Topology
Topology
Topological Spaces
Chapter Structure
1.1 Introduction
1.2 Objectives
1.3 Topological Spaces
1.3.1 Base
1.4 Product Topology and Subspace Topology
1.5 Open and Closed Sets
1.6 Interior
1.7 Continuous Functions
1.8 Glossary
1.9 Bibliography
1.10 Let Us Sum Up
1.11 References for Further Reading
1.12 Chapter End Exercises
1.1 Introduction
Metric spaces are generalization of real line and the usual distance.
We take the process of generalization one more step ahead when we
eliminate the concept of distance and work with the collection of open
sets. In the chapter on metric spaces we saw several concepts like
closure, interior, boundary, continuity etc. which involve concept of
distance. These concepts can be described equivalently using only open
sets and closed sets. Thus a lot of analysis can be carried out only
knowing the open sets (and the complements of open sets) without
using the notion of distance.
1
1.2 Objectives
After going through this chapter you will know:
• Definition and examples of a topological space.
• Definition and examples of basis and subbasis of a topological space.
• Definition and properties of closure of a set.
• Definition and properties of interior of a set.
• Properties of continuous functions.
• ∅ ∈ τ and X ∈ τ.
2
in (X, τ ) are the empty set and X itself. Hence, we cannot get sets U1
and U2 as claimed above.
Examples:
3
1.3.1 Base
Definition 2. Let (X, τ ) be a topological space. A base to the topology
τ is a subfamily B of τ having the following property: given any G ∈ τ
there exists a family {Bλ : λ ∈ Λ} ⊆ B such that G = ∪λ∈Λ Bλ .
(i) X = ∪B∈B B.
4
We do not use subbases extensively in our further discussion. We
move on to our next important concept: the product topology. This is
one of the ways to build new topological spaces out of old ones.
5
Definition 5. Let (X, τ ) be a topological space. Let x ∈ X. A subset
N of X is said to be a neighborhood of the point x if there exists
G ∈ τ such that x ∈ G ⊆ N. Note that in general, N need not belong
to τ. It is called open neighborhood if it belongs to τ. Collection of
all neighborhoods of x is denoted by N (x). It is easy to see that a
nonempty subset G of X is open if and only if it is a neighborhood of
each of its point. Let N x be an open set containing x and contained in
G. Then G = ∪x∈G N x and G is open, being an union of open sets. On
the other hand if G is open then G itself is the neighborhood of each
of its point contained in G.
Examples:
(ii) If A ⊆ B then A0 ⊆ B 0 .
(iii) (A ∪ B)0 = A0 ∪ B 0 .
Proof. (i) For any neighborhood N of any point x of X we have
∅ ∩ N = ∅ and hence ∅ ∩ N \ {x} = ∅. This proves (i).
6
(iii) We have A ⊆ A ∪ B which give A0 ⊆ (A ∪ B)0 Similarly B ⊆
A ∪ B; which gives B 0 ⊆ (A ∪ B)0 . These two together give us
A0 ∪ B 0 ⊆ (A ∪ B)0 .
To get the reverse inclusion suppose x ∈ (A ∪ B)0 but does not
belong to A0 and B 0 . Then there exist neighborhoods N1 and N2
of x with A∩N1 \{x} = ∅ and B ∩N2 \{x} = ∅. Let N = N1 ∩N2 .
Then N is a neighborhood of x satisfying A ∩ N \ {x} = ∅ and
B ∩ N \ {x} = ∅ which given (A ∪ B) ∩ N \ {x} = (A ∩ N \ {x}) ∪
(B ∩ N \ {x}) = ∅ ∪ ∅ = ∅. This shows that x does not belong
to (A ∪ B)0 This proves (A ∪ B)0 ⊆ A0 ∪ B 0 which completes the
proof of (iii).
