Top 2
Top 2
A.B.Sossinsky
TOPOLOGY-2
TABLE OF CONTENTS
Foreword
Lecture 1. Homology functors
Lecture 2. CW-complexes
Lecture 3. Homotopy groups
Lecture 4. Cellular homology
Lecture 5. Simplicial homology
Lecture 6. Properties of simplicial homology
Lecture 7. Singular homology
Lecture 8. Applications of homology
Lecture 9. Cohomology
Lecture 10. Poincaré duality
Lecture 11. Obstruction theory
Lecture 12. Vector bundles and G-bundles
Lecture 13. Miscellany
3
FOREWORD
The present booklet is based on lectures given in the fall semester of 2009
to second year IUM students in Russian and to Math in Moscow students
in English. I prepared and distributed handouts, typeset in LaTeX, at the
end of each lecture (the handouts, written in English, were given to the
American students and to the Russian ones as well). These handouts, with
slight revisions, are gathered in this small book.
The Topology-2 course at the IUM (and in the framework of the Math in
Moscow program) is traditionally an introductory course in algebraic topol-
ogy, mainly about homology theory. The students taking it have already had
Topology-1, which at the IUM is an elementary introduction to topology with
emphasis on its geometric and algebraic aspects. Topology-1 includes a min-
imal amount of general topology (topological spaces and continuous maps,
topological equivalence, compactness, connectedness, separability) and such
geometric topics as plane curves, surfaces, vector fields, covering spaces, ex-
amples of 3-manifolds, homotopy, treated by means of the corresponding
basic algebraic invariants (degree of circle maps, Whitney winding number,
Euler characteristic, Poincaré index, fundamental group, Morse index). For
more details about that course, the reader is referred to the book Topology-1
in the same series written jointly by V.V.Prasolov and myself.
As to the Topology-2 course, the titles of the 13 lectures will quickly give
an idea of its contents. As usual, I regard the ability of solving the problems
(appearing at the end of each lecture and used in the exercise class as well as
for homework assignments) as at least as important as mastering the theory.
In preparing the lectures, my main sources of information (besides my own
memories of the subject) were the wonderful books Course in Homotopic
Topology by A.T.Fomenko and D.B.Fuchs, V.V.Prasolov’s Homology Theory,
and S.V.Matveev’s Lectures on Algebraic Topology.
Teaching the Topology-2 course in the fall and winter of 2009 was a very
satisfying experience: the IUM class of some twenty students was very re-
ceptive and attentive, they corrected errors in the lectures and the handouts,
asked very pointed questions, and politely complained whenever the proofs
were only sketched. I am especially grateful to Vladimir Eisenshtadt and
Daniel Le for numerous useful remarks and to Alexey Deinega (who helped
prepare the first version of the illustrations). I would also like to thank Victor
Shuvalov for the final version of the illustrations and ??? , for rapidly and
efficiently editing this booklet.
4
Lecture 1
HOMOLOGY FUNCTORS
h(gf ) = (hg)f : X → W.
The main protagonists of this course are homology theories, which are
covariant functors from certain topological categories to certain algebraic
ones, and we discuss one of them in a section of its own.
Hn (Sn ) ∼
= Z, Hn (Dn ) ∼
= 0 for any n ≥ 1,
C −−−→ D
g
XOO @
@F
@
i
@
A // B
f
where i denotes the inclusion i(x) = x for all x ∈ A, while the dashed
diagonal arrow is the map that we are searching for. The retraction problem
is the particular case of the extension problem in which A = B and f =id.
The lifting problem is as follows: given maps f : A → B and p : X → B,
to find the lift of f , i.e., a map F : A → X such that p ◦ F = f . The
corresponding triangular diagram has the form
X
~>>
F ~ p
~
~
A // B
f
where again the dashed diagonal arrow indicates the desired map. The
section problem is the particular case of the lifting problem in which we have
f = idB : A = B → B and p : X → B is a fiber bundle (a notion that will
be discussed in the next lecture).
Homology theory often provides easy negative solutions to the basic prob-
lems by reducing them to very simple algebraic questions, in particular, for
cases in which the direct topological (geometric) solutions are hopelessly dif-
ficult. We conclude this lecture by such an example. Many more will be
discussed in the exercise class.
8
f (x)
f (y)
y = r(Y )
1.5. Problems
In order to solve the subsequent problems, you can assume known the
homology groups of Dn , Sn , RP n , CP n , T2 , Mg2 .
1.3. Can the torus be retracted on its meridional circle?
1.4. Is there a retraction of S5 on S4 , where S4 is the “equator” of S5 , i.e.
S5 = ΣS4 (here Σ denotes suspension) ?
1.5. Is there a retraction of RP 5 on RP 4 , where RP 4 ,→ RP 5 is the natural
embedding given by (x1 ; . . . ; x5 ) 7→ (x1 ; . . . ; x5 ; 0) ?
1.6. Is there a retraction of the solid torus S = S1 × D2 on its boundary
dS = T2 = S1 × S1 ?
1.7. Given a continuous map p : RP 2 → S1 × D17 and a homeomorphism
h : S1 × D17 → S1 × D17 , can h be lifted to a map H : S1 × D17 → RP 2 ?
1.8. Prove that Euclidean spaces of different dimensions are not homeo-
morhic.
10
Lecture 2
CW-COMPLEXES
This will complete the proof, because g1 now satisfies condition (iii) (as well
as the conditions (i) and (ii)).
y
−1
χ−1 (y)
(y)
−1(y)
χ
χ
χ
χ
χ −1
−1
−1
(y)
(y)
(y)
D n
g χ
Dm m
ε := {x ∈ R : kxk ≤ ε},
so that Dm = Dm m
1 . For any 0 < ε < 1, the disk Dε is homeomorphic to
m m m
χ(Dε ) ⊂ Y , and so we identify Dε and χ(Dε ) ⊂ Y . The map f is uniformly
continuous on the compact set f −1 (Dm
3/4 ), so we can choose a δ > 0 such that
x, y ∈ f −1 (Dm
3/4 ) ⊂ D
n
& kx − yk < δ =⇒ kf (x) − f (y)k < 1/4.
We shall construct the map g and the homotopy separately for each sim-
plex. In case (a), we set g(v) = f (v) for all vertices of the simplex and extend
the map by linearity. In case (b), we don’t change anything.
For a simplex whose image intersects Sm−11/2 (case (c)), the situation is more
complicated, because g is already defined on some of the faces (namely, on
those satisfying (a) or (b)), and we must extend g to the entire simplex. For
the vertices, we set g(v) = f (v). On a 1-face, if the map is not yet defined,
we extend it from the endpoints by linearity. On a 2-face ∆2 , if the map
is not yet defined, we define it as follows: we cover ∆2 by segments [m, x]
joining its baricenter m to points x on its boundary (at which g is already
defined), then set g(m) = f (m) and extend g to [m, x] by linearity. Further,
we perform the same construction for 3-faces, and so on, thus defining g on
the entire disk Dn .
A look at Fig.2, which shows only a “sector” of the disk Dm , should help
visualize what is going on.
1/4
2/4
3/4
Figure 3. The image of ∆k in Dm
U × FJ / p−1 (U )
JJ
JJ
JJ p
pr1 JJ
JJ
%
U
B −−−→ B 0
f
and p1 (d, e) = p(e). By the Feldbau Theorem, the bundle H ∗ (p) is trivial.
Now consider the identity Dn+1 → Dn+1 as a homotopy Jt : Dn → Dn+1 ,
define
Je0 : Dn → E1
by setting
Je0 (d) = ((d, 0), H0 (d)),
denote Jt0 = Jt |Sn−1 , and define Jet0 ; Sn−1 → E1 by setting
Jet
-E /9 E
qq q8 @ 1 ϕ
sssB
ss
qqq
Je0 qqq He 0 sss
q ′ H ∗ (p) ss p
qqqqq Jet sss
s
Het ′
q J ′ ss
n
qqq n−1 t
ss
D ⊃ S / Dn+1 ⊃ D n
⊃ S n−1
6; B
4
Jt Ht
Case 3. We now suppose that the bundle p : E → B and the pair (X, X 0 )
are arbitrary. The proof readily proceeds by induction on the dimension n
of the cells constituting X n ⊂ X, using the results of Case 2 to define the
required map for each n-cell.
