutator subgroup of 1 (K, b) (for the obvious basepoint b K), for
then the corresponding result will hold in X just by applying f . Thus we deduce
the theorem from the following lemma.
Lemma 8.5. Let K be a compact, connected, 2-dimensional -complex. Suppose
PN
z is the sum of the 2-simplices, and that 2 z = i=1 i for some 1-simplices i
such that i (1) = i+1 (0), i Z/N . Fix a basepoint b; take it to be a vertex lying
on K. Then 2 z represents an element of = 1 (K, b). This element lies in the
normal subgroup [, ] generated by commutators.
Proof. First observe (exercise!) that in general, if we have a free homotopy through
loops t : S 1 X, then we have two fundamental groups = 1 (X, 0 (1)) and
0 = 1 (X, 1 (1)); and [0 ] [, ] iff [1 ] [ 0 , 0 ] 0 .
We now proceed by induction on the number of 2-simplices. The lemma is
obvious when there is only one 2-simplex. When there is more than one, remove a
2-simplex adjacent to the boundary which has the basepoint as one of its vertices,
so as to create a new -complex K which again satisfies the hypotheses(!). Pick a
new basepoint b0 on K 0 which was one of the vertices of . By induction, K 0 is
a product of commutators in 1 (K 0 , b0 ), hence in 1 (K, b0 ). But K is homotopic
through loops to K 0 , so the result follows from our observation.
Remark. The lemma is connected with the geometric interpretation of the algebraic
notion of commutator length. In general, for a group , the commutator length
cl() of [, ] is the least integer g such that is the product of g commutators
in . If = 1 (X, x), then one can show that cl() is the minimal genus g of a
compact oriented surface K bounding . Here by a compact oriented surface I mean
a -complex K, equipped with a map f : K X, which satisfies the conditions of
the lemma and which is locally homeomorphic to R2 . The genus of K is half the
rank of H1simp (K).
30
TIM PERUTZ
9. Homological algebra
Having introduced singular homology, we now need an adequate algebraic language to describe it.
9.1. Exact sequences. We shall work with modules over a base ring R, which
we will assume to be commutative and unital. We write 0 for the zero-module. A
sequence of R-modules and linear maps
a
ABC
is exact if ker b = im a. A longer sequence of maps is called exact if it is exact at
each stage.
0 A B C 0,
are called short exact sequences. In such a sequence, coker a = B/ im a =
B/ ker b. But b induces an isomorphism B/ ker B im b = C. Thus a is
injective with cokernel C, while b is surjective with kernel A.
A short exact sequence is called split if it satisfies any of the following
equivalent conditions: (i) there is a homomorphism s : C B with bs =
idC ; (ii) there is a homomorphism t : B A with ta = idA ; or (iii) there is
an isomorphism f : B A C so that a(x) = f (x, 0) and b(f 1 (x, y)) = y.
If the six-term sequence
a
0ABCD0
ker b while c induces an isois exact then a induces an isomorphism A =
morphism D
= coker b.
Exact sequences are useful because if one has partial information about the groups
and maps in a sequence (in particular, the ranks of the groups) then exactness helps
fill the gaps.
Example 9.1. Suppose one has an exact sequence of Z-modules
i
0 Z A Z/2 0.
What can one say about A (and about the maps)? Choose x A with p(x) 6= 0.
Then 2x ker p = im i. There are two possibilities:
(i) 2x = i(2k) for some k. Let x0 = x i(k). Then 2x0 = 0. We can then
define a homomorphism s : Z/2 A with p s = id by sending 1 to x. Thus the
sequence splits, and so may be identified with the trivial short exact sequence
0 Z Z/2 Z Z/2 0.
(ii) 2x = i(2k + 1) for some k. Let x0 = x i(k). Then 2x0 = i(1) and
p(x0 ) = p(x) 6= 0. Given y A, either y = i(m) for some m, in which case
y = 2mx0 , or else y x0 = i(m) for some m, in which case y = (2m + 1)x0 . Thus
A = Zx0 . Moreover, x0 has infinite order (since Zx0 contains im i). So the sequence
may be identified with the sequence 0 Z Z Z/2 0, in which the map
Z Z is multiplication by 2 and Z Z/2 is the quotient map.
31
9.2. Chain complexes. A chain complex over R is a collection {Cp }pZ of Rmodules, together with linear maps dp : CpL
Cp1 , called differentials, satisfying
dp1 dp = 0. We write C for the sum
p Cp , which is a graded module, and
L
d=
dp : C C (an endomorphism map which lowers degree by 1, satisfying
d2 = 0). We define the homology
H(C ) = ker d/ im d.
L
L
Notice that ker d = p ker dp and im d =
im dp , so H(C ) = p Hp (C ), where
Hp (C ) = ker dp / im dp+1 .
Elements of Zp (C) := ker dp are called p-cycles; elements of Bp (C) := im dp+1 ,
p-boundaries. If H(C ) = 0, we say that C is acyclic.
A chain map from (C , dC ) to (D , dD ) is a linear map f : C D such
that f (Cp ) Dp and dD f = f dC . A chain map induces homomorphisms
Hp (f ) : Hp (C ) Hp (D ). A chain map which induces an isomorphism on homology is called a quasi-isomorphism.
L
9.2.1. Chain homotopies. We need a criterion for two chain maps f and g : C
D to induce the same map on homology. For this we introduce the notion of
chain homotopy. A chain homotopy from g to f is a collection of linear maps
hp : Cp Dp+1 such that
dD hp + hp1 dC = f g.
If x Zp (C) then f (x) = g(x) + dD (hp x), hence [f (x)] = [g(x)] H(D ). If, for
example, there is a chain homotopy from f : C C to the identity map idC , then
f is a quasi-isomorphism. If there is a null-homotopy, i.e., a chain homotopy from
f to the zero-map, then f induces the zero-map on homology. If both possibilities
occur then C must be acyclic, i.e., H (C) = 0.
9.2.2. Short and long exact sequences. We study the effect of passing to homology
on a short exact sequence
a
0 A B C 0
of chain complexes and chain maps.
Hp (a)
Hp (b)
Lemma 9.2.
(i) The sequence Hp (A) Hp (B) Hp (C) is exact.
(ii) Take x Ap with dA x = 0. Then [x] ker Hp (a) iff there exists y Bp+1
such that a(x) = dB y.
(iii) Take z Cp with dC z = 0. Then [z] im Hp (b) iff there exists y Bp1
such that b(y) = z and dB y = 0.
Proof. (i) Take y Bp with dB y = 0 and b(y) = dC z for some z Cp+1 . Then
z = b(y 0 ), say, and b(y dB y 0 ) = dC (z b(y 0 )) = 0, so y dB y 0 = a(x) for some
x Ap , i.e. y im a + im dB , as required.
(ii) is obvious, and (iii) almost so.
Points (ii) and (iii) can be pushed considerably further. Define the connecting
homomorphism
: Zp (C) Hp1 (A)
as follows:
(z) = [x] when there exists y Bp with b(y) = z and a(x) = dB y.
32
TIM PERUTZ
0 A B C 0
induces an exact sequence
Ha
Hb
Ha
Hb
(1)p dimk Cp .
Cp Cp1 Cp2 . . .
P
P
is an exact sequence, and p dim Cp < , then
(1)p dim Cp = 0.
Exercise 9.2: A collection C of Z-modules is called a Serre class if for every short exact
sequence 0 A B C 0 such that two out of the three Z-modules A, B,
C are in C, the third is in C also. Fix a prime p Z. Identify which of the following
properties of Z-modules M define Serre classes: (a) M is torsion; (b) M is torsion-free;
(c) M is torsion but has no p-torsion; (d) every element of M has p-power order; (e);
every element of M is divisible by p; (f) every M is finitely generated. (*) What if we
replace M by an arbitrary commutative ring R (and p by a prime of R?).
Exercise 9.3: Prove or give a counterexample: given two path connected open sets U
and V whose union is X and whose intersection U V is path connected, there exists
a short exact sequence of abelian groups
0 H1 (U V ) H1 (U ) H1 (V ) H1 (X) 0.
Exercise 9.4: If a : A B is a chain map its mapping cone, denoted cone(a) , is
the complex
cone(a)n = An1 Bn
33
with differential dcone(a) (x, y) = dA x + a(x) + dB y (check that this squares to zero).
