Toplogy Notes
Toplogy Notes
Toplogy Notes
M.Thamban Nair†
Visiting Professor
Department of Mathematics
BITS Pilani, KK Birla Goa Campus, Goa 403726
∗
Notes for the course MATH F311: Introduction to Topology, August-December 2023 at BITS
Pilani, Goa Campus.
†
Formerly Professor & Head of the Department of Mathematics at IIT Madras, Chennai.
.
Contents
1 Topology: Motivation, Definition and Examples 6
6 Compact spaces 66
6.1 Definition and implications . . . . . . . . . . . . . . . . . . . . . . . . . 66
6.2 Comapct subsets of R . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
6.3 Locally compact spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
6.4 One-point compactification . . . . . . . . . . . . . . . . . . . . . . . . . 77
7 Product topology 78
7.1 On product of finite number of topological spaces . . . . . . . . . . . . 78
7.2 On product of infinite number of topological spaces . . . . . . . . . . . 81
.
5
1 Topology: Motivation, Definition and Examples
Recall from the course on Real Analysis that if Ω is a metric space with metric d, then
the collection G of all open sets in Ω has the following properties:
1. ∅ ∈ G and Ω ∈ G;
We have seen that many notions introduced in Real Analysis can be expressed using
the notion of open sets and the above three properties of G, without explicit reference
to the metric. For example, look at the following:
6
Definition 1.1. (Topology) Let X be a set and T be a collection of subsets of X.
Then T is called a topology on X if it has the following properties:
1. ∅ ∈ T , X ∈T;
2. A, B ∈ R ⇒ A ∩ B ∈ T ;
[
3. Λ ̸= ∅ and Aα ∈ T ∀ α ∈ Λ ⇒ Aα ∈ T .
α∈Λ
Note that, if T is any topology on a set X, and if T0 and T1 are the indiscrete
topology and discrete topology, respectively, on X, then
T0 ⊆ T ⊆ T 1 .
7
Definition 1.5. Let T1 and T2 be topologies on a set X. Then, T1 is said to be weaker
than T2 if T1 ⊆ T2 , and in that case, T2 is said to be stronger than T1 .
The topology T1 is said to be strictly weaker than T2 if T1 is a proper subcollection
of T2 , and in that case T2 is said to be strictly stronger than T1 . ♢
We have already observed in the beginning of this section that the collection of all
open subsets of a metric space satisfies the requirements for a topology. In other words
the following example provides a class of examples.
Recall that, if Ω is a metric space with metric d, then a subset A of Ω is said to
be an open set if for every x ∈ A, there exists r > 0 such that Bd (x, r) ⊆ A, where
Bd (x, r) is the open ball centered at x and radius r, that is,
Bd (x, r) = {y ∈ Ω : d(x, y) < r}.
Thus, we have the following class of examples for a topology.
Example 1.6. Let Ω be a metric space with metric and Td be the collection of all
open subsets of Ω. Then Then Td is a topology on Ω. ♢
Definition 1.7. (Metric topology) Let Ω be a metric space with metric d. Then
the topology Td defined as in Example 1.6 is called the topology induced by the
metric d, and it is denoted by Td . ♢
Definition 1.8. (Metrizable space) A topological space (X, T ) is called metrizable
if there exists a metric d on X such that T = Td . ♢
Exercise 1.9. Let (Ω, d) be a metric space. Show that a subset A of Ω is open iff A
is a union of open balls in Ω. ♢
We may recall that, on Rk , the usual metric or the euclidean metric d is defined by
p
d(x, y) := (x1 − y1 )2 + · · · (xk − yk )2 ,
for x = (x1 , . . . , xk ) and y = (y1 , . . . , yk ) in Rk . Similarly, the usual metric on C, the
set of all complex numbers is defined by
d(z, w) = |z − w|, z, w ∈ C,
and on Ck , the usual metric d is defined by
p
d(z, w) := |z1 − w1 |2 + · · · |zk − wk |2 ,
for z = (z1 , . . . , zk ) and w = (w1 , . . . , wk ) in Ck .
8
Definition 1.10. On Rk (respectively, Ck ), the topology induced by the usual metric
is called the usual topology on Rk (respectively, Ck ). ♢
Theorem 1.11. The topologies induced by any two equivalent metrics are the same.
Proof. Suppose d and ρ are equivalent metrics on X. Then there exists c1 , c2 > 0 such
that
c1 d(x, y) ≤ ρ(x, y) ≤ c2 d(x, y) ∀ x, y ∈ X.
We have to show that Td = Tρ . So, let A ∈ Td . Let x ∈ A. Then there exists r > 0
such that Bd (x, r) ⊆ A; that is,
y ∈ X, d(x, y) < r ⇒ y ∈ A.
Note that
ρ(x, y) < c1 r ⇒ d(x, y) < r.
Thus, we obtain Bρ (x, c1 r) ⊆ Bd (x, r). In particular, Bρ (x, c1 r) ⊆ A. Thus, for each
x ∈ A, there exists ε := c1 r > 0 such that Bρ (x, ε) ⊆ A. This shows that A ∈ Tρ , thus,
proving Td ⊆ Tρ . Analogously, it can be shown that Tρ ⊆ Td .
Exercise 1.12. Consider the metrics d1 , d2 , d∞ on R2 defined by
9
• S is said to be closed under
Sn finite unions if for any finite number of members of
S, say A1 , . . . , An ∈ S, i=1 Ai ∈ S.
• S is said to be closed
S under arbitrary unions if for any subcollection of S, say
A ⊆ S, we have A∈A A ∈ S.
Thus, we see that, if (X, T ) is a topological space and if C is the collection of all closed
sets in X, then
Theorem 1.16. Let (X, T ) be a topological space and C be the collection of all closed
sets in X. Then
{Aα ⊆ X : α ∈ Λ}
10
T T T
Similarly, the intersection of all members of A, written as A∈A A or A is α∈Λ Aα ,
that is, \ \
A= Aα .
A∈A α∈Λ
(ii) Let T consist of ∅ and all those subsets of X such that Ac is finite, together
with ∅ and X. Then T is a topology on X, called the co-finite topology on X.
(iii) Let T consist of ∅ and all those subsets A of X such that Ac is countable.
Then T is a topology on X, called the co-countable topology on X. ♢
• On a finite set, co-finite topology and co-countable topology are the same. In
general, co-finite topology is weaker than co-countable topology; strictly weaker
if the underlying set is infinite.
Example 1.18. Let T be the collection of all subsets of R such that A ∈ T iff for
every x ∈ A, there exists r > 0 such that [x, x + r) ⊆ A. Then T is a topology on R,
called the lower limit topology on R. ♢
Notation: In the due course, we shall use R with lower limit topology by Rℓ .
Example 1.19. Let T be the collection of all subsets of R such that A ∈ T iff for
every x ∈ A, there exists r > 0 such that (x − r, x] ⊆ A. Then T is a topology on R,
called the upper limit topology on R. ♢
Remark 1.20. We have already noted that a set which is not open need not be closed;
a set which is not closed need not be open; a set can be both open and closed.
For instance, let R be with the topology induced by the usual metric. Then [0, 1)
is neither open nor closed, and the sets ∅ and R are both open and closed. We may
also note that there is no other set which is both open and closed. ♢
Exercise 1.21. Show that the indiscrete topology on a nonempty set X is induced by
a metric iff X is singleton, and in that case this topology is the discrete topology. ♢
Exercise 1.22. Show that the co-finite topology on a nonempty set X is induced by
a metric iff X is a finite set, and in that case this topology is the discrete topology. ♢
11
Exercise 1.23. Show that the co-countable topology on a nonempty set X is induced
by a metric iff X is a countbale set, and in that case this topology is the discrete
topology. ♢
One of the important property of a metric space Ω is its Hausdorff property, that is,
for every pair of distinct points x, y in Ω, there exists disjoint open sets containing x
and y respectively.
A ∩ B = ∅, x ∈ A, y ∈ B.
We may say that: A topological space X is Hausdorff space iff every pair of distinct
points in X can be separated by disjoint open sets.
Example 2.2. Let X be a set. The following can be verified easily (Verify!):
1. The indiscrete topology on X is Hausdorrf iff #(X) ≤ 1, and in that case the
topology is the discrete topology.
2. The co-finite topology on X is Hausdorrf iff X is a finite set, and in that case
the topology is the discrete topology.
From the statements in the above example, the following can be deduced (Exercise).
2
Felix Hausdorff (1868-1942) was a mathematician from Germany.
12
• If X contains more than one element, then the indiscrete topology on X is not
induced by any metric.
• If X is not a finite set, then the co-finite topology on X is not induced by any
metric.
• If X is an uncountable set, then the co-countable topology on X is not induced
by any metric.
Example 2.3. The lower limit topology and the upper limit topology on R are Haus-
dorff:
First consider Rℓ , the set R with lower limit topology. Let x, y ∈ Rℓ with x ̸= y.
Taking r = |x − y|/2, we see that the open sets A := [x, x + r) and B := [y, y + r)
disjoint open sets in Rℓ containing x and y, respectively. Thus, we have shown that Rℓ
is a Hausdorff space.
Similarly, it can be shown that R with upper limit topology is also Hausdorff. ♢
Remark 2.4. In Example 2.3, we have seen that the lower limit topology and the
upper limit topology on R are Hausdorff. However, they are not metrizable. - we shall
prove this fact after introducing some more notions (see Theorem 3.10). ♢
Example 2.5. For x = (x1 , x2 ) and y = (y1 , y2 ) in R2 , define
ρ(x, y) = |x1 − y1 |, x, y ∈ R2 .
Note that, ρ(·, ·) is a semi-metric or pseudo metric3 . For x ∈ X and r > 0, let
Bρ (x) := {u ∈ R2 : ρ(x, u) < ρ}.
Then
T := {A ⊆ R2 : ∀ x ∈ R2 , ∃ r > 0 such that Bρ (x) ⊆ A}
is a topology on R2 , which is not Hausdorff. ♢
Exercise 2.6. Verify the assertions in the above example. ♢
2.2 Convergence
13
• A sequence can have more than one limit!
xn ∈ A ∀ n ≥ n1 and xn ∈ B ∀ n ≥ n2 .
xn ∈ A and xn ∈ B ∀ n ≥ n0 .
This is not possible, since A and B are disjoint. Hence, our assumption that x ̸= y is
wrong. Thus, we have proved that x = y.
Theorem 2.11. Suppose X is with co-countable topology. Then every convergent se-
quence in X is eventually constant.
14
Example 2.12. Let X = {0} ∪ { n1 : n ∈ N} with co-finite topology. Consider the
sequence (1/n). Let G be an open subset of X containing 0. Then X \ G is a finite set.
