Algebraic Topology ADR Mam
Algebraic Topology ADR Mam
(Simplicial Homology)
BY
A. DEB RAY
Throughout this Chapter, we consider subsets of Euclidean n-space, Rn , unless stated oth-
erwise.
It is easy to see that the above definition does not depend on the chosen point a0 . i.e.,
{a1 −a0 , a2 −a0 , . . . , ak −a0 } is linearly independent if and only if {aj −ai : 0 ≤ j ≤ k, j ̸= i}
is linearly independent for all i = 0, 1, . . . k.
Observation 1.1.1.
1
A. Deb Ray, Pure Math, C.U. 2
k
X
(λi − ηi )ai = 0 (1.1)
i=0
k
X
(λi − ηi ) = 0 (1.2)
i=0
k
P
From (1.2), (λi − ηi )a0 = 0 and so, subtracting it from (1.1),
i=0
k
X
(λi − ηi )(ai − a0 ) = 0.
i=1
The linear independence of the set {a1 − a0 , a2 − a0 , . . . ak − a0 } implies that λi = ηi , for each
i = 1, 2, . . . k. Using (1.2), λ0 = η0 .
For each x ∈ H k , the k scalars (λ0 , λ1 , . . . , λk ) representing x are called barycentric
co-ordinates of x with respect to the geometrically independent set {a0 , a1 , . . . ak }. The
portion of the k-dimensional plane which contains points with non-negative barycentric co-
ordinates, is called a k-simplex.
A. Deb Ray, Pure Math, C.U. 4
It is to note that a simplicial complex is a set of simplexes, not points ofSRn . But if one
takes the union of all the points lying on the simplexes of K, i.e., |K| = σ, then |K| is
σ∈K
a subspace of the Euclidean space Rn . This space |K| is called the geometric carrier or
underlying space of K.
Observation 1.1.2. |K|, being a subspace of Rn , is a metric space.
Since any simplex is a closed and bounded subset of an Euclidean space it is compact.
|K|, being a finite union of compact sets, is also compact.
Definition 1.1.7. If σ = ⟨a0 , a1 , . . . , am ⟩ is a m-simplex then cl(σ) (read as, closure of σ)
is the simplicial complex consisting of all possible faces of σ and the skeleton of σ (denoted
by sk(σ)) is cl(σ) \ {σ}.
Exercise 1.1.1. Show that both closure and skeleton of a simplex σ are simplicial complexes.
In fact, for any m-simplex with m ≥ 4, we are unable to visualize the geometry of
the simplex σ m . Hence, we need to formulate something algebraically without involving
the geometry of the said simplex. Also, the definition should be a generalization of the
aforementioned natural concept of orientation.
Before defining it formally, we introduce the ordering of vertices as follows:
v0 < v1 < ... < vm if and only if ⟨v0 , v1 , ..., vm ⟩ is a positively oriented simplex
Definition 1.2.1. Suppose σ m = ⟨v0 , v1 , ..., vm ⟩ is any m-simplex and v0 < v1 < ... < vm is
an ordering of vertices associated with it. These (m+1) vertices can be permuted in (m + 1)!
ways. Among these permutations, half are even and half are odd. If Sm+1 denotes the
symmetric group on the set {0, 1, ..., m} of m + 1 symbols, then for any τ ∈ Sm+1 ,
vτ (0) , vτ (1) , ..., vτ (m) is positively oriented, if τ is an even permutation;
vτ (0) , vτ (1) , ..., vτ (m) is negatively oriented, if τ is an odd permutation.
Looking back to the illustrations described above, we find that this definition works! One
needs to be very careful about the fact that orientation of a k-simplex is determined from
the ordering of its k + 1 vertices, not induced from the orientation of some k + 1 simplex of
which it is a face. The following serves as a working rule for determining the orientation of
an m-simplex,
If ⟨a0 , a1 , . . . , am ⟩ is positively oriented then the orientation of σ formed by vertices a0 , a1 , . . . , am
is positively (respectively, negatively) oriented if the number of interchange of vertices that
takes σ back to ⟨a0 , a1 , . . . , am ⟩ is even (respectively, odd).
Definition 1.2.3. Suppose σ q and σ q+1 are any two simplexes in an oriented simplicial
complex K of dimensions q and q + 1 respectively. Let +σ q = ⟨v0 , v1 , . . . , vq ⟩ and v be the
extra vertex of σ q+1 . The incidence number is an integer associated to the pair σ q and σ q+1 ,
denoted by [σ q+1 , σ q ] and defined by
0,
if σ q is not a face ofσ q+1
[σ q+1 , σ q ] = 1, if + σ q+1 = ⟨v, v0 , v1 , . . . , vq ⟩
−1, if − σ q+1 = ⟨v, v0 , v1 , . . . , vq ⟩
Example 1.2.1. Let K = cl (⟨v0 , v1 , v2 ⟩). If σ 2 = ⟨v0 , v1 , v2 ⟩, σ11 = ⟨v1 , v2 ⟩, σ21 = ⟨v0 , v2 ⟩,
and σ 0 = ⟨v0 ⟩, then [σ 2 , σ11 ] = 1, [σ 2 , σ21 ] = −1, [σ11 , σ 0 ] = 0 and [σ21 , σ 0 ] = −1.
Definition 1.3.2. The group (Cq (K), +) is called the q-dimensional chain group of K,
for 0 ≤ q ≤ dim(K). It is customary to define the chain groups as trivial for all q < 0 or
q > dim(K).
We now observe that the following are certain distinguished members of q-dimensional
chain group (Cq (K), +), corresponding to each positively oriented σ q ∈ K.
q q
+1, if τ = σ
(σ q )(τ q ) = −1, if τ q = −σ q
0, otherwise
From the following theorem we now see that for complete description of the chain group
Cq (K), it is enough to know that these particular q-chains, known as elementary q-chains.
In other words, elementary q-chains generate the q-th chain group.
Proof. Consider any f ∈ Cq (K). Since K is a finite simplicial complex, we may list the
positively oriented q-simplexes of K as σ1q , σ2q , ... , σtqP
. Let f (σiq ) = ni , i = 1, 2, ..., t. Then
q t q
by definition
Pt of a q-chain, f (−σi ) = −ni . Choose i=1 ni σ i ∈ Cq (K). We claim that
f = i=1 ni σ qi .
For any τ q ∈ S˜q , there are two possibilities:
(i) τ q is positively oriented. In such case, τ q = σjq for some j. Therefore,
t
!
X
ni σ qi (τ q ) = nj = f (τ q ).
i=1
t
!
X
ni σ qi (τ q ) = −nj = f (τ q ).
i=1
Remark 1.3.1. Theorem 1.3.1 shows that Cq (K) is a direct sum of infinite cyclic groups
generated by the elementary q-chains. Since upto isomorphism, there is only one infinite
cyclic group, e.g., Z, Theorem 1.3.1 tells us that the q-th chain group is a direct sum of as
many copies of Z as the number of positively oriented q-simplexes in K.
Definition 1.3.3. For 0 ≤ q ≤ dim(K), the q-th boundary map is the map ∂q : Cq (K) →
Cq−1 (K) defined on the generators σ q by
X
∂q (σ q ) = [σ q , σiq−1 ]σiq−1
(where the summation is taken over all positively oriented (q − 1)-faces of σ q ), and extended
to the entire chain group by linearity. i.e.,
q
t t
!
X X X
∂q ( ni σiq ) = ni [σ q , σiq−1 ]σiq−1
i=1 i=1 i=0
It is easy to see that ∂q as defined in Definition 1.3.3 is a homomorphism and so, they
are known as boundary homomorphisms. If q < 0 or q > dim(K), then ∂q are taken as
trivial homomorphisms.
Remark 1.3.3. It is to be noted that there are q + 1 vertices in a q-simplex and so, a (q-
1)-face can be obtained by deleting one vertex from the list of q + 1 vertices. Consequently,
the number of (q-1)-faces of a q simplex is q + 1 and hence the summation in Definition 1.3.3
varies from i = 0 to i = q. But one may take the sum over all possible (q − 1) simplexes of
K instead of (q − 1)-faces of σ q , because, for the rest of the summands, the incidence number
will be 0, thereby contributing nothing extra.
So far, we have constructed a doubly infinite sequence of chain groups and boundary
homomorphisms, called a chain complex of K:
∂q+2 ∂q+1 ∂q ∂q−1
· · · → Cq+1 (K) → Cq (K) → Cq−1 (K) → · · ·
As discussed earlier, only a finite portion of this sequence (0 ≤ q ≤ dim(K)) is non-trivial.
A pivotal result on composition of boundary maps follows from Theorem 1.2.1 :
Theorem 1.3.2. ∂q ◦ ∂q+1 : Cq+1 (K) → Cq (K) is trivial homomorphism. i.e., ∂q ◦ ∂q+1 = 0,
for each q.
Proof. It is enough to show that (∂q ◦ ∂q+1 )(σ q+1 ) = 0 for each generator σ q+1 of Cq+1 (K).
X
(∂q ◦ ∂q+1 )(σ q+1 ) = ∂q [σ q+1 , σiq ]σiq
σiq ∈K
X
= [σ q+1 , σiq ]∂q (σiq )
σiq ∈K
X X
= [σ q+1 , σiq ] [σiq , σjq−1 ]
σiq ∈K σjq−1 ∈K
X X
= [σ q+1 , σiq ][σiq , σjq−1 ]
σjq−1 ∈K σiq ∈K
= 0
Hence by using linearity of boundary homomorphisms the result follows.
A. Deb Ray, Pure Math, C.U. 10
As Cq (K) is an abelian group, Zq (K) = Ker(∂q ) and Bq (K) = Im(∂q+1 (K)) are normal
subgroups of Cq (K), known as the group of q-cycles and the group of q-boundaries re-
spectively.
Proof. bq ∈ Bq (K) ⇒ ∃aq+1 ∈ Cq+1 (K) such that bq = ∂q+1 (aq+1 ) ⇒ ∂q (bq ) = ∂q (∂q+1 (aq+1 )) =
(∂q ◦ ∂q+1 )(aq+1 ) = 0 (Using Theorem 1.3.2)
∂q+2 ∂q+1 ∂q ∂q−1
Definition 1.3.5. Let · · · → Cq+1 (K) → Cq (K) → Cq−1 (K) → · · · be the chain complex
of an oriented simplicial complex K. Then
(
Zq (K)/Bq (K), 0 ≤ q ≤ dim(K)
Hq (K) =
0, otherwise
for any q-simplex ⟨v0 , v1 , . . . , vq ⟩ of K and extending the definition linearly on entire K. i.e.,
!