The point of this construction is to convert questions about chain maps to questions
about chain complexes.
(i) Show that the induced map on homology, a = H(a) : H (A) H (B), is an
isomorphism iff H(cone(a) ) = 0.
(ii) Construct a short exact sequence
0 B cone(a) A1 0,
and identify the connecting map in the resulting long exact sequence.
(iii) Show that to give a map of complexes f = (h, b) : cone(a) C is to give a
chain map b : B C and a chain-homotopy h from b a to the zero map.
(iv) Show that if the map f from (iii) induces an isomorphism on homology then
there is a long exact sequence
a
Hn (A) Hn (B)
Hn (C) Hn1 (A) Hn1 (B)
Hn1 . . .
34
TIM PERUTZ
Z3 Z2 Z1 Z0 0.
So ker i = 0 when i is even and positive; and when i is odd, i+1 is onto. Thus
Hi () = 0 when i > 0. As expected, we find H0 ()
= Z.
Maps between spaces introduce homomorphisms between homology groups. Given
f : X Y , define f# : Sn (X) Sn (Y ) by
f# () = f .
It is clear that this is a chain map: n f# = f# n . Thus there is an induced map
f = Hn (f ) : Hn (X) Hn (Y ).
Notice that if g : Y Z is another map then (g f )# = g# f# , and hence
(g f ) = g f .
Remark. In categorical language, we can express this by saying that Hn defines a
functor from the category Top of topological space and continuous maps to the
category Ab of abelian groups and homomorphisms. That is, Hn associates with
each space X an abelian group Hn (X); with each map f : X Y a homomorphism
Hn (f ) : Hn (X) Hn (Y ); and the homomorphism Hn (g f ) associated with a
composite is the composite Hn (g) Hn (f ). Moreover, identity maps go to identity
maps.
Theorem 10.2. Suppose that F is a homotopy from f0 : X Y to f1 : X Y .
The homotopy then gives rise to a chain homotopy P F : S (X) S+1 (Y ) from
(f0 )# to (f1 )# , that is, a sequence of maps PnF : Sn (X) Sn+1 (Y ) such that
F
n+1 PnF + Pn1
n = (f1 )# (f0 )# .
35
To prove the theorem, we begin with low-dimensional cases, because there the
geometry is transparent.
Proof of the theorem when n is 0 or 1. First, given a 0-simplex : 0 X, note
that we have a 1-simplex P0F () := F : I = 1 Y , and 1 (P0F ()) = f1
f0 .
Next, consider some 1-simplex : 1 X. We want to examine f1 f0 .
Since 1 = I, our homotopy F defines a map on from the square 1 I to Y ,
F ( idI ) : 1 I.
But the square is a union of two 2-simplices along a common diagonal. To notate
this, let the bottom edge be 1 {0} = [v0 , v1 ], and the top edge 1 {1} = [w0 , w1 ].
Thus the square is the convex hull [v0 , v1 , w0 , w1 ] of its four vertices. It is the union
of the two triangles [v0 , v1 , w1 ] and [v0 , w0 , w1 ] along the common edge [v0 , w1 ]. Note
that by expressing these triangles as convex hulls, we implicitly identify them with
the geometric 2-simplex 2 : for the first of them, say, the point t0 v0 + t1 v1 + t2 w1
corresponds to the (t0 , t1 , t2 ) 2 .
Now define a 2-chain P1F () by applying F to each of these two simplices:
P1F () = F ( idI )|[v0 ,w0 ,w1 ] F ( idI )|[v0 ,v1 ,w1 ]
We have
2 P1F () = f1 f0 + P0F (1 ) :
the first terms come fom the top of the square, the second term from the bottom,
and the third from the two other sides. Thus, defining P1F : S1 (X) S0 (Y ) and
P2F : S2 (X) S1 (Y ) by linearly extending the definitions from simplices to singular
chains, we have that
2 P2F + P1F 1 = (f1 )# (f0 )# .
Proof of the theorem in arbitrary dimensions. We proceed in the same way. For an
n-simplex : n X, we have a map
F ( idI ) : n I Y
defined on the prism n I, and we want to express this as a sum of n+1-simplices.
To do this, we think of the prism as the convex hull [v0 , . . . , vn , w0 , . . . , wn ], where
vi is the ith vertex of {0} = , and wi the ith vertex of {1} = . Then
one can check (as Hatcher does) that
n
[
I =
[v0 , . . . , vi , wi , wi+1 , . . . , wn ],
i=0
36
TIM PERUTZ
We have
n+1 PnF () =
ji
l>k
l>k
We want the two sums here to total PnF (n ). But if j < i, then
X
PnF ( i ) =
(1)j+i1 ( idI )|[v0 ,...vj ...vi ,wi ,...,wn ] .
j
If j i then instead
PnF ( i ) =
37
0 S (A + B) S (X) Q 0
and hence a long exact sequence of homology groups
i
Hp+1 (Q ) Hp (S (A + B))
Hp (X) Hp (Q ) . . .
The theorem asserts that i is an isomorphism. From the long exact sequence, we
see that this is equivalent to the assertion that Q is acyclic, i.e., that
Hp (Q ) = 0
for all p Z.
38
TIM PERUTZ
Lemma 11.3 (Subdivision lemma). The geometric p-simplex p can be decomposed as a p-dimensional -complex in such a way that all the p-simplices 1 , . . . , N
in this decomposition have diameter < 1, and such that in the singular chain complex S (p ), one has
X
idp
i im p+1 .
i
Here we regard idp as a p-simplex in p . (In fact, one can take N = p! and the
p
.)
diameters to be p+1
The particular subdivision we have in mind here is called barycentric subdivision.
For the proof of the lemma we refer to Hatcher (it can be extracted from Steps 1 and
2 of the proof of the Excision Theorem). We will at least say what the barycentric
subdivision is. The barycenter b of a p-simplex [v0 , . . . , vp ] is the point
1
(v0 + + vp ).
b=
p+1
We now define the barycentric subdivision by induction on p. When p = 0, the
subdivision of 0 = [v0 ] has just one simplex: [v0 ] itself. When p > 0, the psimplices of the barycentric subdivision of [v0 , . . . , vp ] are of form [b, w0 , . . . , wp1 ],
where [w0 , . . . , wp1 ] is a (p 1)-simplex in the barycentric subdivision of some face
[v0 , . . . , vi , . . . , vp ] of [v0 , . . . , vp ].
The subdivision lemma is a little fiddly to prove. Since we are omitting the proof,
let us emphasize that this is an entirely combinatorial lemma concerning convex
geometry in Euclidean spaces; the target space X does not appear at all.
P
Proof of locality, granting the subdivision lemma. Write = i i Sp (p ). Now,
each i is a map p p (actually an embedding), so we can iterate the subdivisionPprocess, considering the composed maps j i : p p . Lets write
2 = i,j j i , and more generally
X
n =
in i1 .
i1 ,...,in
39
(a ,b )
a+b
0 S (A B) S (A) S (B) S (A + B) 0,
simply because S (A B) = S (A) S (B). This results in a long exact sequence
of homology groups. But Hn (S (A + B)) = Hn (X) by the locality theorem, and
hence the long exact sequence has the form claimed.
Exercise 12.1: Show that the connecting map can be understood as follows. Take
a p-cycle z Sp (X). By locality, there is a homologous p-cycle z 0 = x + y with x a
chain in A and y a chain in B. Then x = y, hence x is a cycle in A B. We
have [z] = [x].
Exercise 12.2: Show that the MayerVietoris sequence is not merely canonical, but also
natural in the following sense. Given an another excisive triad (X 0 ; A0 , B 0 ) and a map
f : X X 0 such that f (A) A0 and f (B) B 0 , the two long exact sequences and
the maps between them induced by f form a commutative diagram.