Hence, there exists k ∈ N such that n1 ∈ G for all n ≥ k. Hence, n1 → 0. This shows
that, a convergent sequence in a co-finite topological space need not be eventually
constant. ♢
Example 2.13. Consider the sequences (1/n) and (−1/n) in R. We observe that both
the sequences converge to 0 with respect to the usual topology and (1/n) converges to
0 in Rℓ . But (xn ) := (−1/n) does not converge in Rℓ : For any x ≥ 0, the open set
[x, x + 1) in Rℓ containing x does not contain any members of the sequence, so that
xn ̸→ x in Rℓ . Also, if x < 0, taking 0 < r < |x|, the open set [x, x + r) can contain
only a finite number of members of the sequence, so that xn ̸→ x in Rℓ . ♢
Example 2.14. Let X be the set of all real valued functions defined on an interval
I := [a, b]. Let T ⊆ 2X be such that A ∈ T iff for every f ∈ A, there exists r > 0
such that {g ∈ X : |g(a) − f (a)| < r} ⊆ A. Then, T is a topology on X, which is not
Hausdorff. With respect to this topology,
fn → f ⇐⇒ fn (a) → f (a). ♢
Example 2.16. On R, consider the usual topology Tu and the lower limit topology Tℓ .
• Tu is weaker than Tℓ .
To see this, let A be an open set with respect to the usual topology, and let x ∈ A.
Then there exists r > 0 such that (x − r, x + r) ⊆ A. This also implies [x, x + r) ⊆ A.
Hence, A is open with respect to the lower limit topology. ♢
Example 2.17. On R, consider the usual topology Tu and
We see that T is a topology on R: To see this, note that {∅, R} ⊆ T and for every
a, b ∈ R,
a < b ⇒ (a, ∞) ∩ (b, ∞) = (a, ∞) ∈ T .
15
Now consider a collection of sets from T , which is of the form {(a, ∞) : a ∈ S} for
some S ⊆ R. Then [
(a, ∞) = (a0 , ∞) ∈ T ,
a∈S
Proof. Suppose (xn ) converges to x with respect to T2 . We show that (xn ) converges
to x with respect to T1 . For this, let A ∈ T1 be such that x ∈ A. Since A ∈ T2 , there
exists n0 ∈ N such that xn ∈ A for all n ≥ n0 . This shows, by definition of convergence
in T1 , that (xn ) converges to x with respect to T1 .
Example 2.19. We know that discrete topology on a set is stronger than any other
topology on it. Now, consider the sequence (1/n) in R.
• (1/n) converges to 0 with respect to the usual topology on R, but it does not
converge with respect to the discrete topology on R.
To see this, assume for a moment that (1/n) converges to some point x0 ∈ R. Consider
the set {x0 }. It is open with respect to the discrete topology on R. But, this open set
contains only one point, and hence (1/n) does not converge to x0 . ♢
Example 2.20. We know that, on R, lower limit topology is stronger than the usual
topology. Consider the sequence (−1/n). In example 2.13, we have seen that the above
sequence converges to 0 with respect to the usual topology. But,
• (−1/n) not converge to any point with respect to the lower limit topology. ♢
Example 2.21. Let T be the topology on R defined as in Example 2.17. We have
seen that this topology is weaker than the usual topology. Thus, every sequence which
converges with respect to the usual topology converges with respect to T as well. We
note that:
16
Indeed, if G is any open set containing x, then G is of the form G = (a, ∞) for some
a < x so that xn ∈ G for all n ∈ N.
• If (xn ) is any increasing sequence of real numbers and x ∈ R and k ∈ N such that
x ∈ [x1 , xk ], then xn → x with respect to T .
Indeed, if G is any open set containing x, then G is of the form G = (a, ∞) for some
a ∈ R, and hence xn ∈ G for all n ≥ k. ♢
Example 2.24. (i) The topology generated by the collection of all open intervals of
the form (a, b) with a < b is the usual toplogy on R.
(ii) The topology generated by the collection of all open intervals of the form (a, ∞)
and (−∞, b) is the usual toplogy on R. ♢
Example 2.25. Let S be the collection of all intervals of the form [a, b) with a < b.
Then the topology on R generated by S is the lower limit topology on R. ♢
Exercise 2.26. Show that the toplogy on R2 defined as in Example 2.5 is weaker than
the topology induced by the usual metric on R2 . ♢
17
2. A point x ∈ X is said to be an interior point of A if there exists an open set
containing x and contained in A. The set of all interior points of A is called the
interior of A, and it is denoted by A◦ or int(A).
We observe (Exercise):
A◦ ⊆ A ⊆ Ā = A ∪ A′ = A ∪ ∂A = A◦ ∪ ∂A.
• It is also true that Ā = A◦ ∪ ∂A. To see this, first we observe that A◦ ∪ ∂A ⊆ ∂A.
To see the other way inclusion, let x ∈ Ā. Then, every open set G containing
x contains some some point of A. If G does not contain any point of Ac , then
G ⊆ A so that x ∈ A◦ . Thus, we have proved that if x ̸∈ ∂A, then x ∈ A◦ . in
other words, we have proved that Ā ⊆ A◦ ∪ ∂A.
Example 2.28. Let R be with usual topology. Then we have the following:
1. If A = (0, 1], then A◦ = (0, 1), Ā = [0, 1], A′ = [0, 1], ∂A = {0, 1}.
2. If A ∈ {Q, Qc }, then A◦ = ∅, Ā = R, A′ = R, ∂A = R.
S ◦ = ∅, S̄ = S, S ′ = S, ∂S = S.
18
(ii) Let A be the open unit unit disc, and B be the closed unit disc, that is,
A◦ = A, Ā = B, A′ = B, ∂A = S,
B ◦ = A, B̄ = B, B ′ = B, ∂B = S. ♢
Example 2.30. Let R be with co-countable topology. Note that, with respect to this
topology, any open subset has to be an uncountable set, and every countable set is
closed. Hence,
Q̄ = Q, Q◦ = ∅, Qc = R, (Qc )◦ = Q◦ . ♢
Theorem 2.32. Let (X, T ) be a topological space, and A and B be subsets of X such
that A ⊆ B. Then
A◦ ⊆ B ◦ , Ā ⊆ B̄.
Proof. Suppose x ∈ A◦ . Then there exists an open set G such that x ∈ G ⊆ A. Since
A ⊆ B, this shows that x ∈ B ◦ . Thus, A◦ ⊆ B ◦ .
Next, suppose x ∈ Ā. Then, for every open set G containing x, G ∩ A ̸= ∅, so that
we also have G ∩ B ̸= ∅, showing that x ∈ B̄. Thus, Ā ⊆ B̄.
19
(4) A is open iff A◦ = A.
Proof. (1). To show that (Ā)c is an open set. If Ā = X, then its compliment is ∅,
which is an open set. So, assume that X̄ ̸= X. Let x ∈ (Ā)c , that is, x is not a closure
point of A. Then there exists an open set Gx containing x, which does not contain any
point of A. Also, if y ∈ Gx , then y ̸∈ Ā, since the open set Gx containing y does not
contain any point of A. Thus, Gx ∩ Ā = ∅, so that Gx ⊆ (Ā)c . Therefore,
[
(Ā)c ⊆ Gx ⊆ (Ā)c .
x∈(Ā)c
Proof. (1) Recall that Ā = A ∪ ∂A, and A is closed iff Ā = A. This shows that
A closed ⇐⇒ A ∪ ∂A = A ⇐⇒ ∂A ⊆ A.
(2) It is enough to show that A is not open iff A contains a boundary point of A.
First we recall that A is open iff A = A◦ .
20
Suppose that A is not open. Then there exists x ∈ A \ A◦ , which implies that
every open neighbourhood of x contains some point from Ac . This shows that x is a
boundary point of A.
Conversely, suppose A contains a boundary point x of A. Then every open neigh-
bourhood of x contains a point from Ac so that x ̸∈ A◦ = A, so that A is not open.
(3) We know that A◦ is an open set. Suppose G is an open set such that G ⊆ A.
Then we have G = G◦ ⊆ A◦ . Thus, A◦ is the largest open set contained in A.
(4) We know that Ā is a closed set. Suppose F is a closed set such that A ⊆ F .
Then we have Ā ⊆ F̄ = F . Thus, Ā is the smallest closed set containing A.
Example 2.35. Let X is with the discrete topology and A ⊆ X. Then A◦ = A = Ā.
♢
Proof. We observe that x ∈ A is an isolated point of A iff there exists an open set V
in X which does not conain any other point of A iff there exists an open set V in X
such that V ∩ A = {x} iff {x} is open is open in A
Proof. Suppose x ∈ A′ . Assume for a moment that there exist an open neighbourhood
G of x such that G ∩ A contains only finite number of members. Since x ∈ A′ , we know
that G ∩ (A \ {x}) ̸= ∅. Let G ∩ (A \ {x}) = {x1 , . . . , xn } for some n ∈ N. Since X
21
is Hausdorff, for each i ∈ {1, . T
. . , n}, there exist disjoint open sets Ui abd Vi such that
x ∈ Ui and xi ∈ Vi . Let U := ni=1 Ui . Then we have
x ∈ U ∩ G, {x1 , . . . , xn } ∩ U = ∅,
so that U ∩ (G ∩ (A \ {x}) = ∅. This is a contradiction to the fact that the open set
U ∩ G must contain some point from A \ {x}.
The converse is obvious from the definition of A′ .
Definition 2.41. (Nowhere dense) Let X be a topological space and A ⊆ X. Then
A is said to be nowhere dense in X if (Ā)◦ = ∅. ♢
Example 2.42. Let X = R with usual topology. Then Z is nowhere dense in X, since
Z is closed and Z◦ = ∅, and Q is not nowhere dense, since (Ā)◦ = R◦ ̸= ∅. ♢
Theorem 2.43. Let X be a topological space and A be a closed set in X. Then A is
nowhere dense in X iff Ac is dense in X.
Proof. We observe that A is nowhere dense in X iff A◦ = ∅ iff A does not contain any
nonempty open set iff every nonempty open set contains some point from Ac iff Ac is
dense in X.
Remark 2.44. In Theorem 2.43, the condition that A is closed cannot be dropped.
To see this, let X = R with usual topology and A = Q. Then we know that Ac = R \ Q
dense, but (Ā)◦ = R◦ = R. ♢
Let (X, T ) be a topological space and Y ⊆ X. Then the topology T induces a topology
on Y in a natural way as per the following theorem.
Theorem 2.45. Let (X, T ) be a topological space, and Y ⊆ X. Then
TY := {G ∩ Y : G ∈ T }
is a topology on Y .
A1 = G1 ∩ Y, A2 = G2 ∩ Y
so that
A1 ∩ A2 = (G1 ∩ G2 ) ∩ Y ∈ TY .
22
Next, suppose {Aα : α ∈ Λ} ⊆ TY . Then, for each α ∈ Λ, there exists Gα ∈ T such
that Aα = Gα ∩ Y . Hence,
[ [
Aα = Gα ∩ Y ∈ TY .