7
• Let {Cλ : λ ∈ Λ} be a family of closed sets. Then for each
λ ∈ Λ the set Gλ = X \ Cλ is an open set in (X, τ ). Hence
∪{Gλ : λ ∈ Λ} ∈ T. But ∪{Gλ : λ ∈ Λ} = X \ ∩{Gλ : λ ∈ Λ}.
This shows that ∩{Cλ : λ ∈ Λ} is closed.
Proof. Let A be the smallest closed set containing A and let c(A) =
A ∪ A0 . We will prove:
(i) A ⊆ c(A),
(ii) c(A) ⊆ A.
1.6 Interior
This concept from metric spaces is extended to topological spaced
as follows.
Let A be a non empty set in (X, τ ). We define Interior of A as the
union of all open subsets of A. We denote it by i(A) or A◦ . i(A) is an
open set by definition. i(A) is in fact the largest open set contained in
A.
8
Theorem 1.6.1. (1.) i(X) = X
(2.) i(A) ⊆ A, for all subsets A of X.
(3.) i(i(A)) = i(A), for all subsets A of X.
(4.) If A ⊆ B then i(A) ⊆ i(B).
(5.) i(A ∩ B) = i(A) ∩ i(B), for all subsets A and B of X.
Proof. (1.) X is open. Hence X is the largest open set contained in
X. This proves (1).
(2.) i(A) is a union of subsets of A and hence is a subset of A. This
proves (2).
(3.) i(A) is open. Hence i(i(A)) = i(A).
(4.) i(A) is an open set contained in A. Hence i(A) is an open set
contained in B. This means i(A) is contained in largest open set
contained in B which is i(B).
(5.) A ∩ B ⊆ A which gives i(A ∩ B) ⊆ i(A). Also A ∩ B ⊆ B which
gives i(A∩B) ⊆ i(B). Together these two imply i(A∩B) ⊆ i(A)∩
i(B). Now let y ∈ i(A) ∩ i(B). By definition of interior there exist
open sets G and H in (X, τ ) satisfying y ∈ G ⊆ A and y ∈ H ⊆ B.
G ∩ H is an open subset of A ∩ B hence G ∩ H ⊆ i(A ∩ B). This
proves y ∈ i(A ∩ B) and hence i(A) ∩ i(B) ⊆ i(A ∩ B). This
completes the proof.
Definition 9. For any set A in the topological space the set c(A) \ i(A)
is defined as the boundary of A denoted by b(A). It is also called frontier
of A.
Check Your Progress Show that a point x of X is a boundary
point of a subset A of X if and only if N ∩ A 6= ∅ and N ∩ (X \ A) 6= ∅
for all N ∈ N (x).
9
Remark 1.7.1. f (N ) ⊆ N e implies N ⊆ f −1 (Ne ) and therefore f −1 (N
e)
becomes a neighborhood of a. Therefore the definition can be restated
as: f is continuous at a if for every neighborhood N e of f (a) the set
f −1 (N
e ) is a neighborhood of a. Since neighborhoods are supersets open
sets containing that point continuity at a point can be rephrased as :
f is continuous at a if given open set H of (Y, τ 0 ) with f (a) ∈ H there
exists open subset G of (X, τ ) containing a such that f (G) ⊆ H.
10
(c) =⇒ (d). Let A be any subset of X. Let C = c(f (A)). C is a
closed subset of (Y, τ 0 ). Therefore by (c), f −1 (C) is a closed subset of
(X, τ ). But we have A ⊆ f −1 (f (A)) ⊆ f −1 (c(f (A))) = f −1 (C).
Now A is a subset of the closed set f −1 (C). Hence c(A) ⊆ f −1 (C)
which gives f (c(A)) ⊆ f (f −1 (C)) ⊆ C. Thus f (c(A)) ⊆ c(f (A)) is
proved.
(d) =⇒ (c). Let C be a closed subset of (Y, τ 0 ). Let A = f −1 (C).