Example 12.2. As a first example of the MayerVietoris sequence, let us prove
that
H (S 1 ) = Z Z
where the first Z is in degree 0, the second Z in degree 1. We have S 1 = A B
where A = S 1 \ {(1, 0)} and B = S 1 \ {(1, 0)}. Then A B ' S 0 . Since A and B
are contractible, and A B the disjoint union of two contractible components, the
exactness of the MayerVietoris sequence
Hp (A) Hp (B) Hp (S 1 ) Hp1 (A B),
tells us that Hp (S 1 ) = 0 for all p > 1. We already know H1 (S 1 ) = Z = H0 (S 1 ) (via
1 and path-connectedness), but lets see that we can recover this by the present
method. The sequence ends with the 6-term sequence
0 H1 (S 1 ) Z2 Z2 H0 (S 1 ) 0,
where the map Z2 = H0 (A B) H0 (A) H0 (B) = Z2 is given by (m, n) 7
(m n, m n). Thus H1 (S 1 ) is isomorphic to the kernel of this map, which is
Z(1, 1), and H0 (S 1 ) to its cokernel, which is also Z.
40
TIM PERUTZ
n 0,
41
42
TIM PERUTZ
=y
=y
Hn (B)
Hn (X)
In
=y
=y
Hn1 (X)
All the vertical arrows but the middle one are isomorphisms, and hence so is
the middle one (this is the 5-lemma). Putting things together, we find that
S (A)/S (A B) S (X)/S (B) is a quasi-isomorphism, which is the result we
want.
Exercise 13.2: Re-derive H (S n )
= Z(0) Z(n) using excision.
43
Example 13.2. When working with spaces X equipped with basepoints x, its
. We put
often useful to work with reduced homology H
n (X) = Hn (X, {x}).
H
Then, by the exact sequence of the pair, the natural map
n (X)
Hn (X) H
is an isomorphism for all n > 0. The end of the sequence looks like this:
H1 (X, {x}) H0 ({x}) H0 (X) H0 (X, {x}) 0.
1 (X) = H1 (X)), so
The map H0 ({x}) H0 (X) is injective (which explains why H
44
TIM PERUTZ
Exercise 13.5: Show that the two spaces S 2 S 4 and CP 2 have isomorphic homology
groups. Likewise the two spaces S 3 S 5 and S(CP 2 ).
Remark. We will see later that the homotopy types of S 2 S 4 and CP 2 can be distinguished by their cohomology rings, but that S 3 S 5 and S(CP 2 ) have isomorphic
cohomology rings.
13.3. Summary of the properties of relative homology.
To each pair of spaces (X, A), and each integer n, it assigns an abelian
group Hn (X, A) (and we write Hn (X) for Hn (X, )).
To each map f : (X, A) (X 0 , A0 ) and each n Z it assigns a homomorphism Hn (f ) : Hn (X, A) Hn (X 0 , A0 ). One has Hn (f g) = Hn (f )Hn (g)
and Hn (id(X,A) ) = idHn (X,A) . If f0 is homotopic to f1 via a homotopy {ft }
such that ft |A = f0 |A , then Hn (f1 ) = Hn (f0 ).
To each pair of spaces (X, A), and each integer n, it assigns a homomorphism
n : Hn (X, A) Hn1 (A).
These maps are natural transformations. That is, given f : (X, A)
(X 0 , A0 ), one has
n Hn (f ) = Hn1 (f ) n
as homomorphisms Hn (X, A) Hn1 (A0 ).
Besides these basic properties, the following also hold:
DIMENSION: If denotes a 1-point space then Hn () = 0 for n 6= 0, while
H0 () = Z.
EXACTNESS: The sequence
n
Hn (A) Hn (X) Hn (X, A)
Hn1 (A) Hn1 (A) . . .
is exact, where the unlabelled maps are induced by the inclusions (A, )
(X, ) and (X, ) (X, A).
EXCISION: if (X; A, B) is an excisive triad then the map
Hn (A, A B) Hn (X, B)
induced by the inclusion (A, A B) (X, B) is an isomorphism.
ADDITIVITY: If (X , A ) is a family of pairs, then one has an isomorphism
M
a
a
Hn (X , A ) Hn ( X ,
A )
given by the sum of the maps induced by the inclusions into the disjoint
union.
(Additivity is easy to check; for finite families, it follows from excision and exactness.)
Remark. As stated, these axioms do not uniquely characterise relative homology.
However, they do characterise it if one restricts the pairs (X, A) to be CW pairs.
Alternatively, if we include one more axiom, that weak equivalences induce isomorphisms on homology, then the axioms uniquely characterise the theory for arbitrary
pairs, because every pair is weakly equivalent to a CW pair. If one omits the dimension axiom, there are many different homology theories on CW pairs, including
stable homotopy, real or complex K-theory, and oriented, unoriented or complex
bordism.
45
46
TIM PERUTZ
is injective.
The equivalence of the two assertions follows from the long exact sequences of
the pairs (Rn , U ) and (Rn , Rn \ {x}). Informally, the second assertion says that any
non-trivial (n 1)-cycle in U must wrap around some point outside U . We prove
the second assertion.
n1 (U ) maps to zero in H
n (Rn \ {x}) for all x
Proof. Suppose s H
/ U . We will
show that s = 0. The notation will be rather heavy: we fix a chain of subspaces
K V V U
47
n1 (K).
where K is compact, V open, V compact, and s is the image of some r H
n
in Hn1 (T \ Ep ).
Proof of the theorem. Any z Hn (M ) defines a continuous section sz : M HM .
If M is connected (hence path-connected), it follows from unique path-lifting that
any continuous section is determined by its value at a point. But z is represented by
a cycle that maps to a compact subset Z M . If M is connected but non-compact,
we can choosing x M \ Z. Then sz (x) = 0, and hence sz = 0.
Thus we take a class z Hn (M ) that maps to zero in some Hn (M, M \ {x}).
We must show that z = 0.
There is some compact set Z M such that z is in the image of Hn (Z). Now,
Z is contained in a finite union U1 Uq of coordinate neighbourhoods Ui ,
and it suffices to show [z] = 0 in Hn (U1 Uq ). We have already proved that
Hn (U1 ) = 0. Now take q > 1, and inductively suppose that weve shown that
Hn (U1 Uq1 ) = 0. MayerVietoris gives an exact sequence
n1 (U V ) H
n1 (U ) H
n1 (V ),
Hn (U ) Hn (V ) Hn (U1 Uq ) H
where V = U1 Uq1 and U = Uq . This reduces to
n1 (U V ) H
n1 (V ),
0 Hn (U1 Uq ) H
n1 (U V ) H
n1 (V ) induced by
and so the task is to show that the map j : H
inclusion is injective. This step is a little tricky. Take r ker j . From the long
exact sequence of the pair (V, U V ),
n1 (U V ) H
n1 (V ),
0 Hn (V, U V ) H
we find a (unique) t Hn (V, U V ) such that t = r. From the long exact sequence
of the pair (U, U V ),
n1 (U V ) 0,
0 Hn (U, U V ) H
we find a (unique) s Hn (U, U V ) with s = r. Now, s and t have images
s0 and t0 in a common group Hn (U V, U V ), and s0 t0 lies in the kernel of
n1 (V ), hence in the image of Hn (U V ): say s0 t0 comes
: Hn (U V, U V ) H
from w Hn (U V ). Since U V is non-compact, the composite Hn (U V )
Hn (M ) Hn (M, M \ {x}) maps w to zero (by the argument at the beginning of
48
TIM PERUTZ
49
Then one has a long exact sequence of the pair. Locality, MayerVietoris and
excision hold just as before, as does homotopy invariance. One has H0 (X; R) =
R0 (X) . The homology of a point, Hi (; R), vanishes for i > 0. For an n-manifold
M , one has Hi (M ; R) = 0 for all i > n and, in the non-compact case, for i = n. All
these assertions follow from routine generalisations of the proofs we have given.
Slightly trickier is the assertion that, for X path connected, one has
1 (X)ab Z R.
H1 (X; R) =
The proof we gave for the case R = Z uses the integer coefficients in a significant
way, but the general case follows from it and the universal coefficients theorem, to
be given in the next lecture. (If you have a simpler proof, tell me!)