19
2.6. Problems
2.1. Let C be the union of all circles of center (1/n, 0) in the xy-plane and
radius 1/n. Prove that C is not homeomorphic to a CW-space.
2.2. a) Find a space satisfying (W) but not (C).
b) Find a space satisfying (C) but not (W).
c) Is the closure of cell necessarily a subspace ?
2.3. Find mimimal CW-complex structure on CP n , RP n , #qi=1 T2i , #qi=1 RPi2 .
2.4. Define S∞ , RP ∞ , CP ∞ and supply them with CW-structure.
2.5. Prove that S∞ is contractible.
2.6. Give an example of a nontriangulable two-dimensional CW-complex.
2.7. Prove that π1 X lives in X (2) (the 2-skeleton of X), i.e., show that
π1 (X) ≡ π1 (X (2) ).
2.8. Prove that a CW-complex is connected iff its 1-skeleton X (1) is con-
nected.
2.9. Prove that a CW-complex is connected iff it is path-connected.
2.10. Prove that πk Sn = 0 for all k < n.
2.11. Show that the Cartesian product, the cone, the suspension, and the
join of CW-complexes have natural CW-complexes structures.
2.12. Prove that any finite CW-complex X n can be embedded in RN , where
N = (n + 1)(n + 2)/2.
2.13. Show that πn (S1 ) are trivial for all n > 1. (Hint : Consider the bundle
p : R1 → S1 whose projection is given by formula p(x) = exp(xi) )
2.14. Prove that a (locally trivial) fiber bundle p : Sn → B whose base B
consists of more than one point is not homotopic to a constant map.
20
Lecture 3
HOMOTOPY GROUPS
Like homology theory, homotopy group theory is a functor from the cat-
egory of topological spaces and their (continuous) maps to the category of
graded groups and their homomorphisms. The construction of homotopy
groups is much simpler than that of homology groups and they have many of
the properties of homology groups, so that they could be used to solve topo-
logical problems in the spirit of Lecture 1. Unfortunately, they are much
more difficult to compute than the homology groups, and this restricts their
applications. Thus the computation of the homotopy groups of spheres is
still an open problem (despite over 50 years of efforts by the world’s best
topologists), but we shall compute one of them π3 (S2 ), which will give us the
occasion to discover (rediscover?) the beautiful Hopf bundle.
In
ϕ
x0
X
Sn
Figure 3.1. Spheroids
ϕ1
I1n
x0
X
I2n
ϕ2
f g
f g
g f
g f
f∗ : πn (X, x0 ) → πn (Y, y0 ).
This is done in the natural way, i.e., by setting f∗ := {f ◦ α} for any spheroid
{α} ∈ πn (X, x0 ), where the curly brackets denote base point preserving ho-
motopy classes. The fact that f∗ is well defined (i.e., does not depend on the
choice α ∈ {α}) is a straightforward verification.
22
X ' Y =⇒ πn (X, x0 ) ∼
= πn (Y, y0 ) for all n > 0.
As pointed out above, although the homotopy groups are easy to define,
they are hard to compute. Here are some examples and properties:
• πn (point) = 0 for all n ≥ 0;
• πk (Sn ) = 0 for all k < n (this easily follows from the Cellular Approxi-
mation Theorem);
• πn (Sn ) = Z for all n ≥ 1 (this is a consequence of the Hurewicz Theorem,
which will be proved in subsequent lectures, and the fact that Hn (Sn ) = Z);
• πn (X × Y ) = πn (X) ⊕ πn (Y ) (obvious);
An important property of the homotopy groups is the fact that there
exists a natural action of the fundamental group π1 (X, x0 ) on πn (X, x0 ),
n ≥ 2. Because of this action, π2 (S1 ∧ S2 ) is not isomorphic to Z, as one
might naively suppose. Indeed, we can modify the spheroid id :S2 → S2
by first mapping S2 to the wedge S2 ∧ I by identifying the parallels within
the Arctic Circle to points and then wrapping the “tail” I around S1 several
times. The element of π2 (S1 ∧ S2 ) thus obtained differs from the spheroid id.
In fact π2 (S1 ∧ S2 ) ∼
= Z ⊕ ··· ⊕ Z ⊕ ....
X
x0
is called exact at the term Gi−1 if Im(ϕi ) = Ker(ϕi−1 ), and simply exact if it
is exact at all its terms. It is easy to prove (see the exercise class) that in
any exact sequence of the form
ϕ
0 −→ G2 −→ G1 −→ 0
α
s0
b0
ϕt
Stn−1
n
S
B
Figure 3.5. Definition of ∂∗
et : Sn−1 → B.
By the covering homotopy theorem, there exists a covering ϕ
We define ∂∗ by setting ∂∗ (α) := ϕ
f1 .
Theorem. The sequence of homomorphisms
i
∗ p∗ ∗ ∂
· · · → πn (F, e0 ) −→ πn (E, e0 ) −→ πn (B, b0 ) −→ πn−1 (F, e0 ) → . . .
inclusion (namely Ker(∂∗ ) ⊂ Im(p∗ )), leaving the other ones for the exercise
class.
We can represent any spheroid α0 : (Sn , s0 ) → (B, b0 ) as a homotopy
αt : Sn−1 → B and (by the Covering Homotopy Theorem) and consider
the covering homotopy α et : Sn−1 → E. If α0 was chosen in Ker(∂∗ ), then
n−1
α
e1 : S → F is homotopic to the constant map. Let βs be the homotopy
in F joining α
e1 to the constant map. Consider the homotopy
(
αe2t if t ∈ [0, 1/2],
ψet =
βe2t−1 if t ∈ [1/2, 1].
i.e., continuous maps such that α(s0 ) = x0 and homotopies Ht such that
Ht (∂Dn ) ⊂ A and Ht (s0 ) = x0 for all t ∈ I. It is often more convenient
to interpret spheroids as (classes of) maps (In , ∂In , s0 ) → (X, A, a0 ) (this
interpretation gives the same result, because the pair In /∂I n is homeomorphic
to (Dn , ∂Dn ) The main definition, that of the product of two relative spheroids
is relegated to the exercise class, together with the proof of the fact that
πn (X, A, a0 ) is a group for any n ≥ 2 and is Abelian for n ≥ 3.
Of course the homotopy group πn (·, ·) of pairs can be regarded, for n ≥ 3,
as a functor from the category of pairs of topological spaces to the category
of Abelian groups; the main definition, that of the induced homomorphism
f∗ : πn (X, A, a0 ) → πn (Y, B, b0 )
corresponding to the map of pairs f : (X, A) → (Y, B), is left as an exercise.
Given a pair of spaces with base point, (X, A, a0 ), we have the inclusion
map i : (A, a0 ) ,→ (X, a0 ), which determines the induced homomorphism i∗ .
Since any absolute spheroid (Dn , s0 ) → (X, a0 ) can be regarded as a relative
25
π3 (S2 ) = Z.
26
To prove this, let us write out part of the homotopy sequence for the Hopf
bundle:
p∗ ∂ i
· · · → π2 (S3 ) −→ π2 (S2 ) −→
∗
π1 (S1 ) −→
∗
π1 (S3 ) → . . . .
Since π2 (S3 ) = π1 (S3 ) = 0 (see the previous section), we obtain the isomor-
phism π2 (S2 ) ∼= π1 (S1 ), and since the fundamental group of the circle is Z,
2 ∼
we have π2 (S ) = Z. (This is particular case of a fact mentioned but not
proved in the previous section.)
Now let us write out another part of the same homotopy sequence for the
Hopf bundle:
i p∗ ∂
· · · → π3 (S1 ) −→
∗
π3 (S3 ) −→ π3 (S2 ) −→
∗
π2 (S1 ) → . . .
Since the two extreme terms are zero, we obtain π3 (S3 ) ∼ = π3 (S2 ), and our
claim follows from the isomorphism (mentioned in the previous section for
any n, not only 3) π3 (S3 ) ∼= Z.