X X
f# ni σiq = ni f# (σiq )
i i
For q < 0 or q > dim(K), f# is defined as the zero map. The collection of maps {f# :
Cq (K) → Cq (L) : q ∈ Z} is called a chain map induced by f .
11
A. Deb Ray, Pure Math, C.U. 12
It is an easy task to verify that these maps are homomorphisms, at every level q. So, we
get a diagram as follows:
∂qK
... / Cq (K) / Cq−1 (K) / ...
f# f#
∂qL
... / Cq (L) / Cq−1 (L) / ...
f# ◦ ∂qK = ∂qL ◦ f#
where ∂qK : Cq (K) → Cq−1 (K) and ∂qL : Cq (L) → Cq−1 (L) are the corresponding boundary
homomorphisms.
Proof. For simplicity of notation, we may sometimes use ∂ to denote both ∂qK and ∂qL and
it will be understood from the context which we are referring to. Also, it is enough to prove
the result on any generator
σ of Cq (K). So, consider any σ = ⟨v0 , v1 , . . . , vq ⟩ from Cq (K).
To show that f# ◦ ∂qK (σ) = ∂qL ◦ f# (σ).
Two possibilities may occur : either all f (vi )’s are distinct or they are not all distinct. In
any case, the dimension of τ = ⟨f (v0 ), f (v1 ), . . . , f (vq )⟩ is ≤ q. We consider three possible
cases separately:
Case 1 : dimension of τ is q.
Case 2 : dimension of τ is q − 1.
Case 3 : dimension of τ is ≤ q − 2.
Now we see that in Case 2 and 3,
In Case 1,
∂qL ◦ f# (⟨v0 , v1 , . . . , vq ⟩) = 0. Therefore, the equality holds for this case too.
Case 2 : Suppose, dimension of τ = q − 1.
This can happen only when there exists a pair of vertices vi and vj such that f (vi ) = f (vj ).
Without loss of generality assume that i < j. Then
f# ◦ ∂qK (σ)
q
!
X
= f# (−1)i ⟨v0 , v1 , . . . , vbi , . . . , vq ⟩
i=0
D E
= (−1)i f (v0 ), f (v1 ), . . . , f[
(vi ), . . . , f (vj ), . . . f (vq )
D E
j [
+ (−1) f (v0 ), f (v1 ), . . . , f (vi ), . . . , f (vj ), . . . , f (vq )
D E
= (−1)i (−1)j−(i+1) f (v0 ), f (v1 ), . . . , f (vj ), . . . , f[ (vi ), . . . f (vq )
D E
+ (−1)j f (v0 ), f (v1 ), . . . , f (vi ), . . . , f[
(vj ), . . . , f (vq )
Here f (vi ) = f (vj ). Since (−1)j and (−1)j−1 are always of opposite sign, the above sum
= 0. Consequently, in this case also, the equality holds.
Corollary 2.1.1. If f# : Cq (K) → Cq (L) is the chain map induced from a simplicial map
f : K → L then
(i) f# (Zq (K)) ⊆ Zq (L)
(ii) f# (Bq (K)) ⊆ Bq (L)
Proof. (i) Let z ∈ Zq (K). Then ∂(z) = 0 and hence (f# ◦ ∂)(z) = 0 ⇒ (∂ ◦ f# )(z) = 0
⇒ ∂(f# (z)) = 0, i.e., f# (z) ∈ Zq (L).
(ii) Let b ∈ Bq (K). Then there exists some a ∈ Cq+1 (K) such that ∂(a) = b. So,
f# (b) = f# (∂(a)) = ∂(f# (a)) ∈ Bq (L).
Using Corollary 2.1.1, we define a map f∗ : Hq (K) → Hq (L) by f∗ (z + Bq (K)) =
f# (z) + Bq (L). This map is well defined, because,
Also, f∗ is a homomorphism:
Theorem 2.1.2. (i) For each q, 0 ≤ q ≤ dim(K), if f# : Cq (K) → Cq (L) and g# : Cq (L) →
Cq (M ) are the chain maps then g# ◦ f# = (g ◦ f )# .
(ii) For each q, 0 ≤ q ≤ dim(K), if I# : Cq (K) → Cq (K) is the map induced from the
identity map I : K → K, then I# is the identity homomorphism.
Proof. It is enough to prove the results on the generators σ ∈ Cq (K).
(i) Let f : K → L and g : L → M be simplicial maps that induce chain maps f# :
Cq (K) → Cq (L) and g# : Cq (L) → Cq (M ). Since g ◦ f : K → M is a simplicial map,
(g ◦ f )# : Cq (K) → Cq (M ) is also a chain map. Let σ = ⟨v0 , v1 , . . . , vq ⟩. Then three
possibilities may arise :
Case 1. All f (v0 ), . . . f (vq ) are not distinct. Then of course, g(f (v0 )), . . . g(f (vq )) are not
distinct too. So,
(g ◦ f )# (σ) = 0
= g# (0)
= (g# ◦ f# )(σ)
Case 2. All f (v0 ), . . . f (vq ) are distinct, but g(f (v0 )), . . . g(f (vq )) are not distinct. Then
(g ◦ f )(v0 ), . . . (g ◦ f )(vq ) are also not distinct. So,
(g ◦ f )# (σ) = 0
= g# (⟨f (v0 ), . . . f (vq )⟩)
= (g# ◦ f# )(σ)
Case 3. All f (v0 ), . . . f (vq ) and all g(f (v0 )), . . . g(f (vq )) are distinct. So all (g ◦ f )(v0 ), . . . (g ◦
f )(vq ) are distinct.
Hence, (g ◦ f )# = g# ◦ f# .
(ii) I# (σ) = ⟨I(v0 ), . . . , I(vq )⟩ = ⟨v0 , . . . , vq ⟩ = σ, as vi ’s are all distinct.
Corollary 2.1.2. (i) For each q, 0 ≤ q ≤ dim(K), if f∗ : Hq (K) → Hq (L) and g∗ : Hq (L) →
Hq (M ) are the homomorphisms induced by f and g respectively then g∗ ◦ f∗ = (g ◦ f )∗ .
(ii) For each q, 0 ≤ q ≤ dim(K), if I∗ : Hq (K) → Hq (K) is the induced by the identity map
I : K → K, then I∗ is the identity homomorphism.
Proof. Follows immediately from the theorem.
Theorem 2.2.1. If K1 and K2 are two oriented simplicial complexes obtained from a sim-
plicial complex K then for each q,
Hq (K1 ) ∼
= Hq (K2 )
Proof. Suppose, σ q ∈ K is a q-simplex. Let σ1q be the oriented q-simplex in K1 arose from σ q
and σ2q be the oriented q-simplex in K2 arose from σ q . Then q q q
( σ2 is either +σ1 or −σ1 . There-
1, if σ1q and σ2q are of same sign
fore, we get a function ϕ : K → {−1, 1} given by ϕ(σ q ) = .
0, otherwise.
Then σ2q = ϕ(σ q )σ1q .
Define f# : Cq (K1 ) → Cq (K2 ) by
!
X q
X
f# mi σ1i = mi ϕ(σiq )σ2i
q
i i
X
= [σ1q , σ1q−1 ]ϕ(σ q−1 )σ2q−1
σ1q−1 ∈K1
X
= [ϕ(σ q )σ2q , ϕ(σ q−1 )σ2q−1 ]ϕ(σ q−1 )σ2q−1
σ1q−1 ∈K1
X
= ϕ(σ q ) [σ2q , σ2q−1 ]σ2q−1
σ1q−1 ∈K1
In the last theorem, we have proved that the homology groups do not depend on the
orientation of the given complex. So, we can choose any suitable orientation on the simplicial
complex whose homology groups we compute.
Example 2.2.1. Consider K = {⟨a0 ⟩ , ⟨a1 ⟩ , ⟨a2 ⟩ , ⟨a0 , a1 ⟩ , ⟨a1 , a2 ⟩ , ⟨a0 , a2 ⟩}. The simplicial
homology groups of K are as follows:
(
Z, if q = 0, 1
Hq (K) =
0, otherwise
Let z ∈ Z1 (K). Then z ∈ C1 (K) such that ∂1 (z) = 0 which implies that z = m0 ⟨a0 , a1 ⟩ +
m1 ⟨a1 , a2 ⟩ + m2 ⟨a0 , a2 ⟩ such that
Exercise 2.2.1. Consider K = {⟨a0 ⟩ , ⟨a1 ⟩ , ⟨a2 ⟩ , ⟨a0 , a1 ⟩ , ⟨a1 , a2 ⟩ , ⟨a0 , a2 ⟩ , ⟨a0 , a1 , a2 ⟩}. Show
that the simplicial homology groups of K are as follows:
(
Z, if q = 0
Hq (K) =
0, otherwise
Example 2.2.2. Consider K = {⟨a0 ⟩ , ⟨a1 ⟩ , ⟨a2 ⟩ , ⟨a0 , a1 ⟩ , ⟨a1 , a2 ⟩ , ⟨a0 , a2 ⟩ , ⟨a0 , a1 , a2 ⟩ , ⟨b0 ⟩ , ⟨b1 ⟩ ,
⟨b0 , b1 ⟩}. Show that the simplicial homology groups of K are :
(
Z ⊕ Z, if q = 0
Hq (K) =
0, otherwise
f g
An exact sequence of the form 0 → N → M → P → 0 is called a short exact
sequence.
f g
A short exact sequence of the form 0 → N → N ⊕ P → P → 0 is called a split exact
sequence.
f g
A short exact sequence 0 → N → M → P → 0 is said to split if M ∼
= N ⊕ P.
Definition 2.3.1. Let K be a simplicial complex and the following a chain complex of K.
∂ ∂q+1 ∂q ∂ ∂
. . . → Cq+1 (K) → Cq (K) → Cq−1 (K) . . . →2 C1 (K) →1 C0 (K) → 0.
Define a map ϵ : C0 (K) → Z on the vertices of K by ϵ(v) = 1 for each vertexv of K and
m
P
then extend the definition to the whole of C0 (K) by extending it linearly. i.e, ϵ ni v i =
i=1
m
P m
P
ni ϵ(vi ), i.e., = ni ). This map is called an augmentation map.
i=1 i=1
We now establish a formula that connects a 0-dimensional homology group with the corre-
sponding 0-dimensional reduced homology group.