15.2. What its good for. Homology with coefficients contains no more information than ordinary homology, but is useful for the following reasons.
When R is a field, algebraic properties of homology simplify. For instance,
over a field, passing to homology commutes with the tensor product and
dualisation operations on chain complexes.
For many spaces, the Z-homology is more complicated than the homology
over the prime fields Z/p or over Q. Moreover, considering these collectively,
one does not lose information. Good examples are RP n and various Lie
groups, notably SO(n).
When working with manifolds, in Z-homology one must distinguish the orientable and non-orientable cases, but in Z/2-homology this is unnecessary.
15.3. The local homology cover. Recall from the last lecture that the local
homology of an n-manifold (now with R-coefficients) is given by
H (M, M \ {x}; R)
= R(n) .
Collectively, these form a covering space HM,R M with fibres Hn (M, M \{x}; R).
(Here one gives Hn (M, M \ {x}; R) the discrete topology.) The sections M,R of
this covering space form an R-module. A class z Hn (M ; R) gives classes zx
Hn (M, M \ {x}; R) for all x, and this determines a homomorphism Hn (M ; R)
M,R .
50
TIM PERUTZ
51
52
TIM PERUTZ
Exercise 15.3: If M is given a smooth structure then it is orientable iff there is a smooth
atlas whose transition functions ij have positive Jacobian determinants det Dx ij .
Exercise 15.4: A manifold M is orientable if it is simply connected, or more generally,
if 1 (M ) is finite and has odd order.
Exercise 15.5: If p is odd then a Z/p-orientation determines, and is determined by, a
Z-orientation.
Exercise 15.6: Show that if a group G acts freely on M = S n , then the orientation
character 1 (M/G) = G Z/2 of the quotient manifold is given by G 3 g 7
deg(g) {1} = Z/2. For which n is RP n orientable?
Exercise 15.7: (i) Show that there is no homeomorphism Hn
= Rn , where Hn =
n
{(x1 , . . . , xn ) R : x1 0}.
(ii) Let M be a space, and N the subspace of points x M which have a neighbourhood U such that there is a homeomorphism U Hn sending x to 0. Say M is
an n-manifold with boundary if M \ N is an n-manifold. In that case write M for
the boundary N . Check that that M is an (n 1)-manifold.
Exercise 15.8: Suppose M is a compact, connected n-manifold with non-empty boundary M . It is a fact (see Hatcher) that there is a neighbourhood U of M and a
homeomorphism U M [0, 1) sending any x M to (x, 0).
(a) Show that Hp (M ) = 0 for p n and Hp (M, M ) = 0 for p > n.
(b) Show that if M \ M is orientable then Hn (M, M )
= Z.
(c) Let X be a space. Suppose h Hn1 (X) is a homology class that is representable by a manifold, in that there is a compact oriented n 1-manifold N with
fundamental class zN and a map f : N X with f zN = h. Show that h = 0 if
N = M for a compact oriented n-manifold with M = N such that f extends to
F : M X.
Exercise 15.9: Let h : S 3 S 2 denote the Hopf map, given by
h(z1 , z2 ) 7 z1 /z2 .
3
Here we think of S as the unit sphere in C2 and of S 2 as the Riemann sphere C{}.
Let be a generator for H3 (S 3 ) = Z. Then h = 0, since H3 (S 2 ) = 0. Can you find
a direct explanation for why the particular class h should be zero? In other words,
can you find a 4-chain bounding a 3-cycle representing h ?
53
Lemma 16.2. Let 0 M1 M2 M3 0 be a short exact sequence of Rmodules. Let Q be another R-module. Then the induced sequence
M1 Q M 2 Q M3 Q 0
is also exact. If the short exact sequence splits (for instance, if M3 is a free module)
then M1 Q M2 Q is injective.
Proof. Take m3 q M3 Q. Say q = p(m2 ). Then (p id)(m2 q) = m3 q.
Hence p id is onto.
Now, (p id) (i id) = (p i) id = 0, and hence there is an induced
map [p id] : (M2 Q)/ im(i id) M3 Q. Exactness of the sequence at
M2 Q is equivalent to the assertion that [p id] is injective. But im(i id) =
im i Q = ker p Q, and hence we can define s : M3 Q (M2 Q)/ im(i id)
by s(m q) = [p1 m q]. We have s [p id] = id, so [p id] is indeed injective.
Now suppose the sequence splits. Then there is a homomorphism l : M2 M2
with l i = id. Thus l id : M2 Q M1 Q satisfies (l id) (i id) = id id,
hence i id is injective.
In general, i id is not injective, but its kernel can be measured by the introduction of the torsion products. For simplicity, we assume R is Z, a field, or more
generally a principal ideal domain (PID). A PID is a commutative ring R such that
(i) xy = 0 implies x = 0 or y = 0, and (ii) every ideal is of form (x) = {ax : a R}.
The simplification in the present context because if R is a PID then every submodule
of a free R-module is free, i.e., has a basis. For the proof see, e.g., Langs Algebra,
54
TIM PERUTZ
Appendix 2 (in the general case, not restricted to finite rank free modules, this uses
Zorns lemma).
This has the following consequence: any R-module M has a two-step free resolution, i.e. an exact sequence
f1
f0
0 F1 F0 M 0
with F0 and F1 free. Just take any generating set S for M , let F0 be free on S, and
let F1 be the kernel of the natural map F0 M .
Applying Q to this sequence, we obtain a complex
f1 id
f0 id
0 F1 Q F0 Q M Q 0
which is exact at M Q and at F0 Q. We prefer to truncate this to the complex
f1 id
0 F1 Q F0 Q 0.
By the lemma, its homology at F0 Q (i.e., the cokernel of f1 id) is naturally
isomorphic to M Q. We define Tor(M, Q) to be its homology at F1 Q:
Tor(M, Q) = ker f1 id.
In brief: to define Tor, we took a free resolution of M , tensored it with Q, and
measured its failure to be exact by taking homology.
The next point is that Tor(M, Q) is independent of the choice of free resolution.
It is convenient to think of 0 F1 F0 as a chain complex F , and of f0 as an
augmentation (F M ) of the complex.
f0
g0
55
0 Hp (C ) Q Hp (C Q) Tor(Hp1 (C ), Q) 0,
where the map j is induced by the inclusion ker dp Q ker(dp idQ ). These
sequences are functorial in C . They always split, but not in a natural way.
Proof of universal coefficients. We let ip : Bp Zp be the inclusion of the pboundaries into the p-cycles. From the free resolution
ip
0 Bp Zp Hp (C ) 0
for the homology Hp (C ), we see that
coker(ip id) = Hp (C ) Q,
0 Zp Cp Bp1 0.
Since the terms are free modules, the sequence splits, and so the sequence
jp id
dp id
0 Zp Q Cp Q Bp1 Q 0
is also exact. It is, moreover, a short exact sequence of chain complexes, so it induces
a long exact sequence of homology groups. With a little thought one identifies the
connecting maps; the long exact sequence reads
ip id
Hp+1 (C Q) Bp Q Zp Q Hp (C Q) . . . ,
with ip : Bp Zp the inclusion. Its exactness tells us that there are short exact
sequences
0 coker(ip id) Hp (C Q) ker(ip1 id) 0,
i.e.,
0 Hp (C ) Q Hp (C Q) Tor(Hp1 (C ), Q) 0.
It is left as an exercise to think through why the map on the left is j. A splitting
dp
56
TIM PERUTZ
Exercise 16.2: Prove that the splittings of the universal coefficient theorem cannot
be chosen naturally in X by means of the following example. Take an embedding
i : R2 RP 2 , let D = i({z R2 : |z| < 1}), and let q be the quotient map
RP 2 RP 2 /(RP 2 \D)
= S 2 . Show that the k = Z/2 universal coefficients sequences
2
2
for RP and S cannot be split compatibly with q.