3.5. Homotopy groups of spheres: some information
The computation of πk (Sn ) for k > n, arguably a useless task, was one
of the central topics of mathematics in the 1950ies and 1960ies, mostly out
of sheer curiosity. The remarkable work of L.S.Pontryagin, J.-P.Serre, and
J.F.Adams eventually turned out to have other, much more useful applica-
tions. Although open questions still remain, the computation of the homo-
topy groups of spheres is no longer very fashionable today. Here we list only
a few results, chosen more or less at random, that may seem very strange at
first glance.
• πn (Sn ) = Z, n ≥ 1;
• πn+1 (Sn ) = Z2 , n ≥ 3;
• πn+2 (Sn ) = Z2 , n ≥ 4;
• πn+3 (Sn ) = Z24 , n ≥ 5;
• πn+4 (Sn ) = 0, n ≥ 6;
• πn+7 (Sn ) = Z240 , n ≥ 9;
• πn+9 (Sn ) = Z2 ⊕ Z2 ⊕ Z2 , n ≥ 11;
• πn+11 (Sn ) = Z504 , n ≥ 13.
The main tool for proving these results are the so-called spectral se-
quences, in particular those due to J.-P.Serre and J.F.Adams. Spectral se-
quences are a very sophisticated and effective means for computing homology
groups, but are beyond the framework of this course.
27
3.6. Problems
Lecture 4
CELLULAR HOMOLOGY
point the (n−2)-skeleton X (n−2) as well as all the (n−1) cells of X (n−1) except
α. Define the map χ b : ∂Dn → Sn−1 by setting χ b := p ◦ χ. By definition, the
incidence coefficient of the cell β to the cell α is the integer
χ) = deg(p ◦ χ) .
[β : α] := deg(b
(Here the orientations of the spheres ∂Dn and Sn−1 are those induced by the
corresponding cells.) The incidence coefficient is correctly defined if all the
maps of the form χ b are neat for all pairs of cells α and β. In that case, we say
that the CW-space X is neat. We will assume that all CW-spaces considered
in the rest of this lecture are neat.
Example. Let X be the CW-space structure on the projective plane
RP 2 consisting of three cells of dimensions 0,1,2. The 2-cell β : D2 → X is
attached to X (1) = S1 by the map χ : ∂D2 = S1 → S1 , eiϕ 7→ e2iϕ . Then the
incidence coefficient [β : α], which expresses the number of times that the
boundary of β wraps around α, is equal to 2.
Remark 1. The key property of neat CW-spaces is that for them the
incidence coefficient of two cells is correctly defined via the definition of degree
of sphere maps. There are many different ways of ensuring this property, and
other authors use, for this purpose, notions other than what I call “neatness”.
Remark 2. The general construction described above becomes unrecog-
nizable in dimension k = 1, so we will treat this case separately. The 1-disk
D1 is the segment [0, 1], which we always assume oriented from 0 to 1, and
any 1-cell σ is determined by a map of its boundary ∂D1 = {0} ∪ {1} to the
0-skeleton (the vertices) of our CW-space. Now if the point {0} is mapped to
some point p of the 0-skeleton, then the corresponding incidence coefficient is
set to be equal to −1, while if {1} is mapped to some point q, the incidence
coefficient taken to be +1; in the case p = q, the incidence coefficient will be
set equal to zero.
and then extending to the entire group Cn (X) by linearity. This definition
makes sense provided we know that the incidence coefficients are well defined;
this will be the case if the CW-space X is neat, but we have already assumed
this.
Chains c ∈ Ck (X) such that ∂k (c) = 0 are called cycles and those for
which there exists a chain c0 ∈ Ck+1 such that ∂k+1 (c) = c0 are called bound-
aries (or are said to be homologous to zero). Two cycles whose difference
is homologous to zero are said to be homologous. Being homologous is, of
course, an equivalence relation.
We claim that the boundary operator satisfies the Poincaré Lemma, i.e.,
we have
∂k+1 ◦ ∂k = 0 for all k ≥ 0.
We omit the proof of this statement.
The Poincaré lemma implies that Im∂k+1 ⊂ Ker ∂k ⊂ Ck (X), and this
allows us to define the nth (cellular) homology group of a (neat) finite CW-
space X by taking the quotient group
Ker ∂k+1
Hk (X) := for all k ≥ 0.
Im ∂k
f∗n : Cn (X) → Cn (Y );
f∗ : Hn (X) → Hn (Y )
4.5. Problems
4.6. Let p and q be two relatively prime positive integers. Consider the
3 2
action of the group Zp with the generator σ on the unit sphere S ⊂3 C
defined by σ(z, w) = exp 2πi/p z, exp 2πiq/p w . The quotient of S by
this action is a 3-manifold. It is called a lens space and is denoted by L(p, q).
Compute H∗ (L(p, q)
4.7. Compute H∗ (D), where D is the dunce hat, i.e., the triangle with the
identifications shown by arrows. Can it be retracted to its circle N M ?
4.8. Can the dunce hat be retracted to its circle N M ?
M
35
S
36
Lecture 5
SIMPLICIAL HOMOLOGY
∂n+1 n ∂ ∂n−1 1 ∂
. . . −−−→ Cn −−−→ Cn−1 −−−→ . . . −−−→ C0
y y y
0 0
∂n+1 ∂0 ∂n−1 ∂0
. . . −−−→ Cn0 −−−
n 0
→ Cn−1 1
−−−→ . . . −−−→ C00
The fact that this is indeed a category (i.e., the two functoriality axioms
in the definition of categories, see Lect.1, hold) follows immediately from
definitions.
take some element c ∈ Ker(∂n ), and let h be its homology class. Define
c0 := fn (c), and denote by h0 the homology class of c0 . Now we can define
the induced homomorphism f∗ : Hn (C) → Hn (C 0 ) by assigning h 7→ h0 . The
fact that f∗ (here and elsewhere we omit the subscipt n) is a well-defined
homomorphism follows by an fairly easy chase in the last diagram.
It follows directly from the definitions that the assignment described
above of homology groups H∗ (C) and induced homomorphisms f∗ to chain
complexes C and their morphisms f is a functor, i.e.,
f ◦ g ∗ = f∗ ◦ g∗ and (idC )∗ = idH∗ (C) .
38
The set Cn (X) of all n-chains, called the nth chain group has an obvious
structure of an Abelian group generated by allLthe (ordered) n-simplices
(actually it is a free Z-module); the direct sum n≥0 Cn (X) of the n-chain
groups is also an Abelian group denoted C∗ (X).
To obtain a chain complex from the groups Cn (X), we must define the
boundary operators or differentials
for its (n − 1)-face obtained by throwing out its jth vertex. We now define
the boundary operator by setting
X X X n
X
∂n c = ∂n zi σin := zi ∂n (σin ) := zi (−1)j [v0 , . . . , vj∨ , . . . , vn ].
i≥1 i≥1 i≥1 j=0
to the image chain); we do this for all simplices appearing in c with nonzero
coefficients and then combine (add together) the coefficients at like simplices
τjk ⊂ Y , obtaining the image chain, denoted by f∗k (c) ∈ Ck (Y ).
It is easy to verify that
is immediate: one simply takes (see Sec.5.2) the homology groups and the
induced homomorphisms of the chain complex constructed above.
We omit the details at this stage and note without proof that we have defined
a functor from the category of pairs of simplicial spaces to the category of
graded groups.
43
5.8. Problems
5.4. Compute the homology of the boundary of the triangle and the bound-
ary of the square (by using the definition of simplicial homology groups).
Compare.
5.5. Compute the homology of the boundary of the tetrahedron and the
boundary of the cube (its faces are triangulated by means of their diagonals).
Compare.
5.6. Prove that the induced homomorphism in simplicial homology theory
is well defined, i.e., does not depend on the representative of the orientation.
5.7. Prove the Poincaré lemma in detail (check that the two simplices
[v0 , . . . , v˘j , . . . , v̆i , . . . , vn ] do appear with opposite signs).
5.8. Prove that a simplicial space X is connected if and only if H0 X = Z.
5.9. Compute the homology groups H∗ (M öb, Z) of the Möbius strip directly
from the definition of simplicial homology.
5.10. Compute the homology of RP 2 modulo 2 (i.e., with coefficients in Z2 ).
5.11. Define the homology theory for ordered simplices (in which each geo-
metric n-simplex yields n! ordered simplices) and prove that it is equivalent
to the theory involving all oriented simplices and to the ordered theory, in
which the vertices of the simplicial space are ordered (so that there is only
one ordered simplex corresponding to each geometric simplex).