α∈Λ α∈Λ
TY ⊆ T ⇐⇒ Y ∈ T .
23
Example 2.49. Let X = R be with usual topology and Y = [0, 1] be with subspace
topology. Then (0, 1) is open in Y and X, though Y is not open in X. ♢
{A ⊆ Y : A ∈ T } ⊆ TY .
{A ⊆ Y : A ∈ T } = TY ⇐⇒ Y ∈ T .
X \ A = (X \ Y ) ∪ (Y \ A).
X \ A = (X \ Y ) ∪ (Y ∩ G) = Y c ∪ (Y ∩ G) = (Y c ∪ Y ) ∩ (Y c ∩ G) = Y c ∩ G.
24
Thus, X \ A is open in X.
Conversely, suppose Y is closed in X and A ⊆ Y is closed in X. Since Y \A = Y ∩Ac ,
where Ac is open in X, we obtain that Y \ A is open in Y , so that A is closed in Y .
Theorem 2.54. Let X be a topological space and Y ⊆ X. Let CX and CY denote the
set of all closed sets in X and Y , respectively. Then,
CY ⊆ CX ⇐⇒ Y ∈ CX .
Y \A=G∩Y
Y \ A = Y \ (F ∩ Y ) = Y ∩ (F ∩ Y )c = Y ∩ (F c ∪ Y c ) = Y ∩ F c .
Bases are, in some sense, building blocks for topological spaces. Recall that, in a metric
space Ω, a subset A of is open iff for every x ∈ A, there exists an open ball B(x, rx )
for some rx > 0 such that B(x, rx ) ⊆ A. Also, we know that every open set is a union
of open balls. Motivated by this, we define the notion of a base for a topology.
25
Definition 2.56. (Base) Let (X, T ) be a topological space and B be a sub-collection
of T . Then B is called a base or basis for T if for every G ∈ T and x ∈ G, there
exists B ∈ B such that x ∈ B ⊆ G. . ♢
Example 2.57. Let B be the collection of all open intervals of the form (a, b) with
a < b. Then B is a base for the usual topology on R. ♢
Example 2.58. Let B be the collection of all intervals of the form [a, b) with a < b.
Then B is a base for the lower limit topology on R. ♢
Example 2.59. Let B be the collection of all intervals of the form (a, b] with a < b.
Then B is a base for the lower limit topology on R. ♢
Example 2.60. If X is a metric space with metric d, then the collection of all open
balls in X is a base for Td . ♢
Example 2.61. If X is a discrete topological space then the set of all singleton sets
is a base for the topology. ♢
Example 2.62. If X is an indiscrete topological space then {X} and {∅, X} are the
only bases for the topology. ♢
Suppose X is a set. Then a given collection of subsets of X need not form a base
for some topology on X. However, we have the following two theorems.
26
Theorem 2.65. Let X be a set and B be a family of subsets of X such that B covers
X and B is closed under finite intersections. Then B is a base for a (unique) topology
on X.
Proof. Let T be the family of all unions of members of B together with the empty set.
Then, T is a topology for which B is a base. Further, by Theorem 2.64, this is the only
topology for which B is a base.
Recall that, if (Ω, d) is a metric space and B := {B(x, r) : x ∈ X, r > 0}, then the
metric topology that we defined is
Motivated by this, for any set X and for a collection B of subsets of X, let us consider
Theorem 2.66. Let X be a set and B be a family of subsets of X such that B covers X
and for every B1 , B2 ∈ B and x ∈ B1 ∩ B2 , there exists B3 ∈ B with x ∈ B3 ⊆ B1 ∩ B2 .
Then
T := {A ⊆ X : for every x ∈ A, ∃ B ∈ B such that x ∈ B ⊆ A}.
is a topology on X and B is a (unique) base for T .
Remark 2.68. We may observe that, in the context of a metric space (Ω, d), the
conditions in Theorem 2.66 are satisfied: Note that
[
Ω= B(x, r)
x∈Ω
27
for any r > 0. Now, consider open balls B(x, r) and B(y, s) for some x, y ∈ Ω and
r, s > 0. Let u ∈ B(x, r) ∩ B(y, s). Then, taking
it can be seen that B(u, t) ⊆ B(x, r) ∩ B(y, s). Thus, we have shown that the family
B := {B(x, r) : x ∈ X, r > 0}
Example 2.70. Let S be the collection of all intervals of the form (a, ∞) and (−∞, b).
Then S is a sub-base for the usual topology Tu on R, but S is not a base for Tu . For
instance, (−1, 1) is an open set containing 0. But, there is no set in S containing 0
and contained in (−1, 1). ♢
Proof. Suppose S is a sub-base for T . Then, by Theorem 2.71 and Theorem 2.63, B
is a base for T . Conversely, suppose B is a base for T . Then, by Theorem 2.63, S is a
subs-base for T .
We may recall that the set P of all prime numbers is infinite. One of the simplest
proofs of this fact is due to Euclid:
28
Suppose P is a finite set, say P = {p1 , . . . , pn } for some n ∈ N. Consider
the number
q = 1 + p 1 p2 · · · pn .
Then none of the pi divides q. Hence, q is a prime. But, q ̸∈ P, which is a
contradiction.
Now, we give a topological proof for the infinitude of primes, by using a topology on
Z as follows:
For (a, b) ∈ Z × N, consider the set
Na,b := {a + bn : n ∈ Z}.
Define
T := {A ⊆ Z : ∀ a ∈ A, ∃b ∈ N such that Na,b ⊆ A}.
The following can be verified (Exercise):
• T is a topology on Z;
This shows that, if P is a finite set, then Z \ {1, −1} is a closed set, which implies the
finite set {1, −1} is an open set. Thus, we arrive at a contradiction.
Exercise 2.73. Show that the set B := {Na,b : a ∈ Z, b ∈ N} is a base for the topology
T above. ♢
29
be a countable collection4 . We may also observe that, if a base for a topology contains
only a finite number of sets, then the topology itself contains a finite number of sets.
Thus, if a topology consists of infinite number of sets, then every base of such topology
must contain infinite number of sets.
In view of the above observations we introduce the following definition.
Example 3.2. The space Rk with usual topology is separable and second countable.
This follows from the above theorem since the countable set D := Q × Q is dense in
R2 . In fact, the family
{B(u, r) : u ∈ D, r ∈ Q+ }
is a countable base for Rk with usual topology. ♢
Remark 3.4. Suppose S is any countable family of subsets of a set X. In the following
if we shall write S = {Sn : n ∈ N}, with the understanding that if S is a finite collection,
say S = {S1 , . . . , Sk } for some k ∈ N, then we take Sn = Sk for all n ≥ k. ♢
For the next theorem, recall that a subset D of a topological space X is dense in X
if D̄ = X, that is, for every x ∈ X, there exists an open neighbourhood G of X such
that G ∩ D ̸= ∅.
Theorem 3.5. Every metric space with a countable dense subset is second countable.
Proof. Let Ω be a metric space with a metric d and let D be a countable dense subset
of Ω. Let (εn ) be a sequence of positive reals such that εn → 0. Consider the family
A := {B(u, εn ) : u ∈ D, n ∈ N}.
Clearly, A is a countable family of open sets. We show that A is a base for X. For
this, let A be an open set in Ω and x ∈ A. Let r > 0 be such that B(x, r) ⊆ A. Let
4
If X = Z with discrete topology, then the cardinality of the base of singleton sets is ℵ0 , whereas
the cardinality of the topology is c = 2ℵ0 , the continuum.
30
n ∈ N be large enough such that εn < r/2. Since D is dense in Ω, the ball B(x, εn )
contains some point u ∈ D. Then
x ∈ B(u, εn ) ⊆ B(x, r) ⊆ A.
so that B(u, εn ) ⊆ B(x, r) as well. Thus, we have proved that A is a countable base
for Ω.
Example 3.7. The metric space X = Rk with usual metric is separable. For example,
the set Qk is a countable dense subset of X. ♢
Example 3.8. Every indiscrete topological space X separable and second countable.
In fact, the only base of the indiscrete space X in {X}, and every nonempty subset of
X is dense in X, in particular, every countable subset of X is dense in X. ♢
We have proved that (see Theorem 3.5) every separable metric space is second
countbale.
Theorem 3.10. The lower limit topology and the upper limit topology on R are sepa-
rable, but not second countable. In particular, they are not metrizable.
31
Proof. Let Rℓ := R with lower limit topology.
(a) It can be easily seen that Q is dense in Rℓ , and hence it is separable (Exercise).
(b) Let B be a base for Rℓ . Then for each x ∈ Rℓ , there exists Bx ∈ B such
that x ∈ Bx ⊆ [x, ∞). Clearly, inf Bx = x. Hence, x ̸= y implies Bx ̸= By . Hence,
{Bx : x ∈ Rℓ } is uncountable, and therefore, B is uncountable. Thus, Rℓ is not second
countable.
(c) Since Rℓ is separable and not second countable, by Theorem 3.5, it is not
metrizable.
We know that, with respect to the usual metric, R is separable. Also, as a subspace
of R, Q is separable. Is R\Q separable? The answer is in the affirmative, which follows
from the following general result.
Proof. Let (Ω, d) be a separable metric space and Ω0 be a subset of Ω. We show that
Ω0 is also separable. Clearly, if Ω0 is countable, then it is separable. So assume that Ω0
is uncountable. Then D has to be countably infinite (Why?). Let D = {xn : n ∈ N}.
Then, for each n ∈ N, we have
∞
[
Ω= B(xi , 1/n)
i=1
so that ∞
[
Ω0 = B(xi , 1/n) ∩ Ω0 ∀ n ∈ N.
i=1
Also, for every n ∈ N, there exists j ∈ N such that B(xj , 1/n) ∩ Ω0 ̸= ∅. Let
We show that D0 is dense in Ω0 . For this, let x ∈ Ω0 and ε > 0 be given. We have to
show that B(x, ε) ∩ D0 ̸= ∅. Note that, for every n ∈ N, there exists j ∈ N such that
x ∈ B(xj , 1/n) so that (j, n) ∈ ∆. Let n be large enough such that 2/n < ε. Then, we
have
1 1
d(x, yj,n ) ≤ d(x, xj ) + d(xj , yj,n ) < + < ε.
n n
Thus, B(x, ε) ∩ D0 ̸= ∅, and the proof is complete.
32
At this point, one may ask: Is subspace of every separable topological space sepa-
rable? The answer to this question is in the negative, as the following example shows.
Example 3.12. Let T consists of ∅ and those A ⊆ R2 such that for every (x, y) ∈ A,
there exists r > 0, s > 0 such that
[x, x + r) × [y, y + s) ∈ A.
for any r > 0. Hence, every singleton subset of X0 is open in X0 , so that the subspace
topology on X0 is the discrete topology. Since an uncountable set with discrete topology
is not seprable, X0 is not seprable. ♢
Definition 3.13. Let (X, T ) be a topological space and x ∈ X. A family Bx of open
neighbourhoods of x is said to be a base at x or local base at x if every open set
containing x contains some member of Bx . ♢
Definition 3.14. (First countable space) A topological space X is said to be first
countable if it has a countable local base at every point in X. ♢
Example 3.15. (i) In R with usual topology, the set Bx := {(x−1/n, x+1/n) : n ∈ N}
is a local base at x ∈ R.