Exercise 16.3: This exercise develops an alternative approach to a special case of universal coefficients. Let C be a chain complex of free abelian groups. Show that
tensoring by the short exact sequence 0 Z Z Z/n 0 gives rise to a short
exact sequence of chain complexes 0 C C C Z/n 0. By analyzing the
resulting long exact sequence of homology groups, deduce a short exact sequence
0 Hp (C ) Z Z/n Hp (C Z/n) {x Hp1 (C ) : nx = 0} 0.
57
We say that A f en results from attaching an n-cell to A via the attaching map f .
The notation Cf refers to the general notion of the homotopy cofiber Cf =
CX f Y of a map f : X Y , where CX = X I/X {1} is the cone on X.
Example 17.2. Let c : S n1 e0 be the constant map. Then S n
= Cc = en c e0 .
Two crucial observations are:
that Cf contains A as a closed subspace; and
that the quotient Cf /A is homeomorphic to en /en , hence to S n .
`
`
More generally, given an indexing set I and a map f = iI fi : iI en A,
we can build a space
a
Cf = (A q
en )/ where fi (x) x for x en .
i
58
TIM PERUTZ
such that each X k+1 is obtained from X k by attaching a (possibly empty) collection
of n-cells. We call X k the k-skeleton of X. A cell complex is a space X together
with a sequence of closed subspaces
= X 1 X 0 X 1 X 2 . . .
such that X = k X k . A cell complex is a CW complex if X has the weak topology,
i.e. U X is open iff U X k is open in X k for each k.
S
(Im not sure how standard my usage of the term cell complex is here. However,
the term CW complex is certainly standard.)
Mostly we shall work with cell complexes with finitely many cells, which are
automatically CW complexes.
Remark. It is perhaps a pity that J. H. C. Whiteheads term CW complex, which
refers to certain technical properties, has not been replaced by something more
descriptive. W is for weak topology; C for closure finiteness, the property
that the image of each closed cell intersects the interiors of only finitely many cells
of lower dimension. In this as in other aspects of algebraic topology, there were
many possible variants of the definitionthe ones that we use lead to a streamlined
theory, and most of the others do not, but this is far from obvious.
Example 17.4. A graph is just a CW complex of dimension 1.
S
Example 17.5. The Hawaiian earring, i.e. the union n1 Cn of circles Cn in
R2 of radius n1 centered at n1 , is naturally a cell S
complex, but the topology
inherited from R2 does not make it a CW complex ( C2n \ {0} is open in the
weak topology but not the subspace topology).
Example 17.6. The 2-torus T 2 has the structure of a CW complex with one 0cell, one 1-cell (so the 1-skeleton is S 1 ) and one 2-cell. The same goes for the Klein
bottle K 2 . The real projective plane RP 2 has the structure of a CW complex with
two 0-cells, one 1-cell and one 2-cell.
Example 17.7. Complex projective space CP n = (Cn+1 \{0})/C is a cell complex
e2n e2n2 e0 . To see this, define
X 2k = {[z0 , . . . , zn ] CP n : zj = 0 for all j > k}.
Thus X 0 X 2 X 2n = CP n , and X 2k
= CP k . We will exhibit X 2k
as the 2k-skeleton of a cell decomposition. If [z] X 2k \ X 2(k1) CP n then
zk+1 = = zn = 0 but zk 6= 0, so [z] = [w1 , . . . , wk1 , 1, 0 . . . , 0] for a unique
(w0 , . . . , wk1 ) Ck . Thus X 2k \ X 2(k1)
= Ck . Thus CP n is a disjoint union of
open cells, one of each even dimension up to 2n. To see that they are attached in
the proper way, think of e2k as {w = (w0 , . . . , wk1 ) Ck : |w| 1} and define
i2k : e2k X 2k2 as follows:
ik (w) = (w0 , . . . , wk1 , (1 |w|2 )1/2 , 0, . . . , 0).
This map extends to a homeomorphism e2k fk X 2k2 X k which restricts to
the inclusion on X 2k2 , where fk : S 2k1 X 2k2 is given by fk (0 , . . . , k ) =
[0 , . . . , k , 0, . . . , 0].
Example 17.8. Let C[t] be the C-vector space of polynomials in t. Denote by CP
its projective space (C[t] \ {0})/C . Then CP is a CW complex with 2k-skeleton
(and 2k + 1-skeleton) CP k .
59
60
TIM PERUTZ
61
We wish to compute H (X) in terms of the cells and their attaching maps.
One of the convenient properties of CW complexes is that the subspace X n1
of X n has an open neighbourhood U which deformation-retracts to X n1 . Also
X n1 is closed in X n . Hence (this was an exercise using the excision theorem), one
(X n /X n1 ) when X n1 6= .
has H (X n , X n1 ) = H
Now, the idea of our calculation is to exploit the fact that X n /X n1 is a very
simple space:
_
_
Dn+1 /Dn+1 =
Sn.
X n /X n1
=
n-cells
n-cells
Thus Hn (X n , X n1 ) = Z{n-cells} .
Define n+1 as the connecting homomorphism
n+1
Hn+1 (X n+1 , X n ) Hn (X n )
from the exact sequence of the pair (X n+1 , X n ). Define qn as the natural map
qn
Hn (X n ) Hn (X n , X n1 ).
Theorem 18.1. Let Cn = Hn (X n , X n1 ) = Z{n-cells} . Define dn = qn1
n : Cn Cn1 . Then one has dn dn+1 = 0. The homology Hn (C ) of the
chain complex
dn+1
n
Cn+1 Cn
Cn1 . . .
is canonically isomorphic to the singular homology Hn (X).
This theorem is very useful as a computational tool, as we will see next time.
For now, we record two simple corollaries.
Corollary 18.2. If X is a d-dimensional CW complex then Hn (X) = 0 for all
n > d. If X is a CW complex with a finite number l of n-cells then the rank of
Hn (X) is finite and l.
Corollary 18.3. If X is a CW with only even-dimensional cells then Hn (X) =
Z{n-cells} for all n. In particular, H (CP n ) = Z(0) Z(2) Z(2n) , where the
subscripts denote degrees of the Z-summands.
Proof of the theorem. We will simplify by assuming that X is finite-dimensional,
i.e., that X = X d for some d 0. We also assume that X 6= , which forces
X 0 6= .
Step 1. We have already mentioned this step. By excision (which applies because
X n1 is a deformation retract of an open neighbourhood in X n ) one has, for n > 0,
_
_
M
k(
k (S n ).
Hk (X n , X n1 )
S n , ) = H
Sn)
H
= Hk (
=
n-cells
n-cells
n-cells
62
TIM PERUTZ
0 Hn (X n ) Hn (X n , X n1 ) Hn1 (X n1 )
(the zero on the left comes from Step 3). So, crucially, the map qn is injective. Let
dn+1 = qn n+1 . Heres the tricky step: one has
Hn (X)
= Hn (X n )/ im n+1
= im qn / im dn+1 ,
since qn carries Hn (X n ) isomorphically to im qn and im n+1 isomorphically to
im dn+1 .
Step 6. One has
im qn = ker n : Hn (X n , X n1 ) Hn1 (X n1 )
= ker dn1 : Hn (X n , X n1 ) Hn (X n1 , X n2 ) .
Thus
Hn (X)
= Hn (X n )/ im
= ker dn / im dn+1 .
Note that dn dn+1 = 0 because n qn = 0.
Remark. Let us make a note of what we have used in the proof. We needed excision
and the long exact sequence of the pair. We also needed the fact that the reduced
homology of a wedge is the direct sum of the reduced homologies. We needed
(S n ) = Z(n) . Recall that this was proved using excision (and
to know that H
homotopy invariance of singular homology) by an inductive argument beginning
k (S 0 ) = Z.
with H
We have Cn = Z{n-cells} and Cn1 = Z{(n1)-cells} , so dn is represented by an
integer-valued matrix (Dnij ):
X
Dn hii =
Dnij hji,
j
63
Proof. We can think of the closure of the cell ek as an k-simplex (since it is homeomorphic to k . Thus the ith Z-summand in Hn (X n , X n1 ) is represented by the
cell ein , considered as a singular chain (notice that it is a relative n-cycle, since its
boundary lies in X n ).