45
Lecture 6
PROPERTIES OF SIMPLICIAL HOMOLOGY
The aim of this lecture is to establish some basic properties of the simpli-
cial homology groups. We begin, however, with some algebraic preliminaries
mostly related to chain complexes (short exact sequences of chain complexes
and the corresponding long homology sequences, some auxiliary statements
such as Steenrod’s five-lemma, chain homotopy, etc.). We also discuss the
notion of acyclic support, a useful geometric tool that yields important in-
formation in algebraic topology, e.g. in constructing chain homotopies.
After this is done, it turns out that a series of fundamental results are
obtained without much extra work. They are the homotopy invariance (and
hence the topological invariance) of homology groups, the exact homology
sequence for pairs, the Hurewicz theorem (which establishes a fundamental
relationship between homology and homotopy groups), the Mayer–Vietoris
sequence (which often allows to compute the homology of a space when we
know the homology of its parts).
i h
0 −−−→ Ker h −−−→ G −−−→ Im h −−−→ 0,
ϕ ψ
0 −−−→ A −−−→ B −−−→ C −−−→ 0,
i p
0 −−−→ A −−−→ A ⊕ C −−−→ C −−−→ 0,
where i is the natural inclusion and p the projection on the second factor,
is equivalent to the condition that ϕ has a left inverse (i.e., there exists a
homomorphism Φ : B → A such that ϕ ◦ Φ = idB ) as well as to the condition
ψ has a right inverse (i.e., there exists a homomorphism Ψ : C → B such
that Ψ ◦ ψ = idB ).
The proof of this lemma is an (easy) exercise.
The next algebraic lemma is used in homotopy and homology theories to
prove that different spaces have isomorphic homotopy (homology) groups. A
few instances of its application will appear in the exercise class.
Lemma (Steenrod’s Five-Lemma). In the commutative diagram
f g h i
A −−−→ B −−−→ C −−−→ D −−−→ E
p q r s
y y y y yt
f0 g0 h0 i0
A0 −−−→ B 0 −−−→ C 0 −−−→ D0 −−−→ E 0
let the rows be exact and the vertical homomorphisms p, q, s, t be isomor-
phisms. Then r, the middle vertical arrow, is also an isomorphism.
The proof is a nice exercise about abstract groups and diagram chasing.
In fact, the assertion of this lemma holds under weaker conditions. This will
be discussed in the exercise class.
47
· · · → Hi (C) → Hi (C 0 ) → Hi (C 00 ) → Hi (C) → · · · → H0 (C 00 ).
ϕk , ψk : Ck (X) → Ck (Y )
Since A(∆0 ) is acyclic, the chain bi ∆0i must be the boundary of some
P
1-chain (which we denote D0 (a∆0 )) lying in A(∆0 ), i.e.,
D0 (a∆0 ) = aD0 (1 · ∆0 ).
50
Now all the simplices of ∂k (∆) are contained in ∆, and therefore A(∆) sup-
ports the chain ∂k (∆) and so supports the chain Dk−1 ∂k (∆). Thus A(∆)
supports the chain ck and the induction hypothesis implies that
The support A(∆) is acyclic, therefore the cycle ck is the boundary of some
chain (which we denote Dk (∆)) supported by A(∆) and satisfying the equal-
ity ∂k+1 Dk (∆) = ck , as required.
Theorem (Homology sequence for pairs). For any simplicial pair (X, A),
we have the following exact sequence:
i ∗ j∗ ∗ ∂ ∗ i
. . . −−−→ Hn (X) −−−→ Hn (X, A) −−−→ Hn−1 (A) −−−→ Hn−1 (X)
j ∂n−1 i j
−−−→ Hn−1 (X, A) −−−→ . . . −−−→ H0 (X) −−−→ H0 (X, A).
Proof. Given a simplicial pair (X, A), consider the three chain complexes
C(A), C(X), C(X, A). There is a morphism of the first complex to the second
one (inclusion), and of the second one to the third (factorization), yielding
the sequence of chain complexes 0 → C(A) → C(X) → C(X, A) → 0, which
can be written in more detail as
y y y
0 −−−→ Cn (A) −−−→ Cn (X) −−−→ Cn (X, A) −−−→ 0
j
∂ ∂ ∂
y∗ y∗ y∗
0
0 −−−→ Cn−1 (A) −−−→ Cn−1 (X) −−−→ Cn−1 (X, A) −−−→ 0
j
y y y
But this is a short exact sequence of chain complexes, and it gives the required
long exact sequence by the Short to Long Exact Lemma.
Now it is easy to see that this short sequence is exact. Using the Short to Long
Exact Lemma, we obtain a long homology sequence, which is exactly the one
appearing in the statement of the theorem.
6.8. Problems
Lecture 7
SINGULAR HOMOLOGY
The setP
Cn (X) has a natural
P 0 G-module structure, with the sum of two
0
chains c = gα σα and c = gα σα defined by
X
c + c0 := (gα + gα0 )σα
57
P
and multiplication by a constant λ ∈ G by λc = (λgα )σα .
Next we define the boundary operator ∂n : Cn (X) → Cn−1 (X) on each
simplex σ by setting
n
X
∂n (σ) = (−1)k σhki,
k=0
where σhki := σ|∆hki denotes the restriction of σ to the kth face of ∆ (which
is of course a singular (n − 1)-simplex); then we extend ∂n to the whole group
Cn (X) by linearity.
Doing this for all n ≥ 0, we obtain a sequence of Abelian groups (in fact,
G-modules) and homomorphisms
∂n+1 n∂ ∂n−1 1 ∂
. . . −−−→ Cn (X) −−−→ Cn−1 (X) −−−→ . . . −−−→ C0 (X),
called the complex of singular chains of X and denoted C = (Cn , ∂n ) .
Lemma (Poincaré) The boundary operator of singular chains satisfies
Now let us consider pairs of topological spaces (X, A) and their maps.
By a map (X, A) → (Y, B) of such pairs we mean a continuous map f from
X to Y such that f i(A) ⊂ B; we write f : (X, A) → (Y, B) in that case.
Any pair (X, A) of topological spaces defines thecorresponding relative
singular chain complex C(X, A) = Cn (X)/Cn (A), ∂n (here by abuse of no-
tation we write ∂n for the boundary operator in C(X), which is well defined
on cosets mod Cn (A)). The homology of this chain complex is called the
relative singular homology of the pair (X, A). We write
∞
M ∞
M
H∗ (X, A; G) = Hn (X, A; G) := Hn (C(X, A)).
n=0 n=0
The relative homology group Hn (X, ∅; G), which can be identified with
Hn (X; G), is sometimes called the absolute homology group of X.
59
X 7→ Hn (X), n = 0, 1, 2, . . .
(X, A) 7→ Hn (X, A), n = 0, 1, 2, . . .
(X, A) 7→ ∂∗ : Hn (X, A) → Hn−1 (A), n = 1, 2, . . .
f :X→Y 7→ f∗ : Hn (X) → Hn (Y ), n = 0, 1, . . .
f : (X, A) → (Y, B) 7→ f∗ : Hn (X, A) → Hn (Y, B), n = 0, 1, . . .
these objects will be called graded homology groups of X and (X, A).
Theorem. The correspondences described above define a covariant func-
tor (called singular homology) from the category T op of topological spaces to
that of graded Abelian groups, and this functor has the following properties.
(I) Dimension: H0 (pt) = G, where pt is the singleton and G is an Abelian
group, and Hn (pt) = 0 for n > 0.
60
is commutative.
(III) Homotopy invariance: if two maps f, g : (X, A) → (Y, B) are homo-
topic, then f∗ = g∗ , the induced homomorphisms coincide in all dimensions,
and so homotopy equivalent spaces have the same homology.
(IV) Exactness: For any pair of spaces (X,A), the following sequence is
exact:
i ∗ j∗ ∗ ∂
∗ i
. . . −−−→ Hn (X) −−−→ Hn (X, A) −−−→ Hn−1 (A) −−−→ Hn−1 (X)
j∗ ∗ ∂ ∗ i j∗
−−−→ Hn−1 (X, A) −−−→ . . . −−−→ H0 (X) −−−→ H0 (X, A).