(ii) If R is with lower limit topology, then Bx := {[x, x + 1/n) : n ∈ N} is a local
base at x ∈ R for every x ∈ R.
(ii) In a discrete space X, the family Bx = {{x}} is a local base at x.
(iii) In an indiscrete space X, Bx = {X} is the only local base at x. ♢
Theorem 3.16. Every metric space is first countable.
Proof. Let (Ω, d) be metric space. Then, for every x ∈ Ω, the family Bx := {Bd (x, 1/n) :
n ∈ N} is a countable base at x: For every open set A containing x, there exists r > 0
such that Bd (x, r) ⊆ A. Taking n ∈ N such that 1/n < r, we have Bd (x, 1/n) ⊆
Bd (x, r) ⊆ A.
Theorem 3.17. Every second countable space is first countable, and the converse is
not true.
33
Proof. Let (X, T ) be a second countable space, and let B be a countable base for X.
Let x ∈ X, and let
Bx := {B ∈ B : x ∈ B}.
Then Bx is a countable base at x. Indeed, for every open set A containing x, there
exists Bx ∈ B such that x ∈ B ⊆ A. Clearly Bx ∈ Bx . This shows that Bx is a
countable base at x.
The converse of the above theorem is not true. To see this, it is enough consider an
uncountable discrete space X. This space is not second countable. However, for each
x ∈ X, the family Bx := {{x}} is a base at x.
Every metric space is first countable, but not necessarily separable and
second countable.
We observe that:
Theorem 3.18. If (xn ) is a sequence in a subset A of a topological space such that
xn → x, then x ∈ Ā.
However, if X is a first countable space, then converse to Theorem 3.18 does hold.
Theorem 3.20. Suppose X is a first countable space and A ⊆ X. Then x ∈ Ā iff
there exists a sequence (xn ) in A such that xn → x.
Gn ∩ A ⊆ Bk ⊆ G ∀n ≥ k.
34
In particular, xn ∈ G for all n ≥ k. Thus, we have proved that xn → x.
Conversely, suppose (xn ) is a sequence in A such that xn → x. Let G be any open
set containing x. Then, by the definition of convergence, there exists k ∈ N such that
xk ∈ G. In particular, G ∩ A ̸= ∅, showing that x ∈ Ā.
Recall that:
Proof. We have already proved that if X is Hausdorff, then every convergent sequence
in X has a unique limit.
Conversely, suppose X is first countable and every convergent sequence in X has
a unique limit. To show that X is Hausdorff. For a moment assume that X is not
Hausdorff. Then there exists u, v ∈ X with u ̸= v such that every pair of open sets
containing u and v, respectively, will intersect. Now, let Bu = {Un : n ∈ N} and
Bv = {Vn : n ∈ N} be countable bases at u and v, respectively. Let
n
\ n
\
An = Uj , Bn = Vj ∀ n ∈ N.
j=1 j=1
Then An and Bn are open sets containing u and v, respectively, and hence An ∩Bn ̸= ∅.
Let xn ∈ An ∩ Bn . We show that xn → u and xn → v, which would contradict the
35
hypothesis. For this, let G be an open set containing u. Since Bu is a base at u, there
exists k ∈ N such that Uk ⊆ G. Hence An ⊆ G for all n ≥ k, so that we also have
xn ∈ An ∩ Bn ⊆ An ⊆ G ∀ n ≥ k.
Thus, xn → u. Analogously, using the fact that Bv is a local base at v, we see that
xn → v. This completes the proof.
4.1 Motivation
From the theory of metric spaces, recall the following: Suppose (X, d) and (Y, ρ) are
metric spaces. Recall that a function f : X → Y is continuous at a point x0 ∈ X if for
every ε > 0, there exists a δ > 0 such that
That is,
x ∈ BX (x0 , δ) ⇒ f (x) ∈ BY (f (x0 ), ε) ⊆ V.
Thus, taking U = BX (x0 , δ), we obtain
36
Conversely, suppose that for every open neighbourhood V of f (x0 ), there exists an
open neighbourhood U of x0 such that
To show that f is continuous at x0 . For this, let ε > 0 be given and V = B(f (x0 ), ε).
Then, by hypothesis, there exists an open neighbourhood U of x0 such that x ∈ U
implies f (x) ∈ V . Let δ > 0 be such that BX (x0 , δ) ⊆ U . Thus, we have
Thus, we have proved that for every ε > 0, there exists δ > 0 such that d(x, x0 ) < δ
imples ρ(f (x), f (x0 )) < ε, that is, f is continuous at x0 .
For our later references, let us recall the following results from real analysis.
37
2. Let X be a nonempty set and T1 and T2 be topologies on X. Let X1 = X with
T1 and X2 = X with T2 . Then the identity map from X1 to X2 is continuous iff
T1 is stronger than T2 . ♢
Remark 4.7. In some of the books on topology, the continuity of a function is defined
in terms of the above characterization. ♢
The proof of the following theorem follows using the definitions of base and sub-
base for a topology: Recall that a family B of open sets is a base for a topology T if
every open set is a union of members of B, and a family S of open sets is a sub-base
for a topology T if the collection of all finite intersections of members of S form a base
for T .
Theorem 4.9. Let X and Y be topological spaces, B be a base and S be a sub-base for
the topology on Y . The, for a function f : X → Y , the following are equivalent:
(i) f is continuous.
38
Example 4.10. Let X be any topological space and R be with usual topology. Then
a function f : X → R is continuous iff
{x ∈ X : a < f (x) < b}
is open in X for any a, b ∈ R with a < b. ♢
Theorem 4.11. Let X and Y be topological spaces, and let f : X → Y be a continuous
function. Then, for any sequence (xn ) in X,
xn → x in X ⇒ f (xn ) → f (x) in Y.
Proof. Let (xn ) be a sequence in X such that xn → x in X. To show that f (xn ) → f (x).
For this, let A be an open set containing f (x). Since f is continuous, f −1 (A) is open
in X. Also, x ∈ f −1 (A). Since xn → x and f −1 (A) is an open set containing x, there
exists n0 ∈ N such that xn ∈ f −1 (A) for all n ≥ n0 . This implies that f (xn ) ∈ A for
all n ≥ n0 . Thus, f (xn ) → f (x).
Remark 4.12. By the above theorem. if we can find a sequence (xn ) in X such that
xn → x for some x ∈ X and f (xn ) ̸→ f (x), then f is not continuous. ♢
• Converse of Theorem 4.11 is not true. To see this, look at the following example.
Example 4.13. Let Rcc = R with co-countable topology and Rdis = R with discrete
topology. Consider the identity function f : Rcc → Rdis , that is, f (x) = x for every
x ∈ R. This function is not continuous. For example, {1} is an open set in Rdis , but
f −1 ({1}) is not open, as its compliment is not countable. However, for every sequence
(xn ) in R,
xn → x in Rcc ⇒ f (xn ) → f (x) in Rdis .
This is true, because, if xn → x in Rcc , then (xn ) is eventually constant, and hence
xn → x in Rdis . ♢
39
Definition 4.15. (Sequential continuity) Let X and Y be topological spaces. Then
a function is said to be sequentially continuous if for any sequence (xn ) in X,
xn → x in X ⇒ f (xn ) → f (x) in Y. ♢
Proof. ⇒): Suppose f is continuous. Let x ∈ Ā. We have to show that f (x) ∈ f (A).
Let V be an open neighbourhood of f (x). To show that V contains some point from
f (A). Since f is continuous, we know that f −1 (V ) is an open neighbourhood of x, and
since x ∈ Ā, the open neighbourhood f −1 (V ) of x contains some point from A, say
u ∈ f −1 (V ) ∩ A. Then f (u) ∈ V ∩ f (A).
40
⇐): Suppose f (Ā) ⊆ f (A) for every A ⊆ X. We have to show that f is continuous.
By Theorem 4.8, it is enough to show that f −1 (B) closed for every closed B ⊆ Y .
That is, taking A = f −1 (B), we have to show that A = Ā. So, let x ∈ Ā. Then, by
hypothesis, f (x) ∈ f (A). But, since A = f −1 (B), we have f (A) ⊆ B, so that
f (x) ∈ f (A) ⊆ B̄ = B.
• f is continuous.
• For every open set V in Y , f −1 (V ) is open in X.
• For every closed set F in Y , f −1 (F ) is closed in X.
•For every set A ⊆ X, f (Ā) ⊆ f (A).
Proof. Let X and Y be topological spaces and X be a first countable space. Suppose
f : X → Y is sequentially continuous. We have to show that f is continuous. By
Theorem 4.17, it is enough to show that f (Ā) ⊆ f (A) for every A ⊆ X. So, let A ⊆ X
and x ∈ Ā. Since X is first countable, by Theorem 3.20, there exists (xn ) in A such
that xn → x. By assumption f (xn ) → f (x). Since f (xn ) ∈ f (A), it follows that
f (x) ∈ f (A), and the proof is complete.
41
Proof. Let X and Y be topological spaces and f : X → Y be a continuous function.
Let D be a dense subset of X. We show that f (D) is dense in f (X). For this, let
y ∈ f (X) and V be an open set in f (X) which contains y. Then V = G ∩ f (X), where
G is open in Y and y ∈ G. We have to show that f (D) ∩ V ̸= ∅.
Let x ∈ X be such that y = f (x). Since f is continuous, f −1 (G) is an open
neighbourhood of x. Since D is dense in X, D ∩ f −1 (G) ̸= ∅. Let u ∈ D ∩ f −1 (G).
Clearly, f (u) ∈ f (D) ∩ G ⊆ f (D) ∩ V . Thus, f (D) ∩ V ̸= ∅ and the proof is
complete.
42
4.4 Restriction and expansion and composition theorems
The proofs of all the items in the following theorem follow from the definition of
continuity. However, we supply brief explanations.
Theorem 4.23. Let X and Y be topological spaces and f : X → Y be a continuous
function. Then the following results hold.
43
Theorem 4.24. (Compositiion thorem) Let X, Y and Z be topological spaces and
f : X → Y and g : Y → Z be continuous functions. Then the composition function
g ◦ f : X → Z is continuous.
(g ◦ f )−1 (V ) = f −1 [g −1 (V )],
which is open in X.
Definition 4.25. Given a function f : X → Y and a subset X0 of X, the function
f0 : X0 → Y defined by
f0 (x) = f (x), x ∈ X0 ,
is called the restriction of f to X0 , and it is usually denoted by f |X0 , and in that
case, f is called an extension of f0 . ♢
Example 4.26. Let X = R be with usual topology and X0 = [0, 1] be with the
subspace topology. Then f0 : X0 → X defined by
f0 (x) = x, x ∈ X0 ,
Example 4.27. Let R be with usual topology and ((0, ∞) is with subspace topology.