Under n , the singular chain ein is mapped to its boundary, which is the image
of the fundamental class [S n ] Hn (S n ) under the cellular attaching map fi : that
is, n ein = (fi ) [S n1 ]. Next we have to apply the quotient map qn : X n1
X n1 /X n2 , since dn (ein ) = (qn1 ) (fi ) [S n1 ] Hn (X n1 /X n2 ).
The projection Hn (X n1 /X n2 ) Z to the jth Z-summand is induced by the
collapsing map
_
j : X n1 /X n2 =
S n1 S n1
(n1)cells
n1
64
TIM PERUTZ
65
= i1
k1 (y0 : : yk1 : 0 : : 0)
= (y0 , . . . , yk2 )
k1
(check the last line for yourself!). If y int(D+
), then we have instead
i1
k1 fk (y) = (y0 , . . . , yk2 ).
We now see that the cellular differential dk is given by dk hek i hek1 i, where
is the degree of the composite map
p
ida
S k1 S k1 S k1 S k1 .
Here p is the pinching map that collapses the equator to the wedge point. The
k1
second map acts as the identity on the first wedge summand S+
and as the
k1
antipodal map a on the second wedge summand S . Now, on fundamental classes
we have p [S k1 ] = [S k1 ]+ + [S k1 ] , whilst (id a) maps [S k1 ]+ + [S k1 ] to
(1 + deg(a))[S k1 ] = (1 + (1))[S k1 ]. Thus = 1 + (1)k .
Hence the cellular complex reads
0
Z Z Z Z 0.
So we find the following: if n is even then
Z,
Z/2,
Hk (RP n ) =
0,
0,
k = 0,
k < n odd
k < n even
p n.
If n is odd then
Z,
k = 0,
Z/2,
k < n odd
n
Hk (RP ) = 0,
k < n even
Z,
k
= n,
0,
p > n.
66
TIM PERUTZ
Taking the last example further, lets now calculate H (RP n ; Z/2). We can do
this in two ways. One way is to use the cellular chain complex reduced mod 2,
which reads
0 Z/2 Z/2 Z/2 0
with a Z/2 in each degree between 0 and n. The maps are all zero. Hence
Hi (RP n ; Z/2) = Z/2 if 0 i n, and it is zero otherwise.
The other method is to use our calculation for Z-coefficients in conjunction with
universal coefficients. For n even, say, the Z/2-module Hk (RP n )Z/2 is Z/2 for
k = 0 or 0 < k < n odd, and 0 otherwise. However, Tor(Hk1 (RP n ), Z/2) is Z/2
if k 1 > 0 with k even, and is zero otherwise. Thus for any k between 0 and n,
the flanking terms in the universal coefficients exact sequence
0 Hk (RP n ; Z) Z Z/2 Hk (RP n ; Z/2) Tor(Hk1 (RP n ), Z/2) 0
always consist of Z/2 and 0 (in some order), hence Hk (RP n ; Z/2) = Z/2.
Exercise 19.1: Compute H (RP n ; Z/4). Do it in two ways: via the Z/4 cellular chain
complex; and via the homology over Z and universal coefficients.
Exercise 19.2: Use cellular homology to compute H (g ), where g is a standard
closed, orientable surface of genus g (defined, for instance, as a quotient of the 4ggon). Also, compute H (g #RP 2 ), the connected sum of g and RP 2 .
Exercise 19.3: Describe how the product X Y of finite CW complexes X and Y
inherits a structure of CW complex.
(1) Compute the Euler characteristic (X Y ) in terms of (X) and (Y ).
(2) Show that Ccell (X Y ) = Ccell (X) Ccell (Y ) as a graded abelian group.
(You are not asked to compute the differential.)
(3) Compute H (S n S n ).
(4) Let X = (S 3 )n = S 3 S 3 . Show that H3p (X) is isomorphic to the
exterior power p Zn , and that Hq (X) = 0 if q is not a multiple of 3.
67
n
En (A) En (X) En (X, A)
En1 (A) Hn1 (A) . . .
is exact, where the unlabelled maps are induced by the inclusions (A, )
(X, ) and (X, ) (X, A).
EXCISION: if (X; A, B) is an excisive triad then the map
En (A, A B) En (X, B)
induced by the inclusion (A, A B) (X, B) is an isomorphism.
ADDITIVITY: If (X , A ) is a family of pairs, then one has an isomorphism
M
a
a
En (X , A ) En ( X ,
A )
given by the sum of the maps induced by the inclusions into the disjoint
union.
Theorem 20.2 (EilenbergSteenrod). Take any homology theory E on CW pairs.
Then for any CW complex X, one has En (X, A)
= Hn (X, A). In fact, there is a
unique natural transformation E H extending a given isomorphism E ()
H (), and this natural transformation is a natural isomorphism.
Remark. The EilenbergSteenrod theorem is one of two general organising principles for (co)homology theories. The other is sheaf cohomology: the idea that one
can use different resolutions of the same sheaf to compute its cohomology. This
68
TIM PERUTZ
where C (X) is the reduced cellular chain complex C (X)/C (b). (This necessitates some slight adjustments in degree 0; otherwise the argument is identical.) To
be precise, the argument shows that we can obtain an isomorphism
n (X) Hn (CE (X))
: E
n (X) lifts to some x
n (X n ). We define (x) as the image of x
as follows: x E
E
n
n1
n (X /X
in E
).
(S n )
The dimension and suspension axioms tell us that E
= Z(n) , and hence that
n
n1
{ncells}
En (X /X
)=Z
. As before, the cellular differential may be identified
69
with a degree matrix, but at this point we hit a snag. The degree matrix is computed
rather than ordinary homology. Do these degrees agree?
using E
Let f : S n S n be a map preserving a basepoint. It has a homology degree
-degree, and we would like to prove that they are the same. This is easy
and an E
to check for S 0 , and quite easy also for S 1 , where we know that every map f is
homotopic rel basepoint to z 7 z d for some d Z. We can try to prove it in
general by induction on n, using naturality of the suspension isomorphism, but this
will only work for maps f homotopic to g for some g : S n1 S n1 . Does this
exhaust all possible maps S n S n ? Yes. This can be seen as a special case of the
Hurewicz theorem, one of the basic principles of homotopy theory. Alternatively,
it can be proved using a little differential topology (see Milnor, Topology from the
differentiable viewpoint).
and H
coincide.
The upshot is that the cellular chain complexes as defined via E
We deduce an isomorphism
(X) H
(X).
E
One now checks that this isomorphism is natural with respect to cellular maps
(those that send k-skeleta to k-skeleta) and so, by cellular approximation, with
(and
respect to arbitrary based maps. We have now succeeded in rebuilding E
hence E ) from the cellular homology of X.
. Indeed, it follows from excision
We now recover E from its reduced theory E
(X/A)
that the quotient map X X/A induces an isomorphism E (X, A)
=E
(X q ).
when A 6= ; and one has E (X) = E
One can also recover the natural transformation n , though this is trickier, and
I will omit it (see May chapters 14 and 8).
70
TIM PERUTZ
dp
C p1 C p C p+1 . . . ,
p Z,
such that dp dp1 = 0 for all p. Its just the same as a chain complex, except
that the indexing runs the other way. We use superscripts
L p for cochain complexes,
subscripts for chain complexes. We write C =
p C . The pth cohomology
module is then
H p (C ) = ker dp / im dp1 ,
L
and we put H (C ) =
H p (C ).
If (D , ) is a chain complex then one obtains a cochain complex by dualisation,
putting C p = HomR (Dp , R) and (dp f )(x) = f (p+1 x).
Remark. It is notQtrue that C = HomR (D , R), unlessLCp = 0 for |p| 0: one has
HomR (D , R) = p C p , which is usually bigger than p C p .
If X is a space, one defines the singular cochains S (X; R) by
S p (X; R) = HomZ (Sp (X), R),
with the differentials dp defined by dualising :
X
(dp c)() = c() =
(1)i c( i ).
i
p
One then puts H (X; R) = H (S (X; R)). Note that singular cochains, and hence
singular cohomology, are contravariantly functorial: a map f : X Y induces
f : H (Y ; R) H (X; R), and one has (f g) = g f .