A. Deb Ray, Pure Math, C.U. 19
S
Theorem 2.4.2. If K is a simplicial complex then |K| = int (σ) and int (σ)∩int (τ ) = ∅
σ∈K
whenever σ ̸= τ . i.e., {int (σ) : σ ∈ K} partitions the geometric carrier |K| of K.
S
Proof. By definition, |K| = σ and int (σ) ⊆ σ, for each σ ∈ K. Hence, one side of the
σ∈K
desired equality is clear.
Let x ∈ |K|. Then there exists some σ ∈ K such that x ∈ σ. If x ∈ int(σ) then we get what
we need to prove. But if x ∈
/ int(σ) then x ∈ γ where γ is a face of σ. Again, if x ∈ int(γ)
then γ being a member of K, we obtain the result. But if x ∈ / int(γ) then it must be on
some face of γ. Repeating this argument finitely many times,Swe must get a simplexSβ (may
be of dimension 0) of K, such that x ∈ int(β). Hence, x ∈ int(σ). i.e., |K| = σ.
σ∈K σ∈K
Suppose σ ̸= τ in K. Then either there exists one vertex of σ which is not a vertex of τ or
one vertex of τ which is not a vertex of σ. Without loss of generality, we assume that there is
one vertex of σ which is not a vertex of τ . So, σ is not a face of τ . Hence, by Theorem 2.4.1,
int(σ) ∩ τ = ∅. So,
int(σ) ∩ int(τ ) ⊆ int(σ) ∩ τ = ∅.
Proof. Exercise.
Theorem 2.4.4. Let K be a simplicial complex. If |K| has k many topological components
and vi , i = 1, . . . k are vertices of K taken exactly one from each component Ki of |K| then
H0 (K) is a free abelian group with a basis {[vi ] : i = 1, . . . k}, where [vi ] stands for the
homology class corresponding to vi ∈ C0 (K). Moreover, H0 (K) ∼ = ⊕ki=1 Z.
If v ∼ w then Cv = Cw .
If v ≁ w then Cv ∩ Cw = ∅.
Cv ’s are disjoint open connected subsets of |K| and therefore form components of |K|.
Step 2. Choose exactly one vertex vi from each component Cv and construct the set
B = {[vi ] : i = 1, 2, . . . , r} (Here [vi ] denotes the homology class vi + Im∂1 in H0 (K)). We
show that B is a basis of H0 (K).
B generates H0 (K) : Let [w] ∈ H0 (K). Then w ∈ C0 (K) and so w =
P
ni wi , where
i
wi ’s are 0-simplexes of K. Suppose wk ∈ Cvj . Then wk ∼ vj . i.e., There is a sequence
of vertices a0 , a1 , . . . an such that a0 = wk , an = vj and ⟨ai , ai+1 ⟩ is a 1-simplex of K.
n−1
P
Consider c = ⟨ai , ai+1 ⟩ ∈ C1 (K). Then
i=0
n−1
!
X
∂1 (c) = ∂1 ⟨ai , ai+1 ⟩
i=0
n−1
X
= (⟨ai+1 ⟩ − ⟨ai ⟩)
i=0
= ⟨vj ⟩ − ⟨wk ⟩
P r
P
So, vj − wk ∈ Im∂1 ⇒ [vj ] = [wk ]. Hence, [w] = ni [wi ] = mj [vj ]. Therefore, B
i j=1
generates H0 (K).
r r
B is linearly independent : Let
P P
mi [vi ] = 0. Then mi vi ∈ Im∂1 . There exists
i=1 i=1
r
P P
d ∈ C1 (K) such that ∂1 (d) = mi vi . Again d ∈ C1 (K) implies that d = dj , where
i=1 j
dj lie in one component Cvj , j = 1, 2, . . . , r. Hence, ∂1 (dj ) = mj vj . (check!)
If we consider the augmentation map ϵ : C0 (K) → Z given by ϵ(a) = 1 for each vertex
a ∈ K, then for each j = 1, 2, . . . r,
groups of K1 = sk(σ 2 ) and have obtained that H0 (K) = H1 (K) = Z and trivial for all other
q ̸= 0, 1. It is natural to ask whether we can extend these results to achieve homology groups
of cl(σ n ) and sk(σ n ) for any n. In this section we answer this question in affirmative.
Definition 2.5.1. Let K be a simplicial complex and w ∈ Rn be such that any ray emanating
from w intersects |K| at atmost one point. Consider w ∗ K as the collection of all simplexes
of the form ⟨w, a0 , a1 , . . . , an ⟩, where ⟨a0 , a1 , . . . , an ⟩ is a simplex in K and all possible faces
of ⟨w, a0 , a1 , . . . , an ⟩. Then w ∗ K is a simplicial complex called a cone over K with vertex
w.
Exercise 2.5.1. (worked-out in class) Show that w ∗ K is a simplicial complex.
Exercise 2.5.2. (worked-out in class) The geometric carrier |w ∗ K| is path connected (and
hence, connected).
Observation 2.5.1. It is easy to observe, if σ 2 = ⟨a0 , a1 , a2 ⟩ and σ1 = ⟨a1 , a2 ⟩, then cl(σ 2 ) =
a0 ∗ cl(σ1 ).
We may generalize this observation for any n as follows: If σ n = ⟨a0 , a1 , a2 , . . . , an ⟩ and
σn−1 = ⟨a1 , a2 , . . . , an ⟩, then cl(σ n ) = a0 ∗ cl(σn−1 ).
Observation 2.5.2. Clearly, cl(σ 2 ) = a0 ∗ cl(σ1 ) (where σ1 has all the vertices of σ2 except
a0 ). Proceeding similarly, one may obtain cl(σ n ) is a cone over cl(σ n−1 ). Hence, for the
homology groups of cl(σ n ), we need to know the homology groups of a cone w ∗ K over K.
Theorem 2.5.1. Let K be an oriented simplicial complex. If w ∈ Rn is such that w ∗ K is
a cone over K with vertex w then for all q,
(
Z, if q = 0
Hq (w ∗ K) =
0, otherwise.
where cq−1 = dq−1 + ∂q (cq ) is a (q − 1)-chain in K. Then ∂q (zq − ∂q+1 [w, cq ]) = ∂q ([w, cq−1 ])
implies (since, ∂q+1 ◦ ∂q = 0 and ∂q (zq ) = 0),
(
cq−1 − ϵ(cq−1 )w, if q = 1
0=
cq−1 − [w, ∂q−1 cq−1 ], if q > 1
Proof. For each n > 0, |cl(σ n )| = |a0 ∗ cl(σ n−1 )|, where σ n = ⟨a0 , a1 , . . . , an ⟩ and σ n−1 =
⟨a1 , . . . , an ⟩. By Theorem 2.5.1, the result is immediate.
Proof. Consider sk(σ n+1 ) and construct the chain complex as follows:
∂ ∂n−1 ∂ ∂
0 → Cn (sk(σ n+1 )) −→
n
Cn−1 (sk(σ n+1 )) −→ . . . C1 (sk(σ n+1 )) −→
1
C0 (sk(σ n+1 )) −→
0
0
A. Deb Ray, Pure Math, C.U. 24
Cq (cl(σ n+1 )) = Cq (sk(σ n+1 )), for all q = 0, 1, . . . , n. Hence, the q-th homology group of
cl(σ n+1 ) and sk(σ n+1 ) are same for q = 0, 1, . . . n − 1. So, we obtain
(
Z, if q = 0
Hq (sk(σ n+1 )) =
0, q = 1, 2, . . . n − 1.
Certainly, Hq (Sk(σ n+1 )) = 0 for any q > n or q < 0. Therefore, we need to calculate the
homology group Hn (Sk(σ n+1 )). Consider the following portions of the chain complexes:
∂n+1 ∂n /
Zσ n+1 ∼
= Cn+1 (cl(σ n+1 )) / Cn (cl(σ3 n+1 )) = Cn (sk(σ n+1 )) Cn−1 (cl(σ n+1 )) = Cn−1 (sk(σ n+1 ))
Now, Hn (sk(σ n+1 )) = Ker∂n /Im(0) ∼ = Ker∂n and Hn (cl(σ n+1 )) = Ker∂n /Im(∂n+1 ). But
n+1
Hn (cl(σ )) = 0. Hence, Ker∂n = Im(∂n+1 ) = ∂n+1 (Zσn+1 ). Since homomorphic image of
a cyclic group is cyclic, Ker∂n becomes a cyclic subgroup of Cn (clσn+1 ). Also, Cn (clσn+1 )
being a free abelian group, it has no torsion element. Therefore, Ker∂n can not be a finite
cyclic group. (For if it is a finite cyclic group, a generator of it would have been of finite order
and hence a torsion element). So, Ker∂n is an infinite cyclic group and hence isomorphic to
Z. i.e., Hn (sk(σ n+1 )) ∼
= Z, as desired.
Chapter 3
Simplicial Approximations
In this chapter we observe that a simplicial map between two simplicial complexes induces a
continuous function between their corresponding geometric carriers |K| and |L|. We shall ex-
plore whether or not any continuous map can be approximated by a simplicial map. For that
matter, we need to make the term ‘approximation’ more precise and search for a satisfactory
answer to this question.
Given a simplicial map f : K → L, we wish to define a map fb : |K| → |L| and establish
that such fb is continuous.
25
A. Deb Ray, Pure Math, C.U. 26
Theorem 3.1.2. Let K and L be two simplicial complexes. Each simplicial map f : K → L
induces a continuous function fb : |K| → |L|.
Proof. Let f : K → L be a simplicial map and take any x ∈ |K|. Without loss of generality,
m
P
we choose σ = ⟨a0 , a1 , . . . , am ⟩ such that x ∈ int(σ). Then x = λi ai where λi > 0, for all
i=0
m
P
i = 0, 1, . . . , m and define fb(x) = λi f (ai ). In view of an earlier observation (see the 2nd
i=0
part of Observation 2.4.1), the unique representation of x ensures that the function fb is well
defined. we need to show that fb is continuous.
Since each simplex σ is a closed set, it is enough to show that f |σ is continuous. Because,
whenever x ∈ σ ∩ τ , fb|σ (x) = fb|τ (x) (using Observation 2.4.1 2nd part, again) and hence,
by pasting lemma, fb is continuous on |K|.
Consider the barycentric representation (λ1 , λ2 , . . . , λm ) of x.