(V) Excision: Suppose that U ⊂ X is an open subset whose closure lies in
the interior of A, where A ⊂ X. Then the inclusion (X \ U, A \ U ) ,→ (X, A)
induces an isomorphism in homology Hn (X \ U, A \ U ) → Hn (X, A) in all
dimensions n.
Proof. Property (I) is obvious.
Property (II) follows by a straightforward verification of definitions.
Let us prove (III) in the particular case A = B = ∅ (the general case is
quite similar). Given homotopic maps f, g : X → Y , we shall construct a
chain homotopy
Dk = Ck (X) → Ck+1 (Y );
this will imply f∗ ≡ g∗ by the Chain Homotopy Lemma. Let H : X × I → Y
be a homotopy between f and g. To construct Dk , consider a singular simplex
σ : ∆k → X and the Cartesian product ∆k × I =: T . The set T has a
canonical triangulation consisting of (k + 1)-simplices all of whose vertices
lie in ∆k × {0} and ∆k × {1}. For k = 1 and k = 2, the simplices are shown
in Fig.1 below.
For an arbitrary k, we denote by 0, 1, . . . k the vertices of ∆k , by 00 , 10 , . . . k0
and 01 , 11 , . . . k 1 those of ∆k ×{0} and ∆k ×{1}, respectively. Then a generic
(k + 1)-simplex of T is of the form [00 , . . . j0 , j 1 , . . . k 1 ]; note that the number
61
j of the last vertex lying in ∆k × {0} is the same as that of the first vertex
lying in ∆k × {1}.
1 1
0 0
∆ ∆
Figure 7.1. Triangulations of ∆k × I
∆k+1 = [0, 1, . . . , k + 1]
that takes j and (j + 1) to τ (j)0 and τ (j)1 , respectively. Thus the right-
hand side of the previous displayed formula is a chain in Ck+1 (Y ). Having
constructed Dk on an arbitrary singular k-simplex of X, we extend it by
linearity to chains from Ck (X).
The fact that this construction indeed yields a chain homotopy, i.e., that
∂k+1 Dk + Dk+1 ∂k = g∗ − f∗ ,
Proof. The argument is similar to the proof in the simplicial case (it is
based on the Short-to-Long Exact Lemma), except that the third term of the
short exact sequence is C∗ (X1 ) + C∗ (X2 ) rather than C(X1 ∪ X2 ).
64
7.4. Problems
Lecture 8
APPLICATIONS OF HOMOLOGY
8.1. Connectedness
Homology theory gives a simple characterization of path connectedness
for arbitrary topological spaces and indicates the number of path connected
components of arbitrary topological spaces.
Theorem. A topological space X is path connected if and only if the
0-homology group H0 (X; Z) is isomorphic to Z; this condition can be replaced
by H0 (X; G) ∼= G provided that the coefficient group G is a ring with unit.
Proof. To prove necessity, choose a basepoint p in X. Then any other
point q can be joined to p by a path, this path may be regarded as a singular
simplex ∆1 ; then the 1-chain 1 · ∆1 has the boundary q − p, so that all
points (regarded as 0-chains) are homological to p, and therefore p generates
H0 (X) ∼= Z. The proof of sufficiency is left for the exercise class.
Corollary. The number of path connected components of an arbitrary
topological space X is equal to the dimension of the linear space H0 (X; R).
Proof. Because of the previous theorem, it suffices to prove that no
two vertices located in different components are homologous. Assuming the
converse, one can easily obtain a path joining the two vertices.
66
8.2. Orientability
Let M n be a triangulated n-dimensional manifold, i.e., a simplicial space
each point of which possesses a neighborhood homeomorphic to Rn . We
assume known that any smooth manifold has a triangulation and so does
any topological manifold of dimension two or three (the last fact is actually
a very difficult theorem proved by Edwin Moise in the 1940ies). We say that
M n is oriented if a coherent orientation of its n-simplices is given, i.e., all its
n-simplices are oriented in such a way that the two orientations induced on
any (n − 1)-simplex by the two adjacent n-simplices are opposite. If such an
orientation exists, the manifold is called orientable. It is easy to see that if
M n is connected, then there are only two possible choices of orientation on
M n.
The simplest example of a non-orientable manifold is the projective plane
RP 2 , while the n-sphere Sn is, of course, orientable.
Theorem. A connected triangulated n-dimensional manifold M n is ori-
entable if and only if its n-homology group Hn (M n ; Z) is isomorphic to Z.
Proof. To prove the “if part”, consider a coherent orientation of all the
n-simplices of M n and take the chain c1 ∈ Cn (M n ; Z) with coefficient +1 on
each of them. This chain is obviously a cycle, and so is any chain obtained
by replacing the plus ones by a fixed integer z ∈ Z. There are no other
cycles, because if the chain has non-constant coefficients, then (using the
fact that Mn is connected) we can find two adjacent n-simplices with different
coefficients; in that case the boundary of the cycle will be nonzero on their
common face. The cycle c1 clearly generates the group Hn (M n ; Z) ∼ = Z.
The “only if” part will be proved in the exercise class.
For an orientable smooth n-manifold M , the choice of the generator of
Hn (M ) ∼= Z defines one of the two orientations on M ; this generator is called
the fundamental class of M . If M and N are both oriented and f : M → N
is a map, then the image under f∗ of the fundamental class ϕ ∈ Hn (M ) is of
the form d ψ, where ψ ∈ Hn (N ) is the fundamental class of N and d is an
integer; this integer is called the degree of f and denoted deg(f ). The fact
that in the case M = N = Sn this definition agrees with the geometric one
for neat maps given in Lecture 4 will be proved in the exercise class.
where b ∈ N is a nonnegative integer, called the rank of G, the Z(i) are copies
of Z, and T is a finite Abelian group called the torsion of G.
The above algebraic fact immediately implies the following theorem.
Theorem. If X is a finite simplicial space, then its integer homology can
be expressed as follows:
b
M
Hk (X; Z) = Z(i) ⊕ T,
i=1
χ(X) := k0 − k1 + k2 − · · · + (−1)n kn .
This integer invariant is the oldest one in topology, it was actually first in-
vented and computed by Descartes (and only a century later by Euler) in
the case of convex polyhedra. Apparently Riemann was the first to general-
ize it to other simplicial spaces, in particular to two-dimensional manifolds
(Riemann surfaces). The reader should know that χ(M ) classifies orientable
2-manifolds M .
Consider the sequence
i ∂
0 −−−→ Zk (X; F) −−−→ Ck (X; F) −−−→ Bk−1 (X; F) −−−→ 0,
68
where F is a field and Ck , Zk , Bk−1 are the chain, cycle, and boundary groups,
respectively; its exactness implies that dim Zk + dim Bk−1 = dim Ck . On the
other hand, we obviously have dim Ck = ak . Therefore
X X X
χ(X) = (−1)k Ck = (−1)k Zk + (−1)k Bk−1 =
X X
= (−1)k dim Zk − dim Bk−1 = (−1)k dim Hk (X; F).
Proof. Consider the chain complex Ck (X; R), ∂k ; denote Zk := Ker ∂k
and Bk := Im∂k+1 . For an appropriate subspace C bk of the vector space Ck ,
we can write Ck = Zk ⊕ Cbk . The operator fk maps Zk to itself and therefore
bk → C
we have a linear operator fbk : C bk such that
Summing this over k with alternating signs, we obtain (1), because the
first and third sum in the right-hand side cancel each other out.
Theorem (Lefschetz). A continuous map f : X → X of a finite simplicial
space to itself with nonzero Lefschetz number Λ(f ) 6= 0 has a fixed point.