Relative cohomology H p (X, A; R) for a subspace i : A X is defined as the
cohomology of ker(i : S (X; R) S (A; R)).
All the familiar properties of homology (long exact sequences of the pair, homotopy invariance, excision, MayerVietoris, etc.) have dual versions in cohomology,
with essentially identical proofs. Note that the connecting maps in long exact
sequences go up not down in degree.
Exercise 21.1: Formulate these properties of cohomology, and think through how you
would prove them.
Theorem 21.1 (Dual universal coefficients). If X is a space, and R any commutative unital ring, there are natural (in X), non-naturally split short exact sequences
j
71
Lemma 21.2. Let 0 M1 M2 M3 0 be a short exact sequence of Rmodules. Let P be another R-module. Then the induced sequence
p
f0
0 F1 F0 M 0.
Applying Hom(, P ) to this sequence, we obtain a complex
f
0
1
0 Hom(M, P )
Hom(F0 , P )
Hom(F1 , P ) 0
1
0 Hom(F0 , P )
Hom(F1 , P ) 0.
72
TIM PERUTZ
x
0 Hp1 (C )tors H p (C ) Hp (C ) 0,
where for any R-module M , M denotes Hom(M, R).
Proof. The proof is similar to that of the universal coefficient theorem for tensor
products. We take for our free resolution of the homology groups
ip
0 Bp Zp Hp (C ) 0
(Bp and Zp are free because Bp Zp Cp ). Thus we have
Ext(Hp1 (C ), R) = coker[ip1 : Hom(Zp1 , R) Hom(Bp1 , R)].
As in the tensor product argument, we have short exact sequences
p
0 Zp Cp Bp1 0
which split and therefore induce short exact sequences
dp
dp
i
p
73
Exercise 21.3: Work over Z and fix distinct primes p and q. Compute (i) Ext(Z, Z/pn );
(ii) Ext(Z/pn , Z); (iii) Ext(Z/pn , Z/q m ); (iv) Ext(Z/pn , Z/pm ).
Exercise 21.4: Working over Z, compute Ext(A, Q/Z) for an arbitrary finitely generated
abelian group A.
Exercise 21.5: For any simply connected space X, one has H 2 (X)
= Hom(H2 (X), Z).
Exercise 21.6: Verify a case of the theorem by computing H (RP 2 ; Z) in two ways.
Do the same for H (RP 2 ; Z/2).
Exercise 21.7: What is the relation between H (X; R) and H (X)?
74
TIM PERUTZ
(a, b) 7 a ^ b.
iI
75
(c, h) 7 c _ h.
D(c) = c _ [M ],
(a, b) 7 ha ^ b, [M ]i,
76
TIM PERUTZ
a 7 (b 7 ha ^ b, [M ]i)
is an isomorphism. If one fixes bases for the free abelian groups H k (M )/T k , it is
the assertion that the matrix of the pairing is square and its determinant is 1.
Proof of the corollary. Recall a corollary of universal coefficients (which used finite
generation): H p (M ) = Hom(Hp (M ), Z) Hp1 (M )tors . Thus H p (M ; R)/T p
=
Hom(Hp (M ), Z). The adjoint map : H p (M ) Hom(H np (M ), Z) sends a to
the map b 7 ha ^ b, [M ]i = hb ^ a, [M ]i = hb, a _ [M ]i. If b is non-zero (mod
torsion) then it evaluates non-trivially on some homology class, and by the Poincare
duality theorem, we may take that class to be of form b _ [M ]. Hence ker = T p .
And is also surjective, because given any homomorphism f H np (M ) Z, we
can represent it as evaluation on some homology class h = D(c), and then f = (c).
77
deg u = 2.
Proof. Induction on n. The space CP 0 is a point, and the result is trivially true
in this case. It is also trivially true when n = 1, and we will start the induction
there. If n > 1, observe that we have an inclusion i : CP n1 CP n , induced
by a linear inclusion Cn+1 Cn+2 . By induction, H (CP n1 )
= Z[t]/tn where
Exercise 23.1: Specify a suitable cohomology class u. Describe uk for all k. (Part of
the exercise is to work out what format the answer should sensibly take.)
Remark. A precisely similar argument, using the mod 2 version of Poincare duality,
shows that
H (RP n ; Z/2)
= (Z/2)[t]/tn+1 , deg t = 1.
The result for projective spaces shows that the cup product makes cohomology
a more powerful invariant.
Example 23.2. CP 2 is not homotopy equivalent to S 2 S 4 . Indeed, H (CP 2 )
contains a degree 2 class u such that u ^ u is non-tivial, whereas the cup-square
of a degree 2 class in H (S 2 S 4 ) is always zero.
Example 23.3. Any homeomorphism h : CP 2 CP 2 preserves orientation. For
it suffices to show that h [CP 2 ] = [CP 2 ] (where [CP 2 ] is the fundamental class),
i.e., that h has degree 1 rather than 1. But h u = u, since these are the two
generators for H 2 . So h (u ^ u) = h u ^ h u = u ^ u, hence h is the identity
on H 4 . Hence the dual map h on H4 is also the identity.
The argument also extends to CP 2n , but not to CP 2n+1 (indeed, it is false for
CP 1 ).
However, it still has limitations:
78
TIM PERUTZ
B
,
and
product
given on homogeneous monomials by
ij
j j
0
(x y) (x0 y 0 ) = (1)|y||x | (x ^ x0 ) (y ^ y 0 ).
Theorem 23.5 (K
unneth formula). Let X and Y be spaces. Define the cross
product map
H (X; k) k H (Y ; k) H (X Y ; k),
where prX and prY the two projection maps from X Y . One checks that it is a map
of graded k-algebras. When X and Y are CW complexes, and H (Y ) is a finitely
generated free k-module (e.g. when k is a field), this map is an isomorphism.
Without the assumption that H (Y ) is a free module, there is still a K
unneth
theorem; it says that the cross product map is injective, and expresses its cokernel
using Tor.
Example 23.6.
H (CP n CP m )
= Z[t, u]/(tn+1 , um+1 )
with |t| = |u| = 2, and
H (RP n RP m ; Z/2)
= (Z/2)[x, y]/(xn+1 , y m+1 )
with |x| = |y| = 1 (the signs become irrelevant when k = Z/2).
23.2. An algebraic application of cup product. A division algebra over a field
F is an F -vector space A equipped with a bilinear map m : A A A, such that
for each x A \ {0} the linear maps m(x, ) : A A and m(, x) : A A are
isomorphisms.
Theorem 23.7 (Hopf). The dimension of a finite-dimensional division algebra
over R is a power of 2.
Proof. Let A be an n-dimensional real division algebra, with multiplication
m : A A A.
Bilinearity of m, and the fact that m(x, y) = 0 implies x = 0 or y = 0, tell us that
there is an induced map
M = Pm : PA PA PA.
79
Here H (g) is the induced map on H (X; Q). What restrictions does this entail for
groups acting freely on CP n ?
p
80
TIM PERUTZ
a ^ b = D# (a b).
(p + q)-cells e
p eq . The cellular chain complex of the product is given by
Ccell (X Y ) = Ccell (X) Ccell (Y ),
with the cellular boundary dXY (e f ) = dX e f + (1)deg e e dY f .
I wont give the proof (see Hatcher p. 268 or May p.99), but the basic point is
that the product of a p-cell and a q-cell is a p + q-cell.
: Ccell
(X) Ccell
(X) Ccell
(X X).
Now, D is not a cellular map, but we have the cellular approximation theorem:
Theorem 24.2. Every map between CW complexes is homotopic to a cellular map,
i.e., one which maps the k-skeleton to the k-skeleton for all k.
So D is homotopic to a cellular map D0 : X X X.
Example 24.3. Make S 1 a CW complex with exactly two cells. Then one can see
from a picture how the diagonal S 1 S 1 S 1 is homotopic to a cellular map.
Exercise 24.1: Find a cellular approximation to the diagonal S n S n S n .
81
82
TIM PERUTZ
Thus S (X) is a differential graded algebra (DGA). For a map f , one has f (a ^
b) = f a ^ f b, so this DGA is natural in X.