Case 1 : Assume that for all i = 0, . . . , m, f (ai )’s are all distinct.
Consider the following diagram, for each j = 0, . . . , m.
fb
(λ0 , λ1 , . . . , λm ) / (λ0 , λ1 , . . . , λm )
πj
πj ◦fb
)
λj
/f(λ0 , λ1 , . . . , λk
b
(λ0 , λ1 , . . . , λm ) + λs , . . . λm )
πk
πk ◦fb *
λk + λs
So πk ◦fb = πk +πs , i.e., the sum of two continuous maps and hence, continuous. Consequently,
fb is continuous.
It is now natural to ask whether a continuous function g : |K| → |L| is achievable from
a simplicial map K → L. Of course it is too ambitious to hope for getting a simplicial map
f : K → L such that the induced continuous function fb is exactly equal to the given function
g. But, one may relax the condition of equality by replacing it with homotopy and frame a
little more convincing question as follows :
“For any continuous function g : |K| → |L| can we get a simplicial map f : K → L, such
that fb is homotopic to g? ”
The answer to this question is still in negative. If such f exists then it is called a simplicial
A. Deb Ray, Pure Math, C.U. 27
approximation of g. In the section that follows next, we explore, under which condition(s)
such ‘simplicial approximations’ exist.
Definition 3.2.1. Let K and L be two simplicial complexes and g : |K| → |L| a continuous
function. K is said to be star-related to L with respect to g, if for each vertex v of K there
exists a vertex w of L such that g(ost(v)) ⊆ ost(w).
k
T
Therefore, ost(vi ) ⊇ int(σ) ̸= ∅.
i=0
k
T
Conversely, let {v0 , v1 , . . . , vk } be such that ost(vi ) ̸= ∅. Then there exists some x ∈
i=0
ost(vi ), for each i = 0, 1, . . . , k. So x ∈ int(σi ), for some simplex σi (i = 0, 1, . . . , k) of K
with vi as a vertex. Since {int(σ) : σ ∈ K} forms a partition of K, σ0 = σ1 = . . . = σk = τ
(say). Since vi is a vertex of σi for each i, it follows that v0 , v1 , . . . , vk are vertices of τ . Then
either ⟨v0 , v1 , . . . , vk ⟩ is τ or a face of τ . In either case, ⟨v0 , v1 , . . . , vk ⟩ ∈ K.
Theorem 3.2.1. Suppose K and L are simplicial complexes and g : |K| → |L|, a continuous
function. If K is star-related to L with respect to g, then there is a simplicial map f : K → L
such that the induced continuous function fb is homotopic to g. (in otherwords, g has a
simplicial approximation)
Proof. K and L are star-related with respect to g implies that for each v in K there exists
w ∈ L such that g(ost(v)) ⊆ ost(w). Define f : vert(K) → vert(L) as follows:
f (v) = w whenever g(ost(v)) ⊆ ost(w).
We claim that f is a simplicial map. For that, we take any σ = ⟨v0 , . . . , vk ⟩ from
K.
Tk Tk
By Lemma 3.2.1, ost(vi ) ̸= ∅. For each i, ost(f (vi )) ⊇ g(ost(vi )) ⊇ g ost(vi ) ̸= ∅.
i=0 i=0
Applying the lemma again, ⟨f (v0 ), f (v1 ), . . . , f (vk )⟩ is a simplex in L. Hence, f is a simplicial
A. Deb Ray, Pure Math, C.U. 28
map.
Let x ∈ |K|. Then there exists σ = ⟨v0 , v1 , . . . , vn ⟩ ∈ K such that x ∈ int(σ). Then
n
P n
P n
P
x = λi vi , where each λi > 0 and λi = 1. Then fb(x) = λi f (vi ). If each f (vi ) is
i=1 i=1 i=1
distinct, then the coefficients of fb(x) (i.e., λi ’s) are all > 0. If for some i ̸= j f (vi ) = f (vj ),
then the i-th coefficient in fb(x) (i.e., λi + λj ) is > 0. So, in either case, barycentric co-
ordinates are > 0. Hence, fb(x) ∈ int (⟨f (v0 ), . . . , f (vn )⟩) (cancelling possible repetitions of
vertices). So, for each i = 0, 1, . . . , n, fb(x) ∈ ost (f (vi )). On the other hand,
Therefore, both g(x) and fb(x) are in ost(f (vi )), for each i = 0, 1, 2, . . . n. i.e., g(x) and fb(x)
n
T
are in ost(f (vi )), hence by Lemma 3.2.1, there exists some simplex τ with vertices f (vi ),
i=0
i = 0, 1, . . . , n, such that g(x), fb(x) ∈ τ . A simplex being a convex set, the line segment
tg(x)+(1−t)fb(x) ∈ τ , for all t ∈ [0, 1]. Therefore, we define a homotopy H : |K|×[0, 1] → |L|
by
H(x, t) = tg(x) + (1 − t)fb(x)
H
It is clear that g ∼ fb.
How far can we extend this result? For obtaining a simplicial approximation of any
continuous function g : |K| → |L|, we need to build some basics.
Definition 3.2.2. Let σ = ⟨v0 , v1 , . . . , vk ⟩ be a simplex. The barycentre of σ, denoted by σ̇
is the point on σ whose barycentric co-ordinates are all equal. i.e.,
k
X 1
σ̇ = vi
i=0
k+1
barycentre of a 2-simplex is the centroid of the triangle formed by the 2-simplex, etc.
We construct a new simplicial complex K (1) from K obeying the rules mentioned below:
1. Take the barycentres of all the simplexes of K as the set of vertices (i.e., 0-simplexes);
2. If σ0 , σ1 , . . . , σn ∈ K, then take ⟨σ˙0 , σ˙1 , . . . , σ˙k ⟩ ∈ K (1) if and only if there is some
permutation t of {0, 1, . . . , n} such that σt(i) is a proper face of σt(i+1) in K.
K (1) is known as the first barycentric subdivision of K.
(1)
The first barycentric subdivision K (1) of K (1) is called the 2nd barycentric subdivision
of K. Hence, by induction, the n-th barycentric subdivision of K, denoted by K (n) is given
(1)
by K (n−1) , for all n ∈ N, where by K 0 we mean the original simplicial complex K.
A. Deb Ray, Pure Math, C.U. 29
The following lemma shows that for calculating the mesh of a complex it is enough to
calculate the lengths of the 1-simplexes of K and take the maximum.
k
P k
P
Proof. Let σ = ⟨v0 , v1 , . . . , vk ⟩ and x, y ∈ σ. Then x = λi vi and y = µi vi , where
i=0 i=0
k
P k
P
λi , µi ≥ 0, λi = µi = 1.
i=0 i=0
Xk k
X k
X
||x − y|| = ||1.x − y|| = ||( µi )x − µi vi || = || µi (x − vi )||
i=0 i=0 i=0
Therefore,
k
X k
X
||x − y|| ≤ µi ||x − vi || ≤ µi max ||x − vi || = max ||x − vi ||.
i i
i=0 i=0
for some vertices vr , vs of σ. Then the diameter of σ = sup ||x − y|| ≤ ||vr − vs ||, for some
x,y∈σ
vertices vr , vs of σ. Clearly, ||vr − vs || ≤ sup ||x − y||. So, diameter(σ) = ||vr − vs ||, as
x,y∈σ
desired.
Theorem 3.2.2. If K is a simplicial complex with dim(K) > 0 then lim mesh K (n) = 0.
n→∞
Proof. Suppose
dim(K) = m. By definition of mesh and applying the above Lemma,
(1)
mesh K = max ||σ̇ − τ̇ ||.
⟨σ̇,τ̇ ⟩
Since ⟨σ̇, τ̇ ⟩ ∈ K (1) , σ is a proper face of τ .
p
1
P
Let τ = ⟨v0 , v1 , . . . , vp ⟩, where p ≤ m. Then τ̇ = p+1
vi is a point on τ . Also, σ̇ is a point
i=0
on σ and hence a point on τ . So, proceeding as the proof of the lemma above, there exists
A. Deb Ray, Pure Math, C.U. 30
Proof. K and L are finite simplicial complexes and therefore, |K| and |L| are compact subsets
of Euclidean spaces. {ost(w) : w ∈ vert(L)} is an open cover of |L| and so, applying
Lebesgue covering theorem, it has a Lebesgue number, say η1 > 0. Using the definition of
Lebesgue number, for each y ∈ |L|, there is some ost(w) such that Bη (y) ⊆ ost(w), where
Bη (y) = {z : ||z − y|| < η} and η = η21 .
Since g is continuous on compact set it is uniformly continuous and hence, for η there exists
some δ > 0 such that whenever ||s − t|| < δ, ||g(s) − g(t)|| < η.
From Lemma 3.2.2 we get that lim meshK (n) = 0. So, for δ > 0, there exists some n0 ∈ N,
n→∞
such that for all n ≥ n0 , mesh K (n) < 2δ .
Choose any k > n0 . Let v ∈ vertK (k) and x, z ∈ ost(v). There exist σ, τ ∈ K (k) with v as a
common vertex and such that x ∈ int(σ) and y ∈ int(τ ). Then ||x−v|| < 2δ and ||z −v|| < 2δ ;
so that ||x − z|| < δ. So, ||g(x) − g(z)|| < η. i.e., g(x), g(z) ∈ Bη (g(x)). Hence, there is some
w ∈ vert(L) such that g(x), g(z) ∈ Bη (g(x)) ⊆ ost(w). In otherwords, for each v ∈ vertK (k) ,
g(ost(v)) ⊆ ost(w), for some w ∈ vert(L). i.e., K (k) and L are star related with respect to
g. Using Theorem 3.2.1, there exists a simplicial map f : K (k) → L such that the induced
continuous function fb is homotopic to g.
Chapter 4
In Chapter 2, we have seen how simplicial homology groups can be computed for a (finite)
simplicial complex K. It has also been observed that homology groups do not depend on
the orientation of the complex and therefore, one may use any suitable orientation of K for
the purpose of computing its homology groups.
A topological space X is said to be triangulable if there exists some simplicial complex K
such that X is homeomorphic to |K|. It is quite clear that one topological space may possibly
have more than one triangulations. In such case, unless we could prove that Hq (K) ∼
= Hq (L),
for each q, where K and L are two triangulations of a space X, we won’t be able to define
the homology group of X by taking Hq (X) = Hq (K). In fact, we assume without proof the
statement of the following result for the time being,
Theorem 4.0.1. If a topological space X has two triangulations K and L then Hq (K) ∼ =
Hq (L), for each q.