Proof. Suppose that the map f has no fixed points. By compactness,
there is a positive lower bound for the distance between points and their
images. Therefore, if we take the iterated baricentric subdivision of X a
sufficient number of times (denoting by X 0 the obtained simplicial space)
and consider the simplicial approximation ϕ : X 0 → X 0 of f , we can assume
that ∆0 ∩ ϕ(∆0 ) = ∅ for any simplex ∆0 ∈ X 0 . But then the diagonal of the
70
assigning to each point x ∈ S2k the intersection with S2k of the ray issuing
from the origin and passing through the end point of the vector v(x).
v(x)
O
f (x)
8.5. Problems
Lecture 9
COHOMOLOGY
Cohomology groups are dual to homology groups in the same sense that
covectors are dual to vectors: they are linear functionals on homology. At
first glance, it seems useless to construct a dual theory which is, in a sense,
equivalent to the original one (in particular, it satisfies a dual version of the
same Steenrod–Eilenberg axioms). However, it turns out that in cohomology
theory there is a multiplication operation (the cup product) which has much
better properties than the corresponding operation (the cap product) in ho-
mology. Moreover, cohomology is the natural setting for other operations
(such as Steenrod squares) and for such constructions as the Poincaré iso-
morphism, and it coincides with the purely analytic cohomology of de Rham
defined on smooth manifolds by means of differential forms.
9.1. Definitions and constructions
For the sake of simplicity, we will construct cohomology theory for the
case of finite simplicial complexes, although the general case of arbitrary
topological spaces can be treated in the same way with a few modifications.
(Beware: some modifications are a bit delicate, e.g. the question of dualizing
infinite sums of Abelian groups.) You will see in what follows that in practice
the dualization of homology theory is formally quite simple, and it consists
in lifting the indices and reversing the arrows.
Let X be a simplicial space and let G be an Abelian group. We shall
write Ck (X) = Ck (X; Z) for the simplicial chain group. A homomorphism
ck : Ck → G is called a k-dimensional cochain with values in G. Cochains ob-
viously form a group (with the group structure “inherited” from G) denoted
by C k (X; G). In more compact notation, C k (X; G) := Hom(Ck (X), G).
Let ck ∈ C k (X; G) and ck ∈ Ck (X). We denote the value of the homo-
morphism ck on the chain ck by hck , ck i and define the coboundary operator
δ : C k (X; G) → C k+1 (X; G), hδck , ck+1 i = hck , ∂ck+1 i.
(Here we have omitted the index k in the notation for δ and ∂ because it is
clear from the context.) To compute the value of a k-cochain, it suffices (by
linearity) to know its value on a k-simplex [v0 , . . . , vk ]. This value is given
by the formula
k+1
X
δc([v0 , . . . , vk ]) = (−1)i hck , [v0 , . . . , vi∨ , . . . vk ]i;
i=0
74
the (easy) proof of this formula will be discussed in the exercise class.
The Poincaré Lemma ∂ ◦ ∂ = 0 implies δ ◦ δ = 0, and so we can define
the cohomology groups
Z k (X; G)
H k (X; G) := , H 0 (X; G) := Z 0 (X; G) ,
B k (X; G)
where Z k and B k are the kernel and image of the corresponding cobound-
ary operators in C k ; elements of Z k are called cocycles and those of B k ,
coboundaries.
We defined cohomology groups on the chain-cochain level, however, for
the case of fields (in that case the homology groups are finite-dimensional
linear spaces), we could have defined them directly as dual spaces to the
homology spaces. This fact is expressed by the following statements, whose
simple proofs are omitted.
Theorem (Duality) If G is the additive group of a field, then H k (X; G)
is the dual space of Hk (X; G).
Corollary. If G is the additive group of a field and the space Hk (X; G)
is finite-dimensional, then H k (X; G) ∼= Hk (X; G).
This corollary begs the following question: what is the use of cohomology
groups if they are (isomorphic to) the homology groups? The answer to
that question was given at the beginning of this lecture – because of the
multiplication operation (see Sec.9.4).
The remaining basic steps of the theory, namely the definition and/or
construction of relative cochains, relative cohomology groups, augmentation,
reduced cohomology, induced homomorphisms in cohomology for spaces and
pairs of spaces, the coboundary operator δ ∗ : H k (X, A; G) → H k (X; G) is
similar (or rather dual) to the corresponding definitions and/or constructions
in homology theory and are left as exercises for the reader.
X 7→ H n (X), n = 0, 1, 2, . . .
(X, A) 7→ H n (X, A), n = 0, 1, 2, . . .
(X, A) 7→ ∂ ∗ : H n (X, A) → H n−1 (A), n = 1, 2, . . .
f :X→Y 7 f ∗ : H n (Y ) → H n (X), n = 0, 1, . . .
→
f : (X, A) → (Y, B) 7→ f ∗ : H n (Y, B) → H n (X, A), n = 0, 1, . . .
these objects will be called graded cohomology groups of X and (X, A).
Theorem. The correspondences described above define a contravariant
functor (called the cohomology functor) from the category of finite simplicial
spaces to that of graded Abelian groups. The contravariance of the functor
means that
(f ◦ g)∗ = g ∗ ◦ f ∗ and (idX )∗ = idH n (X) .
This functor has the following properties.
(I) Dimension: H0 (pt) = G, where pt is the singleton and G is an Abelian
group, and Hn (pt) = 0 for n > 0.
(II) Commutation: the following square diagram
is commutative.
(III) Homotopy invariance: if two maps f, g : (X, A) → (Y, B) are homo-
topic, then f ∗ = g ∗ , the induced homomorphisms coincide in all dimensions,
and so homotopy equivalent spaces have the same cohomology.
76
(IV) Exactness: For any pair of spaces (X,A), the following sequence is
exact:
i j∗ δ∗ i
. . . ←−∗−− H n (X) ←−−− H n (X, A) ←−−− H n−1 (A) ←−∗−− H n−1 (X)
j∗ δ∗ i j∗
←−−− H n−1 (X, A) ←−−− . . . ←−∗−− H 0 (X) ←−−− H 0 (X, A).
(V) Excision: Suppose that U ⊂ X is an open subset whose closure lies in
the interior of A, where A ⊂ X. Then the inclusion (X \ U, A \ U ) ,→ (X, A)
induces an isomorphism in cohomology
H n (X, A) → H n (X \ U, A \ U )
in all dimensions n.
Proof. The proofs of all five assertions of this theorem follow by dual-
ity from the five assertions in the corresponding theorem in homology (see
Sec.7.2). For example, to prove (IV) (exactness), we dualize the short exact
sequence of chains
∗ i p∗
0 −−−→ C(A) −−−→ C(X) −−−→ C(X, A) −−−→ 0,
which is also exact (this follows from the Splitting Lemma applied to the
chain sequence). The (long) exact sequence for cohomology is then obtained
by applying the Short-to-Long Exact Lemma.
Remark. We have formulated the Steenrod–Eilenberg axioms in the
finite simplicial setting; they also hold in the topological setting, and there
are uniqueness theorems similar to those in homology theory.
the given ordering of the vertices of X). It is not hard to prove that this
modification yields the same homology groups and induced homomorphisms
(see Exercise 5.11 above).
Let X be a finite simplicial space and let the coefficient group G be a
commutative associative ring with unit (e.g. Z). Recall that we defined the
chain complex Cn (X), ∂n by using oriented simplices (which form the basis
of each G-module Cn ). We will now consider ordered simplices (v0 , . . . vn ),
where the order of the vertices vi is fixed but the vi are not necessarily
pairwise different and define the ordered chain complex C bn (X; G), ∂n by
taking Cbn (X) to be the linear combination (with coefficients in G) of ordered
simplices and using the same boundary operator ∂ as in the usual theory (see
5.4). The rest of the ordered (co)homology theory is constructed similarly to
the usual simplicial (co)homology theory. Note, however, that the chain
complex in the ordered theory is “infinite to the left” (because repetition of
vertices forces us to consider simplices of arbitrarily high “dimension”), but
this chain complex has no nontrivial homology in dimensions higher than the
dimension of X. Moreover, we have the following theorem.
Theorem. The homology of the two complexes C b∗ (X; G) and C∗ (X; G)
is canonically isomorphic.
We omit the straightforward proof of this theorem.
The cup product is obviously bilinear and associative, it has a unit defined
as the cochain taking the value 1 (where 1 is the unit of the ring G) at
all vertices of X. It can be carried over to cohomology classes due to the
following beautiful (and easily verified) lemma.
Lemma (“Leibnitz rule”). δ(cp ^ cq ) = (δcp ) ^ cq + (−1)p cp ^ δ(cq ).
The proof of this lemma, as well as the fact that it allows to correctly
define the cup product on the cohomology level, is a straightforward exercise.