Thus ^ descends to give a unital ring structure on cohomology H (X) bilinear
product on cohomology, natural in X.
Exercise 24.3: Use singular cohomology to compute H (S 1 S 1 ) as a ring. [You may
find it helpful to think of S 1 S 1 as a -complex.]
The cap product is defined by a similar procedure:
1D#
83
25. Non-commutativity
We sketch a definition of the simplest interesting Steenrod operation,
Sq n1 : H n (X; Z/2) H 2n1 (X; Z/2)
and use it to distinguish two homotopy types.
In the de Rham cohomology theory of smooth manifolds M , which is naturally
isomorphic to M 7 H (M ; R), the wedge product of differential forms is commutative (in the graded sense). However, the cup product of cochains is not commuative.
Lemma 25.1. The cochain-level cup product of singular cochains is not commutative. However, there exist natural maps : S p (X) S q (X) S p+q1 (X) satisfying
the chain homotopy identity
a ^ b (1)|a||b| b ^ a = d(a b) + (da b + (1)|a| a db).
Hence the cohomology cup product is commutative (in the graded sense).
I will not prove this lemma.
Suppose we work with cochains over Z/2. Construct the natural map as in
the lemma. For [c] H n (X; Z/2), define Sq n ([c]) = [c ^ c] H 2n (X; Z/2), and
Sq n1 ([c]) = [(c c)] H 2n1 (X; Z/2).
This makes sense because the identity satisfied by implies that, working mod 2,
we have that d(cc) = 0 when dc = 0, and also that ((c+db)(c+db))(cc)
is exact.
It is eminently plausibleand moreover truethat if (X; A, B) is an excisive
triad then the MayerVietoris connecting maps p : H p (A B) H p+1 (X) satisfy
Sq n1 (n c) = 2n1 (Sq n1 c),
c H n1 (A B).
(Here the Sq n1 on the right is the cup-square, that on the left the one defined
using .) It follows that one has (for n > 0)
Sq n1 (c) = (Sq n1 c) = (c ^ c),
c H n1 (X),
84
TIM PERUTZ
duality
26. Poincare
We provide most but not all of the details of the proof of Poincare duality.
Poincare duality is primarily a statement about compact manifolds. Let us recall
the statement in that case.
Theorem 26.1 (Poincare duality: compact case). Suppose M is a compact, connected, R-oriented n-manifold with fundamental class [M ] Hn (M ; R). Then the
duality map
D : H p (M ; R) Hnp (M ; R),
D(c) = c _ [M ],
where the direct limit is over compact subsets K X. To explain this, note that
the compact subsets of X form a direct system under inclusion. This means that
one has inclusion maps i : K1 K2 , and the composite of inclusion maps is again
an inclusion map. The identity map on X is then a map of pairs
(X, X \ K2 ) (X, X \ K1 ),
and hence induces a homomorphism
H p (X, X \ K2 ; R) H p (X, X \ K1 ; R).
In this way, the modules H p (X, X \ K) become a direct system as K ranges over
compact subsets. We defined Hcp (M ; R) as the direct limit of this system. Thus
one has a canonical map
H p (X, X \ K; R) Hcp (X; R)
for each compact K, and these commute with the maps in the inverse system.
Indeed, Hcp (X; R) is universal with respect to this property.
In practice, one computes Hcp (X) as the direct limit of H p (X, X \ K) as K
ranges over some compact exhaustion, i.e., a family of compact subspaces Ki such
that every compact subspace is contained in some Ki . This is valid by abstract
nonsense about cofinal families.
Example 26.2. Let us compute Hcp (Rn ). A compact exhaustion is given by the
discs Dn (m) centred at 0 and of radius m = 1, 2, . . . . One has H (Rn , Rn \
Dn (m))
= Z(n) . The inclusion of D(m) in D(m + 1) obviously induces an isomorphism on the relative cohomology groups. Hence Hc (Rn )
= Z(n) .
1If you know about de Rham theory, you should think of differential forms with compact
support.
85
p
The cap product _ : H (X) Hq (X) Hqp (X) generalizes to a relative cap
product
H p (X, A) Hq (X, A) Hqp (X),
defined at (co)chain level by the same formula as the original cap product. Indeed,
one has ( _ a) = d _ a _ c. If represents a cocycle rel A, and a a
cycle rel A, then both terms vanish. When p = q, the cap product is essentially
just the evaluation pairing. More precisely,
h1, _ ai = h, ai.
Theorem 26.3 (Poincare duality: general case). If M is an R-oriented n-manifold
then the duality map
D : Hcp (M ; R) Hnp (M ; R),
is an isomorphism. Here D is defined as the direct limit of the maps
DK : H p (M, M \ K; R) Hnp (M ; R),
c 7 c _ [MK ],
KU
KU
Proof of the theorem. We drop the coefficients R from the notation. We shall prove
that D : Hcp (U ) Hnp (U ) is an isomorphism for each open subset D M
(including, eventually, M itself).
Step 1. The result holds when U = Rn .
Indeed, we saw above that Hc (Rn )
= R(n)
= Hn (Rn ). For any compact subset
K, the cap product of a generator for H n (Rn , Rn \ K) by the fundamental class
[RnK ] Hn (Rn , Rn \ K) is (up to a sign) the point class in H0 (Rn ), because of the
relation between relative cap product and the evaluation pairing. Passing to the
direct limit, we find that D is an isomorphism in this case.
Step 2: If the result holds for U , V and U V then it holds for U V .
I claim that there is a covariant MayerVietoris sequence in compactly supported
cohomology, and that one has a commutative diagram with exact rows of which
one portion reads
Hcp (U ) Hcp (V )
Dy
Hcp (U V )
Dy
Hcp+1 (U V )
Dy
86
TIM PERUTZ
KUi
But
[
Hnp (V ) = Hnp ( Ui ) = lim Hnp (Ui ),
because the singular chain complex of an expanding union is the direct limit of
the singular chain complexes, and taking homology commutes with direct limits.
Now D : Hcp (Ui ) Hnp (Ui ) is an isomorphism, and it commutes with the maps
induced by inclusion of Ui in Ui+1 ; hence D = limi (D : Hcp (V ) Hnp (M )) is an
isomorphism.
Step 4: The result holds for open subsets of Rn .
Every open set U Rn is the union of a countable set of open balls. Hence U
is the union of a nested family U1 U2 U3 . . . , where U1 is an open ball, and
Ui+1 is the union of Ui and a convex open set C homeomorphic to Rn . Note that
C Ui is convex, open, and has compact closure, and hence is homeomorphic to a
ball. Thus the result holds for Ui+1 by induction and steps 1 and 2.
Step 5: The result holds for M .
By Steps 1 and 4, and Zorns lemma, there is a non-empty, maximal open set O
for which the result holds. If this werent all of M , we could take the union of O
and a coordinate neighbourhood U
= Rn , disjoint from O; the result would then
hold for O U by steps 1 and 2. This contradicts maximality of O.
Remark. There is an almost trivial proof of duality for compact smooth manifolds
M in the context of Morse theory. The cellular chain complex can be understood
in terms of critical points and gradient flows for a Morse function f . Replacing f
by f does not change the homology (which is just H (M )) but it has the effect of
dualizing the chain complex and changing degree to degree n . Thus H (M )
is isomorphic to the cellular cohomology of M in degree n .
Exercise 26.1: Let M be a compact, connected, oriented 3-manifold. Determine the
graded ring H (M ) when (i) 1 (M ) is finite; (ii) H1 (M )
= Z; (iii) H1 (M )
= Z2 .
Exhibit two such 3-manifolds with non-isomorphic cohomology rings, both of which
have H1
= Z3 .
Exercise 26.2: Show that the Euler characteristic of a compact, orientable, odd-dimensional
manifold is zero.
Exercise 26.3: For which even dimensions 2n is it true that the Euler characteristic of
a compact, connected, orientable 2n-manifold is necessarily even?
Exercise 26.4: Show that for every map f : S 2n CP n , the induced map f on
reduced homology is zero.