(Note that we use the same notation ∼= to mean isomorphism of groups as well as homeo-
morphism of topological spaces. One may figure out from the context the associated meaning
of ∼
=)
In view of this theorem, we define homology groups of a topological space as follows:
Definition 4.0.1. If X is a topological space and K is any triangulation of X then for each
integer q,
Hq (X) = Hq (K).
So, for any topological space X if we can find a simplicial complex K whose homology
groups are known and |K| is homeomorphic to X then we are done. In what follows we find
triangulations of some well known spaces.
We discuss here triangulations of some known compact surfaces (i.e., compact topological
2-manifolds). A triangulated surface can be regarded as having been constructed by gluing
together the various triangles in a specific way. Triangulating a surface is thus similar to
fitting the jigsaw puzzle pieces together to get a nice picture! Two different triangles cannot
have the same set of vertices and therefore, we can specify completely a triangulation of a
surface by numbering the vertices, and then listing which triples of vertices form triangles.
Such a list of triangles completely determines the surface together with how they are joined
and hence give a triangulation up to homeomorphism.
We can show (check) that the following is not a triangulation of a mobius band:
It is not hard to realize that one may have several triangulations for a particular topological
space. So, given a topological space X, one may get two simplicial complexes K1 and K2
such that X ∼ = |K1 | ∼
= |K2 |.
A. Deb Ray, Pure Math, C.U. 34
and (
Z, if q = 0
Hq (cl(σ n )) =
0, otherwise
Hence, (
Z, if q = 0 or n
Hq (S n ) =
0, otherwise.
and (
Z, if q = 0
Hq (Dn ) =
0, otherwise
BY
A. DEB RAY
Singular Homology
In the previous chapters we have seen how simplicial homology groups can be constructed
for ‘triangulable’ spaces. Though simplicial homology groups are easy to calculate, simplicial
theory has an inherent limitation. Let us look back and scrutinize the results that we have
obtained so far, in simplicial homology.
Every simplicial map K → L, induces continuous function |K| → |L| very naturally.
It has been proved that, given a continuous function f : X → Y , where X ∼ = |K| and
Y ∼= |L|, we can construct (in view of ‘simplicial approximation theorem’) a typical
simplicial complex K (n) (known as nth barycentric subdivision of K) and a simplicial
map g : K (n) → L such that the induced continuous function |g| : |K (n) | = |K| ∼
=X→
∼
Y = |L| is homotopic to f .
But the question remains : Does the continuous function f : X → Y induce any homomor-
phism Hq (f ) : Hq (X) → Hq (Y )?
Or, one may put the question more precisely : For a continuous function f , can we define
Hq (f ) by taking the homomorphisms Hq (g), induced by the simplicial approximation g of
f ? This, certainly, is non-trivial and therefore, not coming very naturally in our mind.
So, it is quite expected from the mathematicians to rethink the process of building homology
theory, by taking into consideration the topological space X itself, instead of its triangula-
tion. In doing so, we can get rid of the limitation of taking ‘triangulable spaces’ and may
construct homology groups of any topological space. However, computation may become
more complicated in this new setting.
In this chapter, we shall learn how the idea of simplicial homology has been gradually
extended to achieve singular homology theory.
1
A. Deb Ray, Pure Math, C.U. 2
If K is a (finite) simplicial complex, then any n-simplex can be viewed as a copy of ∆n , and
therefore, we may re-define an n-simplex by saying that it is a continuous injective function
∆n → |K|. In that case, K becomes a list of continuous injections from ∆n to |K|. We
generalize this idea to obtain the definition of a singular n-simplex :
Definition 5.1.2. Let X be a topological space. Any continuous function ∆n → X is called
a singular n-simplex. The collection of all singular n-simplexes on X is denoted by Sn (X).
From Definition 5.1.2 it is clear that there is no need for the topological space X to be
triangulable. Also we note that a singular n-simplex is not in general injective and so, such
a map need not preserve the topology of ∆n . For instance, the following are possible images
of singular1-simplexes, though they are not simplicial 1-simplexes.
We have seen earlier that, intuitively, geometric carrier of a simplicial complex K is achieved
via gluing the simplexes along their boundaries suitably. To extend this idea of formation
of |K| in singular case, we need to define boundaries of singular n-simplexes suitably.
We define for each i = 0, 1, . . . , n, di : Rn → Rn+1 by
di (x0 , . . . , xn−1 ) = (x0 , . . . , xi−1 , 0, . . . , xn−1 ).
We observe that di (∆n−1 ) ⊆ ∆n , and hence obtain di |∆n−1 : ∆n−1 → ∆n . In what follows,
we shall refer to this restriction map di |∆n−1 simply by di . Further, to avoid notational
complexity, we use the same notation di to refer to all the maps ∆n → ∆n+1 that inserts 0
in its i-th co-ordinate, for all n ≥ 0 and understand its domain from the context. Note that,
for di ◦ dj : ∆n−1 → ∆n+1 ,
(
(x0 , . . . , xj−1 , 0, xj , . . . , xi−2 , 0, xi−1 , . . . , xn−1 ) if i > j
di ◦ dj (x0 , x1 , . . . , xn−1 ) = (5.1)
(x0 , . . . , xi−1 , 0, xi , . . . , xj−1 , 0, xj , . . . , xn−1 ) if i ≤ j
Since a (n − 1)-face of ∆n consists of points all of which have 0 in exactly one specific co-
ordinate, we see that di (∆n−1 ) is a (n − 1)-face of ∆n and so, the boundary of ∆n , denoted
by ∂∆n is given by d0 (∆n−1 ) ∪ d1 (∆n−1 ) ∪ . . . ∪ dn (∆n−1 ). i.e., ∂∆n consists of all points in
∆n with at least one co-ordinate 0. So, for a singular n-simplex f : ∆n → X, we get (n+1)
different (n − 1)-simplexes f ◦ di : ∆n−1 → X. These observations lead to the definition of
singular chain groups and boundary maps.
A. Deb Ray, Pure Math, C.U. 3
Definition 5.1.3. Let Cn (X) be the free abelian group on the set Sn (X) of singular n-
simplexes (n ≥ 0). i.e., Cn (X) is the collection of all Z-linear combinations of singular
n-simplexes. The elements of Cn (X) are known as singular n-chains. If n < 0 then define
Cn (X) as the trivial group, denoted by 0.
The rest of the construction of homology groups is similar to that of simplicial homology
groups.
Proof. It is enough to prove (δn ◦ δn+1 )(f ) = 0, for any singular (n + 1)-simplex f . From
(5.1) it follows that
(
dj ◦ di−1 , if i > j
di ◦ dj = (5.2)
dj+1 ◦ di , if i ≤ j
Proof. If f ∈ Im(δn+1 ), there exists g such that f = δn+1 (g) which implies that δn (f ) =
(δn ◦ δn+1 )(g) = 0. Hence f ∈ Ker(δn ).
We define the terms chain complex, cycles, boundaries and finally, singular homology groups
exactly the same way as those defined for simplicial homology.
A. Deb Ray, Pure Math, C.U. 4
Definition 5.1.5. Let X be a topological space. Consider the doubly infinite sequence
{Cn (X), δn } of groups and boundary maps, called a singular chain complex,
δn+1 n δ
. . . Cn+1 (X) → Cn (X) → Cn−1 (X) → . . .
where Cn (X) is the free abelian group on Sn (X), for n ≥ 0 and trivial group otherwise.
Define for each n ≥ 0, Bn (X) = Im(δn+1 ) and Zn (X) = Ker(δn )
3. (
Zn (X)/Bn (X), if n ≥ 0
Hn (X) =
0, otherwise
are called the singular homology groups.
Example 5.1.1. Suppose X is a singleton, i.e., X = {x}. Then for each n ≥ 0, Sn (X) =
{f : ∆n → X : f (t) = x, ∀t ∈ ∆n } = {Cx }, where Cx : ∆n → X is the constant map t 7→ x.
By Definition of Cn (X) (n ≥ 0), it is the free abelian group generated by Cx , and hence it
is ∼
= Z and trivial for all n ≤ 0.
We observe that Cx ◦ d0 = Cx ◦ d1 = . . . = Cx ◦ dn = constant function. Consequently, for
n ≥ 1, (
n
X Cx ◦ d0 , n is even
δn (Cx ) = (−1)i (Cx ◦ di ) =
i=0
0, n is odd
So, for each n ≥ 1, (
0, if n is even
Zn (X) = Ker(δn ) =
Z, if n is odd
and Z0 (X) = C0 (X). Also, for each n ≥ 0
(
0, if n is even
Bn (X) = Im(δn+1 ) =
{Cx ◦ d0 : Cx ∈ Cn+1 (X)} ∼
= Z, if n is odd
Hence, for all n > 0, Hn (X) = Zn (X)/Bn (X) = 0. But H0 (X) = C0 (X)/B0 (X) = Z/0 =
∼
= Z. Consequently, (
Z, if n = 0
Hn (X) =
0, otherwise
Example 6.2.1 can be generalized to any discrete space consisting of finitely many (say, r
many) points, as seen next.
t 7→ xi . Hence Cn (X) is a free abelian group of rank r, i.e., Cn (X) = ⊕ri=1 Zfxi ∼
= ⊕ri=1 Z. As
in Example 6.2.1, f ◦ d0 = f ◦ d1 = . . . = f ◦ dn , for each f ∈ Sn (X). So, for n ≥ 1,
n
(
X f ◦ d0 , n is even
δn (f ) = (−1)i (f ◦ di ) =
i=0
0, n is odd
Therefore, proceeding exactly as before, Ker(δn ) = Im(δn+1 ), for each n ≥ 1, giving all
Hn (X) = 0 except H0 (X), which in this case is a free abelian group of rank r. i.e.,
(
Z ⊕ Z ⊕ . . . ⊕ Z (r copies), if n = 0
Hn (X) =
0, otherwise
In Example 5.1.2, we observe that the underlying space X is a disconnected space with r
path components and the 0-th homology group is the free abelian group whose rank is ex-
actly the number of path components of X. The earlier example is just a special case of this
one, because, a singleton set has only one path component. These two examples inspire us to
expect that the 0-th singular homology group might count the number of path components
of the topological space X. Following theorem asserts that our guess is absolutely right. In
fact, the 0-th singular homology groups provide information about path components of X.