Let f0 : (0, ∞) → R be defined by
1
f0 (x) = , x ∈ (0, ∞).
x
Then f0 is continuous. But, it does not have a continuous extension to R (Why?). ♢
44
Theorem 4.28. Let X, Y be topological spaces, X0 be a subspace of X and f0 : X0 → Y
be continuous. If X0 is open as well as closed, then f0 has a continuous extension to
all of X.
Proof. Let f1 : X0c → Y be any continuous function (e.g., we may take f1 (x) = y0 for
every x ∈ X0c for any given y0 ∈ Y ). Then define f : X → Y by
f0 (x) if x ∈ X0 ,
f (x) =
f1 (x) if x ∈ X0c ,
Now, let V be an open set in Y . Then we have
Hence,
f −1 (V ) = f0−1 (V ) ∪ f1−1 (V ).
Since f0 and f1 are continuous, the sets f0−1 (V ) and f1−1 (V ) are open in X0 and X0c ,
respectively. Now, since X0 and X0c are open in X, the sets f0−1 (V ) and f1−1 (V ) are
open in X as well. Thus, f −1 (V ) is open in X. Thus, we have proved that f is a
continuous extension of f0 .
Exercise 4.29. Does R with usual topology have a nonempty proper subset which is
open and closed? Why? ♢
Theorem 4.28 is a particular case of the following theorem, called the pasting lemma.
Theorem 4.30. (Pasting lemma) Let X and Y be topological spaces, and A and B
be closed subspaces of X such that X = A ∪ B. If f1 : A → Y and f2 : B → Y are
continuous functions such that
f1 (x) = f2 (x) ∀ x ∈ A ∩ B,
Now, if S is closed in Y , then f1−1 (S) closed in A and f2−1 (S) closed in B, and since A
and B are closed, f1−1 (S) and f2−1 (S) are closed in X as well. Thus, we have proved
that inverse every closed set under f is closed in X.
45
Exercise 4.31. Prove the pasting lemma (Theorem 4.30) by assuming that both A
and B are closed, instead of assuming that both are open. ♢
and ρ(f0 (xn ), f (x)) → 0, it is enough to show that ρ(f0 (un ), f0 (xn )) → 0.
Now, let ε > 0 be given. Since f0 is uniformly continuous, there exists δ > 0 such
that
d(v, w) < δ ⇒ ρ(f (v), f (w)) < ε. (1)
46
Since d(xn , x) → 0 and d(un , x) → 0, there exists N ∈ N such that
so that
d(xn , un ) ≤ d(xn , x) + d(x, un ) < δ ∀ n ≥ N. (2)
Hence, by (1) and (2) above,
Thus, we have shown that ρ(f (un ), f (x)) → 0, and the the value f (x) is is uniquely
defined, making f : X → Y a function.
(ii) Note that
f (x) = f0 (x) ∀ x ∈ X0 .
Thus, f0 is an extension of f0 .
(iii) f is continuous: For this, let x ∈ X and ε > 0 be given. Since f0 is uniformly
continuous, there exists δ > 0 such that
For a δ0 > 0, let u ∈ X be such that d(x, u) < δ0 , and let (xn ) and (un ) in X0 such
that d(xn , x) → 0 and d(un , u) → 0. We note that
ρ(f (x), f (u)) ≤ ρ(f (x), f0 (xn )) + ρ(f0 (xn ), f0 (un )) + ρ(f0 (un ), f (u)) (2)
for all n ∈ N. By the definition of f (x) and f (u), we know that ρ(f (x), f0 (xn )) → 0
and ρ(f (un ), f0 (u)) → 0. Hence, there exists N1 ∈ N such that
Note that
d(xn , un ) ≤ d(xn , x) + d(x, u) + d(u, un ) ∀ n ∈ N.
Since d(xn , x) → 0 and d(u, un ) → 0, there exists N2 ∈ N such that
δ δ
d(xn , x) < , d(u, un ) < ∀ n ≥ N2 .
4 4
Hence, taking δ0 := δ/2, we have
δ
d(x, u) < ⇒ d(xn , un ) < δ ∀ n ≥ N2
2
so that by (1),
ρ(f (xn ), f (un )) < ε ∀ n ≥ N2 .
47
Therefore, by (2) and (3),
d(x, u) < δ/2 ⇒ ρ(f (x), f (u)) < 3ε ∀ n ≥ N := max{N1 , N2 }.
Thus, we have proved that f is continuous.
(iv) The function f in (iii) is the unique continuous extension of f0 : Suppose
g : X → Y is any continuous extension of f0 . Then for every x ∈ X, if (xn ) in X0 is
such that d(xn , x) → 0, then we have
g(x) = lim g(xn ) = lim f0 (xn ) = f (x).
n→∞ n→∞
Thus, g = f .
We observe that
X ∼ Y ⇐⇒ Y ∼ X.
48
Example 4.37. (A continuous bijection which is not a homeomorphism) Let
X = R be with discrete topology and Y = R with usual topology. Then the identity
map f : X → Y defined by
f (x) = x, x ∈ X,
is continuous and bijective, but not a homeomorphism: Clearly, f is bijective, and
since X is with discrete topolgy, f is continuous as well. Note that, g := f −1 : Y → X
is defined by g(y) := f −1 (y) = y for every y ∈ Y , and the inverse image of the open
set {1} in X under the map g is not open in X. ♢
Example 4.38. (A homeomorphism between open intervals) For any a, b, c, d
with a < b and c < d, the open intervals (a, b) and (c, d) are homeomorphic with udual
topologies on them: To see this consider the function f : (a, b) → (c, d) be defined by
d−c
f (x) = c + (x − a).
b−a
Then we see that f is continuous, bijective, and its inverse is given by
b−a
f −1 (y) = a + (y − c),
d−c
which is also continuous.
Using the same arguments, [a, b] is homeomorphic with [c, d], [a, b) is homeomorphic
with [c, d) and (a, b] is homeomorphic with (c, d]. ♢
For intervals with end points a and b, one may ask the following questions:
49
• Is [a, b) ∼ (a, b]?
We shall see that answer to only the last question is in the affirmative.
Example 4.39. (Non-homeomorphic spaces) We have following.
(1) With topologies induced by the usual metric, the space X = [a, b] is not home-
omorphic with any of the spaces Y1 = (a, b), Y2 = (a, b], Y3 = [a, b): Note that X
is compact, whereas Y1 , Y2 , Y3 are not compact. Recall that, continuous image of a
compact metric space is compact. Hence, there is no continuous onto function from X
to any of Y1 , Y2 , Y3 .
(2) With topologies induced by the usual metric, the space X = (a, b) is not home-
omorphic with any of the spaces Y1 = (a, b], Y2 = [a, b).
To see this, first consider X and Y1 . Assume for a moment that X and Y are
homeomorphic. Then there is a continuous onto function g : Y1 → X. Let x0 = g(b).
The
A = Y1 \ {b} = (a, b), g(A) = (a, x0 ) ∪ (x0 , b).
Note that A is a connected subset of Y1 whereas g(A) is disconnected, which is a
contradiction to the fact that, continuous image of a connected set has to be connected.
Thus, we have proved that X = (a, b) is not homeomorphic with Y1 = (a, b].
Using similar arguments, it can be shown (Verify) that X = (a, b) is not homeo-
morphic with Y2 = [a, b). ♢
Example 4.40. (Homeomorphic spaces) With topologies induced by the usual
metric, the space X = [a, b) is homeomorphic Y = (a, b]. To see this, let φ : (a, b] →
[−b, −a) be defined by
φ(x) = −x.
This function is a homeomorphism from (a, b] onto [−b, −a), that is [a, b) ∼ [−b, −a).
Hence, in view of Example 4.38, [−b, −a) ∼ [a, b). Hence, by the transitivity of the
relation ∼, we have [a, b) ∼ (a, b]. Since (a, b] ∼ (c, d], we also have [a, b) ∼ (c, d]. ♢
(a, b) ∼ (c, d), [a, b) ∼ [c, d), (a, b] ∼ (c, d], (a, b] ∼ [c, d), [a, b] ∼ [c, d]
(a, b) ̸∼ [a, b], (a, b) ̸∼ [a, b), (a, b) ̸∼ (a, b], (a, b] ̸∼ [a, b], [a, b) ̸∼ [a, b].
Remark 4.41. Bear in mind that, to show that two spaces are homeomorphic, it is
enough to find one homeomorphism from X onto Y , whereas to show that two spaces
are not homeomorphic, it is necessary to show that there is no homeomorphism from
X onto Y . ♢
50
Example 4.42. (A homeomorphism) Let (0, 1) and (1, ∞) be with usual topology,
and f : (0, 1) → (1, ∞) be defined by
1
f (x) = , x ∈ (0, 1).
x
We note that f is bijective with its inverse defined by
1
f −1 (y) = , y ∈ (1, ∞).
y
Further, both f and f −1 are continuous (using a result from Real Analysis). Thus, f
is a homeomorphism.
Now, let c ∈ R. Let fc : (0, 1) → (c, ∞) be defined by
1−x 1−x
fc (x) = c + −1=c+ , x ∈ (0, 1).
x x
Then fc is a homeomorphism from (0, 1) to (c, ∞) (Verify). We know that (see Example
4.38), the space (a, b) is homeomorphic with (0, 1). Hence, it follows that (a, b) is
homeomorphic with (c, ∞).
Next, let gc : (c, ∞) → (−∞, c) be defined by
g(x) = −x, x ∈ (c, ∞).
This is a homeomorphism from (c, ∞) to (−∞, c). Hence, it follows that (a, b) is
homeomorphic with (−∞, c).
Using similar arguments, it can be seen that [a, b) is homeomorphic with [c, ∞) and
to (−∞, c] (Verify). ♢
Example 4.43. (A homeomorphism) Let (−1, 1) and R be with usual topology,
and f : (−1, 1) → R be defined by
x
f (x) = , x ∈ (−1, 1).
1 − x2
This function is a homeomorphism: It can be easily seen that f is continuous (using a
result from Real Analysis). Also, for y ∈ R and x ∈ (−1, 1),
x 2y
f (x) = y ⇐⇒ 2
= y ⇐⇒ x = .
1−x 1 + (1 + 4y 2 )1/2
This shows that f is bijective and the function g : R → (−1, 1) defined by
2y
g(x) = , y∈R
1 + (1 + 4y 2 )1/2
is the inverse of f , and g is continuous (Verify).
We know that (see Example 4.38), the space (a, b) is homeomorphic with (−1, 1).