It is also easy to prove that
f ∗ (α ^ β) = f ∗ (α) ^ f ∗ (β).
78
Theorem. The cup product supplies the graded group H ∗ (X; G) with
the structure of a graded G-module; the cup product is skew commutative
in the sense that
α ^ β = (−1)pq (β ^ α).
ω1 ∧ ω2 = (−1)pq (ω2 ∧ ω1 )
and it supplies the graded vector space Λ∗ M n with the structure of a graded
R-algebra. Any smooth map f : M n → N m induces a linear map
f ∗ : Λk N m → Λk M n .
d : Λk M n → Λk+1 M n .
The equality d ◦ d = 0 implies that any exact from is closed, and the quotient
space of the space of closed k-forms by exact k-forms is, by definition, the de
k
Rham cohomology HdR (M n ) in dimension k.
The exterior multiplication operation can be carried over from the level
of forms to the cohomology level, because here also the Liebnitz rule holds:
for any k-form ω1 and any l-form ω2 ,
9.5. Problems
9.1. Prove that
k+1
X
δc([v0 , . . . , vk ]) = (−1)i hck , [v0 , . . . , vi∨ , . . . vk ]i;
i=0
Lecture 10
POINCARÉ DUALITY
σσσ111111
σ∗∗11∗∗11∗∗11
σσσ
σ
σσσ000000
σσσ∗∗22∗∗22∗∗22
Note that that each simplex σ intersects its dual cell transversally, their
only intersection point being the baricenter of σ.
We have not only supplied M n with a CW-space structure (which we
denote by M∗ ), but we have defined a bijection between the k-simplices σ
of the given triangulation and the dual (n − k)-cells σ∗ of M∗ . Moreover, if
the orientation of M n is fixed arbitrarily (i.e., a continuously varying basis,
which we will call positively oriented, is given at each point), then each ori-
ented simplex σ determines a corresponding orientation of the dual cell in
the natural way: we orient the dual cell so that its basis (attached to the
baricenter) added to the orienting basis of σ has the same orientation as the
positive basis of M n at the baricenter.
This allows us to construct, for any k = 0, 1, . . . n a homomorphism
δ : Ck (M ; Z) → Cn−k (M∗ ; Z)
are commutative. This last fact implies in turn that the corresponding ho-
mology and cohomology groups are isomorphic. Thus we have proved the
following theorem.
Theorem (Poincaré Duality). Let M n be a smooth oriented compact
closed manifold and R be a commutative associative ring with unit. Then
the homology and cohomology groups of complementary dimensions are iso-
morphic, i.e., for all k = 0, 1, . . . , n,
Hk (M ; R) ∼
= H n−k (M ; R)
The theorem does not hold for the case in which the manifold M is non-
orientable, but if we replace the ring R by Z2 (the integers mod 2), then we
still have an isomorphism, namely
Hk (M ; Z2 ) ∼
= H n−k (M ; Z2 ) for all k = 0, 1, . . . , n.
a _ b := ϕ−1
p+q ϕ p (a) ^ ϕ q (b) .
From this definition and from the properties of the cup product, it imme-
diately follows that the homology H∗ (M ) of a smooth compact closed man-
ifold M supplied with the cap product is a graded R-algebra with skew-
commutative multiplication.
This statement is not, as would seem at first glance, a useless dualiza-
tion. It turns out that the cap product has a simple very visual geometrical
interpretation, which sometimes makes its computation simpler than that of
the cup product. Also, as we shall see shortly, it gives a simple description
of Poincaré’s duality isomorphism on the homology-cohomology level.
Now let us give a sketch of the geometric interpretation of the cap product.
Suppose a ∈ Cp (M ) and b ∈ Cq (M ) are cycles. To simplify the exposition,
we assume that a and b are cycles with coefficients 1 on some simplices and
zero on the others; let A be the set of all simplices with coefficient 1 for the
cycle a and B the corresponding set for b. Further assume that A and B
are in general position, which means that two (open) simplices σ p ∈ A and
τ q ∈ B either do not intersect at all or intersect in an open cell of dimension
p + q − n. Then A ∩ B is a cell space. Orient all the cells of A ∩ B as follows:
first to an orienting basis of σ p ∩ τ q (it has p + q − n vectors) add n − p vectors
to form an orienting basis of σ p , then add n − q vectors to form an orienting
basis of τ q , in such a way that the n vectors thus obtained determine the
chosen orientation of M n (Fig.10.2).
333333
2 τq
σp
_ : Hp (M n ; R) × Hq (M n ; R) → Hp+q−n (M n ; R).
The homology cap product discussed above is more visual but also less
general than its other version, the cohomology-homology cap product (usu-
ally referred to simply as the cap-product). This operation, also denoted by
_, is defined for the homology and cohomology of path connected simpli-
cial spaces X (not only smooth manifolds) with coefficients in an associative
commutative ring R with unit. We will be using ordered simplices; to fix
an order, we number all the vertices of X once and for all; the vertices of
any simplex will always be listed in increasing order. In this setting, the
cap-product is a bilinear map
defined as follows.
Suppose that cp ∈ C p (X; R); as usual, we first define the cap product on
simplices; we put
as required.
lk(C1 , C2 ) := [d _ c1 ] ∈ H0 (S3 ) = Z.
The fact that the linking number is well defined (does not depend on the
choice of d) is the subject matter of an exercise.
Actually lk(C1 , C2 ) is a so-called finite-type invariant in the sense of Vas-
siliev and can be computed in an elementary way from the diagram of the
link C1 ∪ C2 ⊂ S3 by counting the “signs” of the crossing points of the two
curves (as shown in Fig.3)
+ −
+
+
−
− lk(c1 , c2 ) = −2
−
Figure 10.3. Counting the Gauss linking number
88
10.6. Problems
10.1. Let ϕ : Ck (M ) → C n−k (M∗ ) be the natural homomorphism between
the simplicial chain group of a smooth closed manifold M n and the cochain
group of the dual cell decomposition M∗ of M n . Prove the following commu-
tation relation δn−k ◦ ϕk = ϕk−1 ◦ ∂k , where ∂ and δ are the boundary and
coboundary operators.
10.2. Prove that if H1 M 3 = 0 for a closed 3-manifold M 3 , then M 3 is a
homology 3-sphere, i.e., its homology groups as S3 .
10.3∗ . Give an example of a homology 3-sphere which is not homeomorphic
to the 3-sphere.
10.4. Let the torus be presented as a CW-space with four cells, let c1 be
the cochain taking one of the 1-cells to 1 and the other to 0, while c1 is the
chain with coefficient 1 at the second 1-cell and 0 at the other. Compute
c 1 _ c1 .
10.5. Prove the Leibnitz rule for the cap product.
10.6. Prove the following properties of the cap product:
(i) f∗ (f ∗ (b) _ a) = b _ f∗ (a), where f : X → Y is a simplicial map,
a ∈ H∗ X, and b ∈ H ∗ Y ;
(ii) haq , bp _ cp+q i = haq _ bp , cp+q ;
(iii) ap _ (bq _ cp+q+r ) = (ap ^ bq ) _ cp+q+r .
10.7. Give an example of two simplicial complexes with the same homology
but with different cap products.
10.8. Prove that the GaussP linking number of two oriented circles embedded
3
in R given by lk(C1 , C2 ) = εi , where ε = ±1 is as in Fig.3, is well defined.
10.9. Prove the Poincaré duality theorem for nonoriented manifolds.
89
Lecture 11
OBSTRUCTION THEORY
f∗ : πn (B n ) → πn (X)
Their composition f∗ ◦ h−1 takes the homology class determined by the cycle
∂(∆n+1 ) =: z to cn+1 (f )(∂∆n+1
P
i i ). But
X
∂(∆n+1
i ) = ∂∂∆n+2 = 0,
i.e., the cycle z is zero and therefore its image by the homomorphism f∗ ◦h−1 is
also zero as claimed.
b n → Y can be extended to X
Theorem (Eilenberg) The map f : X b n+1 if
and only if Γn+1 (f ) = 0.
We omit the detailed proof of this theorem, which is rather straightfor-
ward. Indeed, in order to extend the map f , we must be able to extend it
to each (n + 1)-simplex from its boundary (on which f is already defined).