ϵ(s) = 1
P P
for s ∈ S0 (X) and extend it by linearity on C0 (X). Then ϵ(c) = ni , where c = ni s i ,
si ∈ S0 (X). Clearly, ϵ is a homomorphism. For any σ ∈ S1 (X), ϵ(δ1 (σ)) = ϵ(σ ◦d0 −σ ◦d1 ) =
ϵ(σ ◦ d0 ) − ϵ(σ ◦ d1 ) = 1 − 1 = 0.
Define augmentation map ϵ∗ : H0 (X) → Z by
ϵ∗ (z + B0 (X)) = ϵ(z)
Then ϵ∗ is well defined, because, z1 +B0 (X) = z2 +B0 (X) ⇒ z1 −z2 ∈ B0 (X) ⇒ ϵ(z1 −z2 ) = 0.
Also ϵ∗ is a homomorphism. In fact, it is a surjective homomorphism.
A. Deb Ray, Pure Math, C.U. 6
H0 (X) ∼
= Z.
X X X
z − δ1 ( nx λx ) = nx Cx − nx δ1 (λx )
X
= nx (Cx − δ1 (λx ))
m
P
z ∈ Zn (X) ⊆ Cn (X), z = zj , where zj ∈ Cn (Yj ), for each j = 1, 2, . . . , m.
j=1
Xm
δn (z) = 0 ⇒ δn ( zj ) = 0
j=1
m
X
⇒ δn (zj ) = 0
j=1
⇒ δn (zj ) = 0
⇒ zj ∈ Zn (Yj )
m
P
Now, b ∈ Bn (X) implies that b = δn+1 (a) for some a ∈ Cn+1 (X). Then a = aj , where
j=1
m
P m
P
aj ∈ Cn+1 (Yj ). So, δn+1 (a) = δn+1 ( aj ) = δn+1 (aj ).
j=1 j=1
i.e., b is uniquely expressible as linear sum of elements of Bn (Yj ). Hence, z + Bn (X) =
Pm
(zj + Bn (Yj )) uniquely. So,
j=1
Hn (X) = ⊕m
j=1 Hn (Yj ).
H0 (X) ∼
= ⊕m
i=1 Z.
Proof. Since Sn (X) generates the free abelian group Cn (X) and Cn (f ), δn and ∂n are ho-
momorphisms, it is enough to prove the result on the set of generators. Thus, we take any
s ∈ Sn (X) and show that (Cn−1 (f ) ◦ δn )(s) = (∂n ◦ Cn (f ))(s) holds.
for each n ≥ 0.
A. Deb Ray, Pure Math, C.U. 9
Proof. By (i) of Corollary 5.2.1 it is clear that image of any point of Hn (X) under the map
Hn (f ) is inside Hn (Y ). That Hn (f ) is well-defined, follows from Corolloary 5.2.1(ii), because,
The following theorem shows that the induced homomorphisms satisfy functorial prop-
erties:
Theorem 5.2.4. For each n ≥ 0,
(1) If I : X → X is the identity map then Hn (I) : Hn (X) → Hn (X) is the identity
homomorphism.
(2) If f : X → Y and g : Y → Z are continuous functions then Hn (g ◦ f ) = Hn (g) ◦ Hn (f ).
Proof. (1) By definition, Hn (I) : Hn (X) → Hn (X) is such that Hn (I)(z + Bn (X)) =
Cn (I)(z) + Bn (X) where Cn (I)(s) = I ◦ s for each s ∈ Sn (X). But (I ◦ s)(t) = s(t), for all
t ∈ ∆n and so, I ◦ s = s, for each s ∈ Sn (X). As Sn (X) generates Cn (X), Cn (I)(ϕ) = ϕ,
for all ϕ ∈ Cn (X). i.e., Hn (I)(z + Bn (X)) = z + Bn (X), proving Hn (I) is the identity
homomorphism.
(2) For continuous functions f : X → Y and g : Y → Z,
and
Hn (f ) ◦ Hn (g) = Hn (f ◦ g) = Hn (IY ) = IHn (Y )
This shows that the homomorphism Hn (f ) (for each n) is invertible, i.e., Hn (f ) : Hn (X) →
Hn (Y ) is an isomorphism.
If we take a homotopy equivalence f : X → Y instead of a homeomorphism then is Hn (f ) :
Hn (X) → Hn (Y ) an isomorphism? To answer this question, we need to know what happens
to the induced homomorphisms corresponding to homotopic maps.
Therefore, δ1 ( 1j=0 (−1)j (dj × id) ◦ α0 )(1) = (d0 × id)(G0 (1)) − (d1 × id)(G0 (1)) = (0, 1, 1) −
P
(0, 1, 0) − (1, 0, 1) + (1, 0, 0) δ1 (G1 (1)) = G1 (d0 (1)) − G1 (d1 (1)) = (0, 1, 1) − (0, 1, 0) −
and P
(1, 0, 1) + (1, 0, 0). So, δ1 G1 − 1j=0 (−1)j (dj × id) ◦ α0 = 0. i.e., G1 − 1j=0 (−1)j (dj ×
P
id) ◦ α0 ∈ Ker(δ1 ). Since ∆1 × I is a convex set, Hk (∆1 × I) = 0 for all k > 0. So,
Ker(δ1 ) = Im(δ2 ) and consequently, there exists some α1 : ∆2 → ∆1 × I such that
1
X
δ2 (α1 ) + (−1)j (dj × id) ◦ α0 = G1 .
j=0
Suppose that we have already obtained αn−1 : ∆n → ∆n−1 × I satisfying the condition
n−1
X
δn (αn−1 ) + (−1)j (dj × id) ◦ αn−2 = Gn−1 .
j=0
Then
n
! n n
!
X X X
δn (−1)i (di × id) ◦ αn−1 = (−1)k (−1)i (di × id) ◦ αn−1 ◦ dk
i=0 k=0 i=0
n
X
= (−1)i (di × id) ◦ δn (αn−1 )
i=0
n n−1
!
X X
= (−1)i (di × id) ◦ Gn−1 − (−1)j (dj × id) ◦ αn−2
i=0 j=0
Xn
= (−1)i (di × id) ◦ Gn−1
i=0
n X
X n−1
− (−1)i+j (di × id) ◦ (dj × id) ◦ αn−2
i=0 j=0
n
X
= (−1)i (di × id) ◦ Gn−1
i=0
n X
X n−1
− (−1)i+j ((di ◦ dj ) × id) ◦ αn−2
i=0 j=0
n
X
= (−1)i (di × id) ◦ Gn−1 ,
i=0
Lemma 5.3.2. If f , g are two continuous maps from X into Y which are homotopic, then
there exists a family {Φn : Cn (X) → Cn+1 (Y )}n≥0 of homomorphisms such that for n > 0
and
∂1 Φ0 = C0 (g) − C0 (f ).
i.e., In diagrams,
δn+1 δn
Cn+1 (X) / Cn (X) / Cn−1 (X) and C0 (X)
0 / 0
y Φn y Φn−1 z Φ0 } 0
Cn+1 (Y ) ∂ / Cn (Y ) / Cn−1 (Y ) C1 (Y ) / C0 (Y ) / 0
n+1 ∂n ∂1 0
and for n = 0, δ1 (α0 ) = G0 . So, for every n ≥ 0 we have the following composite map in
Cn+1 (Y ):
αn σ×id F
∆n+1 −→ ∆n × I −→ X × I −→ Y
For each n ≥ 0, define Φn : Cn (X) → Cn+1 (Y ) by Φn (σ) = F ◦ (σ × id) ◦ αn , for each
σ ∈ Sn (X) and extending it by linearity on the whole of Cn (X). Each Φn is indeed a
homomorphism. We claim that such a sequence satisfies the desired condition. If n > 0, for
any σ ∈ Sn (X),
Similarly, if n = 0, then for any σ ∈ S0 (X), (∂1 Φ0 )(σ) = F ◦ (σ × id) ◦ G0 . For any z ∈ ∆n
(n ≥ 0), F (σ(z), 1) − F (σ(z), 0) = g(σ(z)) − f (σ(z)) so that for n > 0
and for n = 0
i.e., ∂1 Φ0 (σ) = (C0 (g) − C0 (f ))(σ) and for n > 0, ∂n+1 Φn + Φn−1 δn = Cn (g) − Cn (f ), as
desired.
Definition 5.3.1. Let {Cn , δn }n≥0 and {Dn , ∂n }n≥0 be two chain complexes and ϕn , ψn :
Cn → Dn (n ≥ 0) be two chain maps. A family of homomorphisms {Φn : Cn → Dn+1 }n≥0 is
called a chain homotopy between {ϕn } and {ψn } if
∂n+1 Φn + Φn−1 δn = ψn − ϕn
In view of this definition of chain homotopy, we may re-state Lemma 5.3.2 by saying that
two homotopic chain maps induce a chain homotopy between the given chain complexes.
Now we see how this existence of chain homotopy helps to prove our main result:
Proof. Since f and g are homotopic, there exists a chain homotopy {Φn : Cn (X) → Cn+1 (Y )}n≥0 .
If c ∈ C0 (X),
Finally, using Theorem 5.3.1 we show next that Hn (f ) is an isomorphism (for all n ≥ 0),
if f is a homotopy equivalence. In other words, topological spaces which are of the same
homotopy type have the same homology groups in each dimension.
and
Hn (f ) ◦ Hn (g) = Hn (f ◦ g) = Hn (IY ) = IHn (Y )
So, Hn (f ) and Hn (g) are inverse of each other and hence, isomorphisms.
Chapter 6
In this chapter we first prove Mayer-Vietoris Theorem and then compute homology groups
using Mayer-Vietoris sequence. This theorem guarantees the existence of a long exact se-
quence of homology groups for a topological space X which is expressible as union of two
of its open subspaces. Such a long exact sequence is known as Mayer-Vietoris sequence (in
short, M-V sequence).
fn : Hn (U ) ⊕ Hn (V ) → Hn (X) by
gn : Hn (U ∩ V ) → Hn (U ) ⊕ Hn (V ) by
In what follows, for ease of notation, we shall denote these maps by f and g instead of fn
and gn and understand the domain from the context.