Hence, it follows that (a, b) is homeomorphic with R. ♢
51
Example 4.44. (A continuous bijection which is not a homeomorphism) Let
X = [0, 1) with usual topology and
Y = S 1 := {(a, b) : a2 + b2 = 1}
Then f is bijective and continuous, but its inverse is not continuous, and hence it is
not a homeomorphism (Verify). ♢
Let X ∈ {(c, ∞), (−∞, d), R} and Y ∈ {[c, ∞), (−∞, d]}. Then
(a, b) ∼ X, [a, b) ∼ Y, X ̸∼ Y .
52
Indeed, since f is bijective, for y ∈ Y ,
Hence, using the fact that f −1 is continuous, it follows that f (A) is open if A is open,
and f (A) is closed if A is closed.
Functions having the properties (1) and (2) in Theorem 4.47 have special names.
Example 4.49. The identity map on a topological space is open, closed, and a home-
omorphism. ♢
Exercise 4.51. Write details of the proofs of the above two theorems.
(Hint: for every A ⊆ X, (f −1 )−1 (A) = f (A). ) ♢
53
Example 4.52. Let X = (0, 1) and Y = R with usual metric topologies and leet
f : X → Y be the inclusion map. That is, f (x) = x, x ∈ X. This is an open map
which is not a closed map:
Any open set in X is also open in Y ; hence it is an open map. Note that, A = (0, 1/2]
is closed in X which is not closed in Y . Hence, f is not a closed map. ♢
Example 4.53. Let X = [0, 1] and Y = R with usual metric topologies and leet
f : X → Y be the inclusion map. That is, f (x) = x, x ∈ X. This is an closed map
which is not an open map:
Any closed set in X is also closed in Y ; hence it is an closed map. Note that,
A = [0, 1/2) is open in X which is not open in Y . Hence, f is not an oepn map. ♢
Example 4.54. Let X = [0, 1) and Y = R with usual metric topologies and let
f : X → Y be the inclusion map. That is, f (x) = x, x ∈ X. This is an neither an
open map nor a closed map: In this case [0, 1) is open and closed in X, but it is neither
open nor closed in Y . Hence, f is neither an open map nor a closed map. ♢
Example 4.55. Let X = (0, 1) and Y = (1, ∞), both with usual topology, and
f : X → Y be defined by f (x) = 1/x, x ∈ X. Then f is a homeomorphism,. ♢
Proof. The proof follows from Theorem 2.48 and Theorem 2.54.
Exercise 4.57. Let R and R2 be with usual topologies. Show that the map f : R → R2
defined by
f (x) = (x, 0), x ∈ R,
is a closed map, but not an open map. ♢
2. The sets A = {(x, y) ∈ R2 , 1 < x2 + y 2 < 2} and B = {(x, y) : 1 < |x| + |y| < 2}
are homeomorphic with respect to the subspace topologies induced by the usual
topology on R2 .
54
3. The sets A = {(x, y) ∈ R2 , x2 + y 2 < 2} and B = {(x, y) : 1 < |x| + |y| < 2} are
not homeomorphic with respect to the subspace topologies induced by the usual
topology on R2 .
Appendix
Consider the Example 4.44. Since t 7→ cos 2πt and t 7→ sin 2πt are continuous on [0, 1),
the map f : t 7→ (cos 2πt, sin 2πt) defined from [0, 1) to R2 also continuous. Hence,
f : X → Y is continuous. Now, for s, t ∈ [0, 1), we have
(cos 2πt, sin 2πt) = (cos 2πs, sin 2πs) ⇐⇒ (cos 2πt − cos 2πs, sin 2πt − sin 2πs) = (0, 0)
⇒
cos 2π(s − t) = cos 2πs cos 2πt + sin 2πs sin 2πt = cos2 2πt + sin2 2πt = 1
⇒
2π(s − t) = 0 ⇒ s = t.
Thus, we have proved that f is one-one.
For (a, b) ∈ S 1 with a ̸= 0,
b 1
(cos 2πt, sin 2πt) = (a, b) ⇒ tan 2πt = ⇒t = tan−1 (b/a).
a 2π
If (a, b) ∈ S 1 with a = 0, we have b = 0 so that f (0) = (0, 1). Thus, we have shown
that f is bijective and continuous.
To see that its inverse is not continuous, consider the open set [0, 1/4) in X. Its
inverse image under f −1 is
55
5.1 Connectedness
Theorem 5.5. A nonempty topological space is connected iff the whole space is its only
nonempty subset which is both open and closed.
56
where E ∩ V1 and E ∩ V2 are nonempty disjoint sets. Since V1 and V2 are open in X0 ,
there exists open sets G1 and G2 in X such that V1 = X0 ∩ G1 and V2 = X0 ∩ G2 .
Hence,
E ∩ V1 = E ∩ X0 ∩ G1 = E ∩ G1 , E ∩ V2 = E ∩ X0 ∩ G2 = E ∩ G2 .
Thus, E ∩ G1 and E ∩ G2 are disjoint nonempty subsets of E such that
E = (E ∩ G1 ) ∪ (E ∩ G2 ),
where G1 and G2 are open in X. This shows that E is disconnected in X.
Conversely, suppose that E disconnected in X. Then there exists open sets G1 and
G2 in X such that
E = (E ∩ G1 ) ∪ (E ∩ G2 ),
where E ∩ G1 and E ∩ G2 are nonempty disjoint sets. Note that
E ∩ G1 = E ∩ X0 ∩ G1 = E ∩ V1 , E ∩ G2 = E ∩ X0 ∩ G2 = E ∩ V2 ,
where V1 = X0 ∩ G1 and V2 = X0 ∩ G2 . Hence,
E = (E ∩ V1 ) ∪ (E ∩ V2 ),
where E ∩ V1 and E ∩ V2 are nonempty disjoint sets with V1 and V2 are open in X0 .
This shows that E is disconnected in X0 .
Remark 5.7. In view of the above theorem, while talking about connected sets in
a topological space, we do not specify explicitly whether it is connected in the whole
space or in a subspace. ♢
Theorem 5.8. Continuous image of a connected set is connected.
57
Example 5.9. Let X = R with usual topology and Y be a discrete space. Then the
only continuous function from X to Y is the constant function: Suppose f : X → Y
is a continuous function. Then f (X) is connected, as X is connected. Since Y is a
discrete space, f (X) has to be a singleton set. Thus f is a constant function. ♢
T If {Aα : α ∈ Λ} S
Theorem 5.11. is a family of connected subsets of a topological space
X such that α∈Λ Aα ̸= ∅, then α∈Λ Aα is connected.
T S
Proof. Let x0 ∈ α∈Λ Aα . Suppose X0 := α∈Λ Aα is not connected. Then there exists
disjoint nonempty open sets V1 and V2 in X0 such that X0 = V1 ∪ V2 . Let G1 and G2
be open sets in X such that V1 = G1 ∩ X0 and V2 = G2 ∩ X0 . Since V1 ∩ V2 = ∅, either
x0 ∈ V1 or x0 ∈ V2 , but not in both. Assume that x0 ∈ V1 .
Now, let α ∈ Λ. Then
We shall answer this question affirmatively by making use of the following proposition.
58
Proof. Given that X = A ∪ B, where A and B are disjoint nonempty open subsets of
X such that A = A ∪ B. Then we have
E = (E ∩ A) ∪ (E ∩ B). (∗)
Note that E ∩ A and E ∩ B are open in E, and they are disjoint as well. Hence, by
the connectedness of E,
E ∩ A = ∅ or E ∩ B = ∅.
If E ∩ A = ∅, then from (∗), we see that E = E ∩ B, so that E ⊆ B. Similarly,
E ∩ B = ∅ implies E ⊆ A.
Theorem 5.14. Let X be a topological space. Then X is connected iff every pair of
points in X is contained in a connected subset of X.
By Theorem 5.12, N
S
n=1 En is connected. Thus, we have shown that there is a connected
set containing {x, y}, completing the proof.
59
Theorem 5.16. (Intermediate value theorem) Let X be a connected topological
space and f : X → R be a continuous function (with usual topology on R). Suppose
x1 , x2 ∈ X be such that f (x1 ) < f (x2 ). Then, for every γ ∈ R such that
Proof. Let γ ∈ R be such that f (x1 ) < γ < f (x2 ). Suppose there is no x ∈ X such
that f (x) = γ. Then we see that
Theorem 5.17. Every convex subset of Rn is connected with respect to the usual
topology on Rn .
Then f is continuous and [0, 1] is connected, its range, namely, the set
60
Proof. Let E be a connected subset of a topological space X. We have to show that
Ē is connected. Assume, on the contrary that Ē is not connected. Let (A, B) be a
separation of Ē. Hence,
E ⊆ Ē = A ∪ B = (Ē ∩ G1 ) ∪ (Ē ∩ G2 )
E = (E ∩ G1 ) ∪ (E ∩ G2 ). (∗)
Note that
Ē ∩ G1 ̸= ∅ ⇒ E ∩ G1 ̸= ∅, Ē ∩ G2 ̸= ∅ ⇒ E ∩ G2 ̸= ∅.
(If x ∈ Ē ∩ G1 , then the open set G1 which contains a closure pint x of E must
contain some point of E.) Thus, (∗) shows that E has a separation, contradicting the
connectedness of E.
B = (B ∩ G1 ) ∪ (B ∩ G2 ),
A = (A ∩ G1 ) ∪ (A ∩ G2 ),
Note that
B ∩ G1 ̸= ∅ ⇒ Ā ∩ G1 ̸= ∅ ⇒A ∩ G1 ̸= ∅.
Similarly,
B ∩ G2 ̸= ∅ ⇒ Ā ∩ G2 ̸= ∅ ⇒A ∩ G2 ̸= ∅.
Hence, (A ∩ G1 , A ∩ G2 ) is s separation of A, contradicting the connectedness of A.
Theorem 5.20. A topological space is disconnected iff there exists a continuous func-
tion from X onto a two-point set with discrete topology.
61
Proof. Let X be a topological space. Suppose X is disconnected. Then there exists
disjoint nonempty open subsets A and B such that X = A ∪ B. Let {0, 1} be with
discrete topology and f : X → {0, 1} be defined by
0, x ∈ A,
f (x) =
1, x ∈ B.
Then f is onto. Since A and B are also closed sets, by pasting lemma, f is continuous.
Conversely, suppose f : X → {0, 1} be a surjective continuous function, where
{0, 1} is endowed with discrete topology. Let
Then A and B are nonempty open sets and X = A ∪ B,. Hence, X is disconnected.
E = (E ∩ G1 ) ∪ (E ∩ G2 ).
Note that E∩G1 and E∩G2 are disjoint open subsets E containing x and y, respectively.
Hence, E is disconnected.
Analogously, it can be shown that every subset of R \ Q containing more than one
elements is disconnected. ♢
For proving that every convex subset of Rn is connected, what we have actually done
was that every pair of points in Rn can be joined by a straight line segment and used
theorem Theorems 5.8 and 5.14. Similar idea can be used prove connectedness of more
general sets in topological spaces, as the following theorem shows.