Roughly speaking, the condition Γ = 0 says that the image of the boundary
of each such simplex is homotopy trivial in Y , so that the extension of the
map to the (n + 2)-simplex is possible.
X, Y ←→ H n (X; πn (Y )).
Examples:
• the circle S1 is a K(Z, 1) space;
• any surface M 2 other than S2 and RP 2 is a K(π1 (M 2 ), 1) space;
• RP ∞ is a K(Z2 , 1) space;
• CP ∞ is a K(Z2 , 2) space;
• the infinite-dimensional lens space L∞ m is a K(Zm , 1) space.
The validity of these specific examples will be discussed in the exercise
class. The existence of many other K(π, n) spaces will be established in the
next theorem.
Theorem (Existence of K(π, n) spaces). For any finitely presented group
π and any n there exists a K(π, n) space.
Proof. The proof is by a direct construction, in which we first construct
a space with zero homotopy groups in all dimensions up to n−1, then supply
it with the necessary n-homotopy group (isomorphic to π), and finally kill
all the higher homotopy inductively by “chasing them away to infinity”.
Let hg1 , . . . , gr |R1 = · · · = Rs = 1i be a presentation of π. Denote by K n
the wedge of r copies of the (triangulated) n-sphere. Take s copies of the
(n + 1)-disk and glue it to the wedge of spheres in accordance to the relations
R1 , . . . , Rs , obtaining a simplicial space that we denote by K n+1 . Clearly,
Hn (K n+1 ) ∼ = π. By the Hurewicz Theorem, we have πn (K n+1 ) ∼ = π.
However, our construction is not finished because K n+1 can have nontriv-
ial higher (than n) homotopy groups. We kill these groups inductively by
gluing disks onto their generators.
Theorem (Uniqueness of K(π, n) spaces). Two K(π, n) spaces with the
same π and n are homotopy equivalent.
Proof. It suffices to prove that any K(π, n) space is homotopy equivalent
to the one constructed in the proof of the Existence Theorem. This is done
inductively on the dimension of the skeletons of the given space, the key
point is the application of obstruction theory, which is easy because the
corresponding obstructions vanish.
93
X, Y ←→ H n (X; π).
H n (K(π, n); π 0 ) ∼
= Hom(Hn (K(π, n); π 0 )),
11.5. Problems
11.1. Prove that Sp × Sq = (Sp ∧ Sq ) f Dp+q , where f is some map
S
f : Sp+q → Sp ∧ Sq non homotopic to the constant map.
11.2. In the notation of Sec.11.2 of the lecture, let f, g : K b n → Y be
two maps that coincide on K b n−1 . Denote by dn (f, g) ∈ C n (K n , A; πn (Y ))
the cochain (called distinguishing) defined on a simplex ∆n as follows: take
two copies of ∆n , glue then along their boundary, map one of the copies via
f and the other via g to Y , thus determining an element of πn (Y ). Prove
that
dn (f, g) = cn (g) − cn (f ).
Use this fact to prove the Eilenberg Theorem.
11.3. Prove that dn (f, g) + dn (g, h) + dn (h, f ) = 0, where dn is the
distinguishing cochain defined in Problem 11.2.
11.4. Prove that S1 is a K(Z, 1) space.
11.5. Let M 2 be a closed surface other than S2 or RP 2 . Prove that M 2
is a K(π1 (M 2 ), 1) space.
11.6. Prove that RP ∞ is a K(Z2 , 1) space.
11.7. Prove that CP ∞ is a K(Z, 2) space.
11.8. Prove that L∞ m is a K(Zm , 1) space, where the infinite-dimensional
lens space L∞ m is defined as follows: ∞ ∞
P 2 consider S as the subspace of C of all
points (z1 , z2 , . . . ) such that |zi | = 1; introduce the equivalence relation
(z1 , z2 , . . . ) ∼ (εz1 , εz2 , . . . ), where ε = exp(2πi/m); then L∞ ∞
m := S / ∼.
11.9. Prove that
(
Zm for odd n;
Hn (L∞ ) =
0 for even n.
Lecture 12
VECTOR BUNDLES AND G-BUNDLES
In this lecture, we study and classify vector bundles (a key notion in dif-
ferential topology, the main example being the tangent bundle of a smooth
manifold). and principal G-bundles (which play an important role in geomet-
ric topology, K-theory, and theoretical physics). Vector bundles are classified
over paracompact spaces by using the partition of unity, a generalization of
the Feldbau Theorem, and the Gauss map to the canonical Grassmann bun-
dle, while G-bundles are classified via the beautiful Milnor construction and
the notion of classifying space of a topological group.
Φ|p−1
1 (b)
: p−1 −1
1 (b) → p2 (ϕ(b))
and f ∗ (p)(b, e) := b.
E1 := (b, e) ∈ X × E : f (b) = p(e)
Ekm := (L, r) ∈ Gm m
k ×R : r ∈ L ,
and γkm is the natural projection (L, r) 7→ L. There are obvious inclusions
m+1
Gmk ⊂ Gk which allow to define G∞ ∞
k (called the Grassmannian) and γk by
passing to the inductive limit.
B, G∞
k ←→ Vectk (B) ,
Rev ⊕ Rod of the subspaces generated by the basis vectors with even and odd
coordinates. For the particular case in which the given bundle is the tangent
bundle τ of a smooth manifold M k , the proof is easy: we embed M k in some
RN , regard GNk as consisting of k-dimensional linear subspaces of R
N
and
k
apply the Gauss map, assigning to each point x ∈ M the linear subspace
obtained by parallel transport of the tangent plane Tx M k to the origin.
A good reference for the proofs is the book Fiber Bundles by D.Husemoller
(Rassloennye prostranstva, M., Nauka, 1970).
ωG : EG → BG = EG /G
is called the universal G-bundle, its base is called the classifying space of the
n
group G. In a similar way, one defines the bundle ωG : EGn → BG n
called
(briefly) the n-universal G-bundle, its base is called the n-classifying space
of the group G (in less shortened form, the classifying space up to dimension
n).
Examples:
(1) The classifying space of the group S1 is CP ∞ and ES1 = S∞ . The
k-classifying space of S1 is CP k and ESk1 = S2k+1 .
(2) The classifying space of the group Z2 is RP ∞ and EZ2 = S∞ . The
k-classifying space of Z2 is RP k and ESk1 = Sk .
between the homotopy classes of maps of B into the classifying space BG and
the isomorphism classes of principal G-bundles over B.
We will not prove this theorem in the present course (see the book by
Hosemueller, loc.cit).
Lecture 13
MISCELLANY
Hk (X; G) ∼
= Hk (X) ⊗ G ⊕ Tor Hk−1 (X), G
H k (X; G) ∼
= Hom Hk (X), G ⊕ Ext Hk−1 (X), G .
Hk (X; G) ∼
= Hk (X) ⊗ G,
and
H k (X; G) ∼
= Hom Hk (X), G .
Hk (X × Y ) ∼
M M
= Hl (X) ⊗ Hm (Y ) ⊕ Tor(Hl (X), Hm (Y )) .
l+m=k l+m=k−1
For the case in which the coefficient group is G, the additive group of one
of the fields Q, R, C, then we have the simpler relation
Hk (X × Y ) ∼
M
= Hl (X) ⊗ Hm (Y ) .
l+m=k
103
e k (M ) ∼
H =H e k (M ) ∼
e n−k−1 (Sn \ M ), H =He n−k−1 (Sn \ M ) .
Chain groups (say with integer coefficients) Ck (Nω ) are defined as the sets
of all linear combinations of k-simplices, the boundary operator is defined as
in simplicial homology, and it yields the homology groups Hk (Nω ).
It should be intuitively clear that if X is a “nice” enough and the covering
ω is “sufficiently fine”, then the nerve Nω is a “good approximation” of the
topology of X so that Hk (Nω ) is a good approximation of the k-homology of
the space X.
Alexandrov’s second main idea, roughly speaking, is based on the fact
that if one covering is inscribed in another, there is a natural projection of
the homology of the inscribed nerve onto the homology of the other nerve,
and the definition of the homology groups is obtained by taking the projective
limit of these groups.
It can be proved that the C̆ech homology functor satisfies the Steenrod-
Eilenberg axioms and therefore coincides, say, with the singular homology
functor.