We state a homological analogue of Van Kampen’s theorem, without proof, that we shall
use frequently:
Lemma 6.1.1. (Chain Splitting) If s ∈ Cn (X) and X = U ∪ V where U and V are open
in X, then there are sU ∈ Cn (U ) and sV ∈ Cn (V ) such that s is homologous to sU + sV .
i.e., There is some chain t ∈ Cn+1 (X) such that s − sU − sV = δn+1 (t).
It is easy to observe that Hn (i) ◦ Hn (k) = Hn (j) ◦ Hn (l). In other words, the following
diagram commutes:
Hn (k)
Hn (U ∩ V ) / Hn (U )
Hn (j) Hn (i)
Hn (V ) / Hn (X)
Hn (l)
Hence,
f (g(x)) = f ((Hn (k)(x), Hn (l)(x))) = Hn (i)(Hn (k)(x)) − Hn (j)(Hn (l)(x)) = 0
Therefore, Im(g) ⊆ Ker(f ).
Let c ∈ Cn (X) be such that c+Bn (X) ∈ Hn (X). Then c is a cycle and so, δn (c) = 0. By chain
splitting, there exists cU ∈ Cn (U ) and cV ∈ Cn (V ) such that c − cU − cV = δn+1 (t) for some
t ∈ Cn+1 (X). Therefore, δn (c − cU − cV ) = δn (δn+1 (t)) = 0. i.e., 0 = δn (c) − δn (cU ) − δn (cV )
which gives
δn (cU ) = −δn (cV )
The left side of this expression shows that it is a chain in U and the right side of the expression
shows that it is a chain in V . Hence, both are chains in U ∩ V . Also, δn (cU ) ∈ Cn−1 (U ∩ V )
with δn−1 (δn (cU )) = 0 implies that δn (cU ) ∈ Zn−1 (U ∩ V ). Hence, for each c + Bn (X) ∈
Hn (X), we find an element δn (cU ) + Bn−1 (U ∩ V ) ∈ Hn−1 (U ∩ V ). It is not hard to show
that such an element depends on c and not on the choice of cU .
Proof. c is homologous to cU +cV implies that there exists p ∈ Cn+1 (X) such that c−cU −cV =
δn+1 (p). Similarly, c is homologous to bU +bV implies that there exists q ∈ Cn+1 (X) such that
c − bU − bV = δn+1 (q). Therefore, subtracting one from the other we get, cU + cV − bU − bV =
δn+1 (q − p). By chain splitting on a = q − p, there exist aU ∈ Cn+1 (U ) and aV ∈ Cn+1 (V )
such that a − aU − aV = δn+2 (s), for some s ∈ Cn+2 (X). Therefore,
So, δn+1 (a) = δn+1 (aU )+δn+1 (aV ) which in turn gives cU +cV −bU −bV = δn+1 (aU )+δn+1 (aV ).
i.e.,
(cU − bU ) − δn+1 (aU ) = (bV − cV ) + δn+1 (aV )
The left side expression shows that it is a chain in U and the right side shows that it is
a chain in V . Hence both are chains in U ∩ V with boundary δn (cU − bU − δn+1 (aU ) =
δn (cU ) − δn (bU ). Consequently, δn (cU ) − δn (bU ) ∈ Bn−1 (U ∩ V ). Hence,
Lemma 6.1.3. For each n ≥ 1, there exists a map η : Hn (X) → Hn−1 (U ∩ V ), described by
Proof. To show that such assignment is well-defined. Let c + Bn (X) = d + Bn (X) ∈ Hn (X).
Then c − d ∈ Bn (X). So there is some t ∈ Cn+1 (X), such that c − d = δn+1 (t). By chain
splitting on c and d, there are cU , dU ∈ Cn (U ) and cV , dV ∈ Cn (V ) such that c − cU − cV =
A. Deb Ray, Pure Math, C.U. 19
δn+1 (y) and d − dU − dV = δn+1 (z) for some y, z ∈ Cn+1 (X). Then (c − d) − (cU − dU ) −
(cV − dV ) = δn+1 (y − z). Hence,
(cU − dU ) + (cV − dV ) = δn+1 (t − y − z)
Using chain splitting on (t − y − z), we get tU ∈ Cn+1 (U ) and tV ∈ Cn+1 (V ) such that
(t − y − z) ∼ tU + tV . In particular, δn+1 (t − y − z) = δn+1 (tU ) + δn+1 (tV ).
Let X be any contractible space. Then X is of the same homotopy type with a point space
{a}. So, by a result proved in the previous chapter, Hn (X) ∼
= Hn ({a}). Using the example
above, we then have
Example 6.2.2. For any contractible space X,
(
Z, if n = 0
Hn (X) =
0, otherwise
If p = (0, 1) and q = (0, −1), then U = S 1 \ {p} and V = S 1 \ {q} are open subsets of X such
that X = U ∪ V . Certainly, U and V both are homeomorphic to R and R is contractible.
So, Hn (U ) = 0 = Hn (V ) for each n ̸= 0. Again U ∩ V is`the disjoint union of two spaces P
and Q, each of which is homeomorphic to R. So, H0 (P Q) = H0 (P ) ⊕ H0 (Q) = Z ⊕ Z.
Applying M-V sequence,
If p = (0, 0, 1) and q = (0, 0, −1), then U = S 2 \ {p} and V = S 2 \ {q} are open subsets of
X such that X = U ∪ V . U and V both are homeomorphic to R2 and R2 is contractible.
So, Hn (U ) = 0 = Hn (V ) for each n ̸= 0. Again U ∩ V is homeomorphic to R × S 1 , which is
A. Deb Ray, Pure Math, C.U. 22
. . . → 0 → Hn (X) → 0 → . . . → 0 → H2 (X) → Z → 0 ⊕ 0
→ H1 (X) → Z → Z ⊕ Z → Z → 0
Consider an open set U obtained by deleting any point from the left circle (not lying on the
right circle) and an open set V obtained by deleting any point from the right circle (not
lying on the left circle).
Clearly, X = U ∪ V , with U and V both homotopy equivalent to S 1 and U ∩ V homotopy
equivalent to a point space, say, {p}. So, Hn (V ) = Hn (U ) = Hn (S 1 ) = Z, for n = 0, 1 and
trivial for any n > 1. Also, X being path connected, H0 (X) = Z. By M-V sequence,
f4 is surjective.
So, (Z ⊕ Z)/Ker(f4 ) ∼
= Im(f4 ) = Z and so, Ker(f4 ) = Z.
Im(f3 ) = Ker(f4 ) = Z.
Z/Ker(f3 ) ∼
= Im(f3 ) = Z and so, Ker(f3 ) = 0.
A. Deb Ray, Pure Math, C.U. 23
f1 f2 f3 f4 f5
0 → Z ⊕ Z → H1 (X) → Z → Z ⊕ Z → Z → 0.
Im(f2 ) = Ker(f3 ) = 0.
f2 = 0.
One may also note that in some cases, M-V Theorem alone may not give the precise
homology groups of a space. For example, if we consider a torus and apply M-V sequence
then we shall get a list of probable candidates for its homology groups, thereby get an
estimate, not exact ones. Using other known results and facts we may of course determine
its homology groups precisely.
6.3 Applications
Among various applications of singular homology groups, we cite only a couple of very
important applications of these groups to resolve purely topological problems. We observe
how elegantly such non-trivial fundamental questions can be answered via homology groups.
Hn (f −1 ) Hn (f )
Z −→ 0 −→ Z
It has been observed before that a special case of Brouwer’s Fixed Point Theorem can
be achieved using Fundamental groups. In what follows next, we see that for any n > 0 this
theorem can be completely resolved via homology groups.
Proof. Suppose, f : Dn → Dn has no fixed point. i.e., f (z) ̸= z, for all z ∈ Dn . Consider the
ray (1−t)f (z)+tz (t ≥ 0) that starts from f (z) and passes through z. This ray will meet the
boundary S n−1 at some point, say g(z). Then we get a continuous function g : Dn → S n−1
A. Deb Ray, Pure Math, C.U. 24
Case 1 : n > 1.
This sequence of continuous functions induces another sequence of homomorphisms between
the respective (n-1)-th homology groups. As Dn is convex, Hm (Dn ) = 0 for all m. So, we
have
Hn (i) Hn−1 (g)
Z∼= Hn−1 (S n−1 ) −→ Hn−1 (Dn ) ∼
= 0 −→ Hn−1 (S n−1 ) ∼=Z
By functorial properties of fundamental groups,
Hn−1 (g) ◦ Hn−1 (i) = Hn−1 (g ◦ i) = Hn−1 (IS n−1 ) = IHn−1 (S n−1 ) showing that identity map on
Z is factoring through the trivial group 0, which is impossible. Hence, such a g can not exist
and so, there is some z ∈ Dn such that f (z) = z.
Case 2 : n = 1.
In this case, D1 = [−1, 1] and S 0 = {−1, 1}. As D1 is connected, image of D1 under the
continuous function g is either {1} or {−1}, but not the whole of S 0 . So, g(z) = z for all
z ∈ S 0 can not be satisfied - a contradiction to the construction of g.
Before we conclude, we point out that there is a theorem due to Eilenberg and Steenrod
that says, for any triangulable space, the singular homology groups are same as the simplicial
homology groups.
Therefore, one may raise a question: If Singular Homology is more general then why
we study simplicial homology at all?
This is indeed a pertinent question. But for computing homology groups of a triangulable
space one may use its simplicial structure, avoiding the structural complications of singular
homology, whereas the efficiency of singular homology theory lies in the results involving
induced homomorphisms, M-V sequence, etc. This justifies the study of both.
Chapter 7
In this course of Algebraic Topology, we have studied fundamental groups of pointed topo-
logical spaces and have seen the construction of simplicial and singular homology groups of
topological spaces. A question certainly arises, whether any relation exists between funda-
mental group of a space and any of its homology group. We can not expect them to be
equal, because, homology groups are abelian but fundamental groups may not.
In this chapter, we shall learn that for path connected spaces, there always exists a ho-
momorphism from fundamental group onto its first homology group. This homomorphism
is known as Hurewicz homomorphism. Moreover, it can also be established that the
abelianization of the fundamental group of a path connected space is algebraically same
with the first homology group of the space.
7.1 Abelianization
Definition 7.1.1. Let G be a group and a, b ∈ G. The commutator of a and b is the element
aba−1 b−1 .
Let A = {aba−1 b−1 : a, b ∈ G} and let [G, G] be the subgroup of G generated by A.
[G, G] is called the derived or commutator subgroup of G.
Theorem 7.1.1. The derived subgroup [G, G] of a group G is a normal subgroup of G and
G/[G, G] is abelian.