62
Theorem 5.24. Let X be a topological space such that for every x, y ∈ X, there exists
a continuous function f : [a, b] → X such that f (a) = x and f (b) = y. Then X is
connected.
Proof. Under the given assumption, we have to show that X is connected. By Theorem
5.14, it is enough to show that every pair of points in X is contained in a connected
subset of X. So, let x, y ∈ X. By the assumption on X, there exists a continuous
function f : [a, b] → X for some a, b ∈ R with a < b such that f (a) = x and f (b) = y.
Since [a, b] is connected and f is continuous, the set f ([a, b]) is connected in X with
x = f (a) ∈ f ([a, b]) and y = f (b) ∈ f ([a, b]), and the proof is complete.
Definition 5.25. Let X be a topological space and x, y ∈ X. By a path from x to y,
we mean, a continuous function f : [a, b] → X such that f (a) = x and f (b) = y, and
in that case, we say that x and y are joined by a path. ♢
We observe that:
In view of the above two definitions, Theorem 5.24 can be re-stated as follows.
Theorem 5.27. Every path connected space is connected.
Exercise 5.28. Every convex subset of Rn is path connected with respect to the usual
topology. ♢
63
Then g is continuous and it satisfies
b−a
h(t) = a + (t − c), t ∈ [c, d],
d−c
then g : [c, d] → X defined by
Since f1 and f2 are continuous functions defined on closed subsets of R (with respect
to the usual topology) and f1 (1) = f2 (1) = y, by pasting lemma, g is continuous. Also,
we have g(0) = x and g(2) = z. Thus, g is a path from x to z.
64
Note that Ē = E ∪ F , where F = {(x, y) ∈ R2 : x = 0, −1 ≤ y ≤ 1}. We show
that there is no path from the origin to any point E. For this, assume on the contrary
that there is a continuous function f : [a, b] → Ē such f (a) is the origin and f (b) is a
particular point in E. Consider the the set
J = {t ∈ [a, b] : f (t) ∈ F }.
Since J is closed and bounded, it has a maximum in J, say c = max J. Now, consider
the restriction map f : [c, b] → Ē. Note that f (c) ∈ F and f (b) ∈ E. Let us assume,
without loss of generality, that [c, b] = [0, 1], and write
Then, we have x(0) = 0 and for 0 < t ≤ 1, y(t) = sin(1/x(t)). Our idea is to find a
sequence (tn ) in the open interval (0, 1) such that tn → 0, but (x(tn )) does not converge,
which would contradict the continuity of f at 0. For this, for n ∈ N, let sn such that
0 < sn < x(1/n) and sin(1/sn ) = (−1)n . Since x(·) is continuous, there exists tn such
that
1
0 < tn < and x(tn ) = sn .
n
Thus, we have tn → 0 but (y(tn )) does not converge. ♢
Remark 5.32. The set Ē in Example 5.31 is called the topologist’s sine curve. ♢
Theorem 5.33. Let A and B be path connected subsets of a topological space such that
A ∩ B ̸= ∅. Then A ∪ B is path connected in X.
The following two theorems are also immediate consequences of Theorem 5.33, and
the details of their proofs are left as exercises.
65
Theorem 5.35. Let E0 , E1 , . . . , En be path connected
Sn subsets of a topological space X
such that Ei−1 ∩ Ei ̸= ∅ for i ∈ {1, . . . , n}. Then i=0 Ei is path connected.
We have seen that R \ Q is totally disconnected with respect to the usual topology
(see Example 5.23). What about its counter part in R2 ? However, the set
E = {(x1 , x2 ) ∈ R2 : x1 , x2 ∈ R \ Q}
Thus, we have shown that for every pair of points x, y ∈ E, there exists a path in E
joining x and y. In other words, E is path connected.
6 Compact spaces
Theorem 6.4. Let X be a topological space and E ⊆ X. Then E is a compact set iff
for every family of open sets in X which covers E contains a finite subcollection which
also covers E.
66
Proof. Suppose E is compact. Let {Gα : α ∈ Λ} be a family of open sets in X which
covers E, that is, [
E⊆ Gα .
α∈Λ
Then [ [
E=E∩ Gα = (E ∩ Gα ).
α∈Λ α∈Λ
Thus, E ⊆ ni=1 Gαi . Thus, we have proved that every family of open sets in X which
S
covers E contains a finite subcollection which also covers E.
Conversely, suppose every family of open sets in X which covers E contains a finite
subcollection which also covers E. We have to prove that E is compact. For this, let
{Vα : α ∈ Λ} be a family of open sets in X which covers E, that is,
[
E= Vα .
α∈Λ
Thus, we have proved that every family of open sets in E which covers E contains a
finite subcollection which also covers E, showing that E is compact.
67
• With respect to the usual topology on R, every open interval is non-compact.
Example 6.5. Every co-finite topological space is compact: To see this, let X be a
co-finite topological space and let G be an open cover of X. If X ∈ G, then there is
nothing to be proved. So, suppose that X ̸∈ G. Let G0 ∈ G. Then Gc0 is a finite set,
say Gc0 = {x1 , . . . , n0 } for some n0 ∈ N. Let G1 , . . . , Gn in G be such that xj ∈ Gj for
j = 1, . . . , n0 so that
n0
[
c
G0 = {x1 , . . . , n0 } ⊆ Gj .
j=1
Then we have n0
[
X = G0 ∪ Gc0 ⊆ G0 ∪ Gi .
j=1
68
Theorem 6.8. Closed subset of a compact space is compact.
{G0 } ∪ {G : G ∈ G}
Hence,
n
[ c
Gc0 ∩ Gi = ∅,
i=1
so that n
[
C= Gc0 ⊆ Gi .
i=1
Thus, the open cover G of C contains a finite sub-collection which also covers C,
showing that C is compact.
1. In every indiscrete space with more than one elements, singleton sets are compact,
but not closed.
2. On the set X = {a, b, c}, consider the topology T = {∅, {a}, {a, b, c}}. We
observe that the set {b} is compact, but not closed.
Let T be the family of all subsets A of R2 such that, for every x ∈ A, there exists
r > 0 such that Br (x) ⊆ A. Then, T is a topology on R2 (Verify). With respect
to this topology, singleton sets are compact, but they are not closed.
69
The topological spaces in the above listed examples are not Hausdorff. This raises
the following question:
• Do we have the affirmative answer to the last question if the topological space is
Hausdorff?
x ∈ Uy , y ∈ Vy , Uy ∩ Vy = ∅.
Then, we have
n
\
x ∈ U := Uyi .
i=1
Since Uyi ∩ Vyi = ∅, we have U ∩ Vyi = ∅. This is true for each i = 1, . . . , n. Hence,
U ∩V = ∅. This implies U ∩K = ∅ so that U ⊆ K c . Thus, for each x ∈ K c , we obtain
an open set containing x and contained in K c , showing that K c is an open set.
While proving Theorem 6.9, we have actually proved the following theorem.
Theorem 6.11. Let X be a Hausdorff space, K be a compact set and x ∈ X such that
x ̸∈ K. Then there exist disjoint open sets containing x and K, respectively.
Theorem 6.12. If K1 and K2 are two disjoint compact sets in a Hausdorff space X,
then there exist disjoint open sets containing K1 and K2 , respectively.
Proof. By Theorem 6.11, for each x ∈ K1 , there exist disjoint open sets Ux and Vx such
that
x ∈ Ux , K2 ⊆ Vx .
70
T hen{Ux : x ∈ K1 } is an open cover of K1 . Since K1 is compact, there exist x1 , . . . , xn
in K1 such that
n
[
K1 ⊆ U := Uxi .
i=1
Tn
Let V = i=1 Vxi . Then U and V are disjoint open sets containing K1 and K2 , respec-
tively.
Proof. Let Ω be a metric space and E be a compact subset of Ω. For any given r > 0,
we have [
E⊆ B(x, r).
x∈E
Hence,
d(x, y) ≤ d(x, xk ) + d(xk , xℓ ) + d(xℓ , y) ≤ r + d(xk , xℓ ) + r
Hence, d(x, y) ≤ 2r + max{d(xi , xj ) : i, j ∈ {1, . . . , n}}. Hence, E is bounded.
By Corollary 6.10 and Theorem 6.13, every compact subset of a metric space is
closed and bounded. However, a closed and bounded subset of a metric space need not
be compact. A simple example is the discrete metric space on an infinite set. Here is
another simple example.
Example 6.14. Let Ω = (0, 2] with usual metric. Then the set E = (0, 1] is closed
and bounded, but not compact. ♢
G0 := {G ∩ X0 : G ∈ G}.
71
Then G0 is a family of open sets in X0 that covers E. By hypothesis, there exists
a finite subcollection of G0 that covers E. This finite subcollection is of the form
{G1 ∩ X0 , . . . , Gn ∩ X0 } for some G1 , . . . , Gn in G. Hence,
n
[ n
[
E⊆ (Gi ∩ X0 ) ⊆ Gi .
i=1 i=1
Thus, we have proved that the cover V consisting of open sets in X0 has a finite
subscollection that also covers E. Consequently, E is compact in X0 .
We have seen examples of bijective continuous functions (see, e.g., Example 4.44
and Example 4.46) which are not homeomorphisms. However, if the domain space is
compact and the co-domain space is Hausdorff, then such maps are homeomorphisms.
More precisely, we hae the following theorem.
Theorem 6.16. Every bijective continuous function from a compact space onto a Haus-
dorff space is a homeomorphism.
F (x) = y. (∗)
Now, suppose (yn ) is a sequence in Y such that yn → y for some y ∈ Y . For each
n ∈ N, let xn ∈ X be such that F (xn ) = yn . One would like to know whether xn → x.
72
Theorem 6.16 shows that the answer is in the affirmative if X is compact and Y is
Hausdorff.
If X and Y are metric spaces with metrics d and ρ, respectively, with X compact,
and if F : X → Y is a continuous function, then the above consideration shows that
small error in y can lead only lead to small error in the solution x. In other words, the
solution of the equation (∗) depends continuously on the data y.
Since X is Hausdorff, Kα is closed for every α ∈ Λ, so that the above inclusion shows
that {Kαc : α ∈ Λ0 } is an open cover of the compact set Kα0 . Hence, there exist
α1 , . . . , αN in Λ0 such that
N
[ N
\ c
K α0 ⊆ Kαc i = K αi .
i=1 i=1
Hence,
N
\
K α0 ∩ Kαi = ∅.
i=1
Remark 6.19. We may recall from real analysis that if (Fn ) is a sequence of non-
emptyTclosed sets in a complete metric space X such that Fn ⊇ Fn+1 for all n ∈ N,
then ∞n=1 Fn ̸= ∅. ♢
Let X be a Hausdorff space. We have seen, in Theorem 6.11, that a point in X and
a compact set K which does not contain x can be separated by disjoint open sets, and
73
in Theorem 6.12, every pair of disjoint compact sets in X can be separated by disjoint
open sets. Spaces having these properties with compactness replaced by closedness
are called T3 -spaces and T4 -spaces, respectively. More precisely, we have the following
definitions.