Proof. By definition, [G, G] is a subgroup of G generated by the commutators aba−1 b−1 , for
all a, b ∈ G. It is enough to show that ∀a, b, g ∈ G, g(aba−1 b−1 )g −1 ∈ [G, G]. Now,
This homomorphism will induce another surjective homomorphism π(X, x0 )ab → H1 (X).
To show that such homomorphism is actually an isomorphism, known as Hurewicz
isomorphism.
Theorem 7.2.1. Hurewicz Theorem for n=1 The abelianization of the fundamental
group of a path connected space is isomorphic to its 1st Homology group.
1δ
. . . −→ C1 (X) −→ C0 (X) −→ 0
Theorem 7.2.2. If α, β : [0, 1] → X are two paths such that α(1) = β(0) then α+β−α ∗ β ∈
B1 (X).
Proof. Consider an isosceles triangle ∆ with vertices a0 , a1 , a2 and the midpoint of the base
[a0 , a1 ] (wherever we write [a, b], we mean the line segment joining a with b) of the triangle is
x0 . We join the vertex a2 with x0 . Suppose the line segment [a0 , a2 ] denotes the image of α
in X and the line segment [a2 , a1 ] denotes the image of β in X. Define a function F : ∆ → X
as follows:
(i) Choose any point (s, t) ∈ ∆ and draw a line l parallel to [a2 , x0 ] through (s, t).
(ii) If l intersects [a0 , a2 ], then assign F (s, t) = the value of α at that point of intersection.
(iii) If l intersects [a2 , a1 ], then assign F (s, t) = the value of β at that point of intersection.
Clearly, F is a continuous function, because α and β agree at a2 . Certainly, α = F ◦ <
a0 , a2 > and β = F ◦ < a2 , a1 >.
A. Deb Ray, Pure Math, C.U. 28
Consider the 2-simplex < a0 , a1 , a2 >: ∆2 → X. Then we claim that the 2-simplex F ◦ <
a0 , a1 , a2 >: ∆2 → X is such that δ2 (F ◦ < a0 , a1 , a2 >) = α + β − α ∗ β.
δ2 (F ◦ < a0 , a1 , a2 >)(s, t)
= (F ◦ < a0 , a1 , a2 > ◦d0 − F ◦ < a0 , a1 , a2 > ◦d1 + F ◦ < a0 , a1 , a2 > ◦d2 )(s, t)
= F ◦ < a1 , a2 > (s, t) − F ◦ < a0 , a2 > (s, t) + F ◦ < a0 , a1 > (s, t)
= −β(s, t) − α(s, t) + α ∗ β(s, t)
Hence, −β − α + α ∗ β ∈ B1 (X).
From Theorem 7.2.2, we get the following:
H H
H◦<a0 ,a1 ,a2 >
# H◦<a0 ,a2 ,a3 >
#
X X
Therefore, we consider τ = H◦ < a0 , a1 , a2 > +H◦ < a0 , a2 , a3 >∈ C2 (X). We claim
that δ2 (τ ) = α − β. Viewing H◦ < a0 , a3 > and H◦ < a1 , a2 > as singular 1-simplexes,
we get (H◦ < a0 , a3 >)(s, t) = x0 = (H◦ < a1 , a2 >)(s, t).
3. If α is a path then by Theorem 7.2.2, α + α−1 − α ∗ α−1 ∈ B1 (X). Also, α ∗ α−1 ∼ Cx0 ,
α ∗ α−1 − Cx0 ∈ B1 (X). Since B1 (X) is a group, α + α−1 − Cx0 ∈ B1 (X). Therefore,
α + α−1 ∈ B1 (X).
If [Cx0 ] is the identity element in π(X, x0 ), then Cx0 + B1 (X) = B1 (X) (= identity
element of H1 (X)).
ϕ([α]) = α + B1 (X).
ϕ is well defined : Follows from the fact that α ∼ β implies α − β ∈ B1 (X).
ϕ is a homomorphism : If α and β are two loops based at x0 , then by Theorem 7.2.2,
(α ∗ β) + B1 (X) = (α + B1 (X)) + (β + B1 (X)). Hence, ϕ([α] ◦ [β]) = ϕ([α]) + ϕ([β]). i.e., ϕ
is a homomorphism.
So, we can define a homomorphism ψ : π(X)ab → H1 (X) with the help of ϕ as follows:
Denote the commutator subgroup of π(X, x0 ) by G. Then any element of π(X)ab is denoted
by [α]/G, where α is a loop based at x0 . So, α is a member of H1 (X). Hence,
ψ([α]/G) = α.
We define a homomorphism η : H1 (X) → π(X)ab and show that such η is the inverse of ψ,
proving ψ to be an isomorphism.
For any singular 1-simplex σ, if σ(0, 1) = a and σ(1, 0) = b, then ωa ∗ σ ∗ ωb−1 is a loop
based at x0 .
Hence, define λ : S1 (X) → π(X)ab by sending σ to the element of π(X)ab represented by the
loop ωa ∗ σ ∗ ωb−1 and extend it to C1 (X) by linearity.
i.e., Define λ : S1 (X) → π(X)ab by λ(σ) = [ωa ∗ σ ∗ ωb−1 ]/G and extend it by linearity to
C1 (X).
Of course, it is not hard to see that λ is a homomorphism. If δ2 : C2 (X) → C1 (X) and
λ(δ2 (S2 (X))) = 0 then we can define a homomorphism η : H1 (X) → π(X)ab in a very
natural way. So, we proceed to show that λ(δ2 (S2 (X))) = 0.
Let ϵ : ∆2 → X be a singular 2-simplex. We shall show that λ(δ2 (ϵ)) = 0. Let ϵ(1, 0, 0) =
a, ϵ(0, 1, 0) = b, ϵ(0, 0, 1) = c. One can easily check that (ϵ ◦ d0 )(s, t) =< b, c > (s, t),
A. Deb Ray, Pure Math, C.U. 30
(ϵ◦d1 )(s, t) =< a, c > (s, t) (and therefore, (−ϵ◦d1 )(s, t) =< c, a > (s, t)) and (ϵ◦d2 )(s, t) =<
a, b > (s, t)
If di : ∆1 → ∆2 are the maps that insert 0 in i-th place (i = 0, 1, 2), we get
λ(δ2 (ϵ)) = λ(ϵ ◦ d0 − ϵ ◦ d1 + ϵ ◦ d2 )
= λ(ϵ ◦ d0 ) ◦ λ(−ϵ ◦ d1 ) ◦ λ(ϵ ◦ d2 )
= [ωb ∗ (ϵ ◦ d0 ) ∗ ωc−1 ]/G ◦ [ωc ∗ (−ϵ ◦ d1 ) ∗ ωa−1 ]/G ◦ [ωa ∗ (ϵ ◦ d2 ) ∗ ωb−1 ]/G
= [(ωb ∗ (ϵ ◦ d0 ) ∗ ωc−1 ) ∗ (ωc ∗ (−ϵ ◦ d1 ) ∗ ωa−1 ) ∗ (ωa ∗ (ϵ ◦ d2 ) ∗ ωb−1 )]/G
= [(ωb ∗ (ϵ ◦ d0 ∗ −ϵ ◦ d1 ∗ ϵ ◦ d2 ) ∗ ωb−1 )]/G
= [Cx0 ]/G
Because, ϵ ◦ d0 ∗ −ϵ ◦ d1 ∗ ϵ ◦ d2 is a loop based at b lying in the path connected space ϵ(∆1 ), it
is homotopic to the constant loop at b. Hence, ωb ∗ (ϵ ◦ d0 ∗ −ϵ ◦ d1 ∗ ϵ ◦ d2 ) ∗ ωb−1 is homotopic
to a constant loop at x0 . So, Im(δ2 ) ⊆ Ker(λ).
Define η : H1 (X) → π(X)ab by
η(σ + B1 (X)) = λ(σ).
Since σ is a 1-cycle, σ(1, 0) = σ(0, 1) = a (say), and therefore, λ(σ) = [ωa ∗ σ ∗ ωa−1 ]/G.
We now show that the map η is well defined.
Let σ1 = σ2 in H1 (X). Then
σ1 − σ2 ∈ B1 (X) = Im(δ2 ) ⊆ Ker(λ)
⇒ λ(σ1 − σ2 ) = 0
⇒ [ωa ∗ (σ1 − σ2 ) ∗ ωa−1 ] ∈ G
⇒ [ωa ∗ σ1 ∗ ωa−1 ]/G = [ωa ∗ σ2 ∗ ωa−1 ]/G
⇒ λ(σ1 ) = λ(σ2 )
It can now be checked that such η is the inverse of ψ:
(η ◦ ψ)([α]/G) = η(ϕ([α])) = η(α + B1 (X)) = λ(α) = [ωx0 ∗ α ∗ ωx−10
]/G = [α]/G, since α is a
−1
loop based at x0 , it is homotopic to ωx0 ∗ α ∗ ωx0 .
(ψ ◦η)(α+B1 (X)) = ψ(λ(α)) = ψ([ωa ∗α∗ωa−1 ]/G), where α is the loop corresponding to α ∈
Z1 (X), based at a ∈ X. Then (ψ ◦ η)(α + B1 (X)) = ϕ([ωa ∗ α ∗ ωa−1 ]) = ωa ∗ α ∗ ωa−1 + B1 (X).
We now show that ωa ∗ α ∗ ωa−1 − α ∈ B1 (X) :
Applying Theorem 7.2.2, ωa ∗ α − α − ωa ∈ B1 (X) and ωa ∗ α ∗ ωa−1 − ωa ∗ α − ωa−1 ∈ B1 (X).
Since B1 (X) is a group, ωa ∗ α ∗ ωa−1 − ωa−1 − α − ωa ∈ B1 (X). Also. ωa−1 + ωa ∈ B1 (X) and
therefore, ωa ∗ α ∗ ωa−1 − α − ωa ∈ B1 (X). Hence, we arrive our desired result.
This shows that ψ : π(X)a b → H1 (X) is an isomorphism. Hence, for a path connected
topological space, 1st Homology group and the abelianization of its fundamental group are
same.
Im(f3 ) = Z
Ker(f3 ) = Z
Im(f2 ) = Z
Ker(f2 ) = Z
Im(f1 ) = Z
Hn (T 2 ) = Z ⊕ Z, if n = 1
= Z, if n = 0, 2
= 0, otherwise