Recall that, in a topological space X, if every pair of distinct points can be separated
by disjoint open sets, then the space is called a Hausdorff space. A topological space
satisfying a slightly weaker notion than Hausdorff condition is the so called T1 -space.
Definition 6.22. A topological space X is said to satisfy T1 -axiom, if for every pair
of distinct points x and y in X, there there exist open sets U and V such that x ∈ U \V
and y ∈ V \ U , and in that case X is called a T1 -space. ♢
T4 ⇒ T3 1 ⇒ T3 ⇒ T2 ⇒ T1 ,
2
74
6.2 Comapct subsets of R
Theorem 6.24. For a, b ∈ R with a < b, the closed interval [a, b] is compact.
β := sup S < ∞.
We first show that β ∈ S. For this, first observe that there exists G ∈ G such that
β ∈ G. Hence, there exists α ∈ [a, β) such that (α, β] ⊆ G. Therefore, by the definition
of supremum, there exists y ∈ S such that
α < y ≤ β.
Since
[a, β] = [a, y] ∪ [y, β],
where y ∈ S and [y, β] ⊆ G, it follows that β ∈ S.
Now, it is enough to show that β = b. To show this, assume the contrary that
β < b. Let G0 ∈ G be such that β ∈ G0 , and let t ∈ [β, b) be such that [β, t] ⊆ G0 .
Since
[a, t] = [a, β] ∪ [β, t],
it follows that t ∈ S. This is a contradiction to the fact that β = sup S.
Proof. Let E be a closed and bounded subset of R. Since E is bounded, there exists
E ⊆ [a, b] for some a, b ∈ R with a < b. Since [a, b] is compact, and since closed subset
of a compact set is compact, we obtain that E is compact.
We already know that every compact subset of a Hausdorff space is closed. Also,
every compact subset of a metric space is bounded. Hence, we obtain the following.
Theorem 6.26. A subset of E of R is compact with respect to the usual topology iff it
is closed and bounded.
Remark 6.27. We shall see that, a subset of E of Rn is compact (with respect to the
usual topology) iff it is closed and bounded. ♢
75
6.3 Locally compact spaces
• A topological space X is locally compact iff for every x ∈ X, there exists open
set V and a compact set K such that
x ∈ V ⊆ K.
3. With respect to the usual topology, the spaces R, Z, N are locally compact, but
not compact.
4. With respect to the usual topology, every nonempty open subset of Rn is locally
compact, but not compact. ♢
Example 6.31. With respect to the usual topology, the spaces Q and R\Q are neither
compact nor locally compact. First let us show that Q is not locally compact. To see
this, let x ∈ Q. Suppose there exists an open set V in Q such that x ∈ V and clQ (V)
is compact in Q. Let G be an open set in R such that V = G ∩ Q, and let a, b ∈ R \ Q
such that a < b and
x ∈ (a, b) ∩ Q ⊆ G ∩ Q ⊆ V.
Note that U := (a, b) ∩ Q is a closed subset of Q and clQ (U) ⊆ clQ (V). Hence,
clQ (U) = U is compact in Q. Hence, U is compact in R as well, which is not true, as
compact subset of R are closed and bounded in R, but U is not closed in R. ♢
The conclusion in the above example also follows from the following general result.
76
Theorem 6.32. Let X be a Hausdorff space and X0 be a locally compact dense subspace
of X. Then X0 is open in X.
Proof. Let x ∈ X0 . We show that there exists an open subset of X containing x and
contained in X0 , that will complete the proof.
Since X0 is locally compact, there exists a compact set K ⊆ X0 and an open set V
in X0 such that
x ∈ V ⊆ K.
Since V is open in X0 , there exists an open set G in X such that V = G ∩ X0 . We
show that G ⊆ X0 . In fact, we show that G ⊆ G ∩ X0 so that
G ⊆ G ∩ X0 = V̄ ⊆ K̄ = K,
where the last equality follows, since K is a compact subset of a Hausdorff space. For
showing G ⊆ G ∩ X0 , let y ∈ G and let H be an open subset of X containing y. Since
H ∩ G is open set containing y and X0 is denset in X, we have H ∩ G ∩ X0 ̸= ∅.
That is, the open neighbourhood H of y contains some point from G ∩ X0 . Hence,
y ∈ G ∩ X0 .
• The dense subspaces Q and R \ Q of R (with respect to the usual topology) are
not open in R, and hence are not locally compact.
We know that the topological space X = (0, 1] with usual metric is a locally compact
Hausdorff space which is not compact. Now, if we add one point to it, we obtain a
compact space Y = [0, 1] with the usual topology. Note that X is dense in Y .
Given a locally compact Hausdorff space X which is not a compact space, can we
obtain a compact space in which X is a dense subspace?
For example, consider the space X = (0, 1), which is a locally compact Hausdorff
space, which is not compact, with respect to the with usual topology. Let us consider
the set
Y = X ∪ {p}, p := {0, 1}.
Consider the topology on Y for which the base consists of all open sub-intervals of
(0, 1) together with sets of the form Y \ [a, b] for a, b ∈ (0, 1). Then, it can be seen
that, with this topology, Y is a compact Hausdorff space. Also, since every set of the
form Y \ [a, b] for a, b ∈ (0, 1) contains points from X, X is a dense subspace of Y .
77
The above method can be adopted to obtain a compact Hausdorff space out of a
ny locally compact Hausudorff space, as described in the following theorem.
Theorem 6.33. (One-point compactification) Let (X, TX ) be a locally compact
Hausdorff space which is not compact. Then there exists a compact Hausdorff space
Y ⊇ X such that Y \ X is a singleton set, X is a subspace of Y and X̄ = Y .
Let Y = X ∪ {p}, where p ̸∈ X, and let TY be the family of all those subsets A of
Y such that either A is open in X or A = Y \ K for some compact subset K ⊆ X.
Note that for a compact set K ⊆ X,
A = Y \ K ⇐⇒ A = {p} ∪ (X \ K).
7 Product topology
X = X 1 × · · · × Xn .
V := {V1 × · · · × Vn : Vi open in Ti , i = 1, . . . , n}
V := {V1 × · · · × Vn : Vi open in Ti , i = 1, . . . , n}
78
The proof of the following theorem is obvious from the definition of a product topology.
πi (x) = xi , x := (x1 , . . . , xn ) ∈ X,
is continuous and surjective, and the product topology on X is the weakest topology
such that π1 , . . . πn are continuous.
πi (x1 , . . . , xn ) = y.
πi−1 (G) = V1 × · · · × Vn ,
Definition 7.3. The maps πi in the above theorem are called projection maps. ♢
Example 7.4. Let Xi =Q R for i = 1, . . . , n with usual topology. Then the product
topological space X := ni=1 Xi is nothing n
Qn but R with usual topology. For this
topology, the set of all sets of the from i=1 (ai , bi ) forms a base. ♢
Proof. It is enough (why?) to show that product of two connected spaces is connected.
Let X and Y be connected topological spaces. We show that X × Y is connected.
For this, let us fix an element (x0 , y0 ) ∈ X × Y , and consider the subspaces X × {y0 }
and and for each x ∈ X, the subspace {x} × Y . Note that
S
1. X × Y = x∈X Zx , Zx := (X × {y0 }) ∪ ({x} × Y ),
79
3. (x0 , y0 ) ∈ Zx for every x ∈ X.
Now, since X × {y0 } and {x} × Y are homeomorphic with the connected spaces X
and Y , respectively, and (x, y0 ) is a common point in these spaces, the subspace Zx is
conencted for each x ∈ X. Also, since (x0 , y0 ) ∈ Zx for each x ∈ X, their union X × Y
is connected.
For its proof we shall make use of the following lemma, known as tube-lemma7 .
Proof of Theorem 7.6. It it enough to prove for the case of n = 2. So, let X and Y
be compact topological spaces, and X × Y be with product topology. Let G be an open
cover of X × Y . For each x ∈ X, since {x} × Y is compact, there exists G1 , . . . , Gk in
G such that
[k
{x} × Y ⊆ Gi .
i=1
Hence
ℓ
[
X ×Y = (Vxi × Y ),
i=1
80
7.2 On product of infinite number of topological spaces
Next suppose we have a countably infinite collection of topological spaces, say (Xi , Ti )
for i ∈ N. Consider the cartesian product of X1 , X2 , . . ., that is, the set
∞
Y
X := Xi .
i=1
Q∞
• i=1 Xi is the set of all sequences of the form (x1 , x2 , . . .) with xn ∈ Xn for n ∈ N.
Q∞
For each n ∈ N, let πn : i=1 Xi → Xn defined by
Let
S := {πn−1 (G) : G ∈ Tn , n ∈ N}.
Q∞
• The topology on i=1 Xi generated by S is the weakest topology such that each
πn is continuous.
Q∞
Definition 7.8. The topology on i=1 Xi generated by
S := {πn−1 (G) : G ∈ Tn , n ∈ N}
Definition 7.9. Let Λ be a non-empty set and for S each α ∈ Λ, let Xα be a non-empty
set. Then the collection of all functions f : Λ → α∈Λ Xα such that f (α) ∈ Xα for each
α ∈ Λ is called
Q the cartesan productQof Xα , α ∈ Λ, and this cartesian peroduct is
denoted by α∈Λ Xα . An element f ∈ ∞ α∈Λ Vα is usually denoted by (xα ) or (xα )α∈Λ
with the understanding that xα = f (α) for α ∈ Λ. ♢
81
Definition 7.10.
Q Let {(Xα , Tα ) : α ∈ Λ} be a family of topological spaces. Then the
topology on α∈Λ Xα generated by
S := {πβ−1 (G) : G ∈ Tβ , β ∈ Λ}
Q
is called the product topology on α∈Λ Xα . ♢
Q
We may define another topology on α∈Λ Xα which seems more natural than the
product topology, the so called box topology.
Suppose that, for each α ∈ Λ,Q(Xα , Tα ) is a topological space, and let U be the
collectionQof all sets of the form ∞ α∈Λ Vα , where each Vα is open in Xα . Then U
∞
contains α∈Λ Xα andQit is closed under finite intersections, and hence U is a base for
a unique topology on α∈Λ Xα .
Q
Definition 7.13. The topology on α∈Λ Xα for which
nY o
U := Vα : Vα ∈ Tα
α∈Λ
Q
is a base is called the box topology on α∈Λ Xα . ♢
• If Λ is a finite set, then product topology and box topology on the product set is
the same.
82
References
[1] J. R. Munkres, Topology, 2nd Edition, Pearson Edu. 3rd. Impression, 2007.
[2] K.D. Joshy, Introduction to General Toplogy (Revised), new Age International,
new Delhi, 1999.
83