We have f (c(A)) ⊆ c(f (A)) ⊆ c(f (f − 1(C)) ⊆ c(C) = C. Now apply
f −1 to f (c(A)) ⊆ C we get f −1 (f (c(A))) ⊆ f −1 (C) = A. But c(A) ⊆
f −1 (f (c(A))). Therefore we get c(A) ⊆ A. That is, A is a closed subset
of (X, τ ).
(c) =⇒ (a). Let x be any point of X and let H be any open subset
of (Y, τ 0 ) containing f (x). Let C = Y \ H then we have f −1 (H) =
X \ f −1 (Y \ H) = X \ f −1 (C). By the result (c), f −1 (C) is closed and
hence f −1 (H) = X \ f −1 (C) is open. This proves that f is continuous
at every point x in X.
1.8 Glossary
• Interior of a set.
11
1.9 Bibliography
12
is continuous at a point a of X if for any neighborhood N e of f (a) in
(Y, τ ) there corresponds a neighborhood N of a in (X, τ ) such that
f (N ) ⊆ N e . We say that f is continuous on a subset A of (X, τ ) if it
is continuous at every point a of A. Continuity of f on X has several
equivalent forms. Some are listed below.
3. Let I be the set of all bounded open intervals and let I ∗ be the
subfamily of I consisting of all open intervals of I having rational
end points. Prove that the topology on R generated by I ∗ is the
same as that generated by I. (Note that the topology generated
by I is the usual topology of R.)
13
5. Let B be a base to a topological structure τ on X and let Y be a
nonempty subset of X. Let By = {B ∩ Y : B ∈ B}. Is By a base
to the subspace topology τy on Y? Justify.
14
Chapter 2
Chapter Structure
2.1 Introduction
2.2 Objectives
2.3 Homeomorphism and Heredity
2.4 Cardinality
2.5 Separation Axioms
1.6 Hausdroff topological Spaces
2.7 Regular and Normal Topological Spaces
1.8 Glossary
2.9 Bibliography
2.10 Let Us Sum Up
2.11 References for Further Reading
2.12 Chapter End Exercises
2.1 Introduction
15
2.2 Objectives
After going through this chapter you will know:
• Countability axioms and their consequences
• Separation axioms and examples of metric spaces that satisfy them.
• Definition, examples and properties of Hausdorff spaces.
• Regular and normal topological spaces and their properties.
2.4 Cardinality
Though cardinality is not our main topic of discussion we need to
revise some results in cardinality to understand the countability axioms.
Cardinality of N, the set of natural numbers and any set which is in
bijection with N is denoted by ℵ0 . These sets are called denumerable.
16
It can be proved that the set of rational numbers is denumerable. A
set which is either denumerable or finite is called as countable. The set
of real numbers R and any interval in R are uncountable sets. It can
be proved that cardinality of R, R2 and any interval in R are same. We
denote it by C. Continuum hypothesis states that there is no infinite
set having cardinality between ℵ0 and C. An important property of
countable sets is that all its elements can be written in the form of a
sequence.
17
• Let (X, d) be any metric space. Then B(x) = {B(x, 1/n) : n ∈
N} is a countable base to the complete neighborhood system of
each x in X.
Definition 14. We say that the topological space (X, τ ) is first count-
able space if for each x in X the complete neighborhood system N (x)
has a countable base. This countable base is called the fundamental
system of neighborhoods of x. We also say that such a space is a C1 -
space.
Topological spaces discussed in examples 2 and 3 discussed above
are C1 - spaces. In addition the discrete space (X, P (X)) is C1 be-
cause {{x}} is a fundamental system of neighborhoods for any x in
X and the indiscrete space (X, {∅, X}) is C1 - spaces because X is
the only neighborhood of every point x in X. If {Nn : n ∈ N } is
a fundamental system of neighborhoods of a point x then putting
f1 = N1 ∩ N2 ∩ N3 · · · we get that {N
N fn : n ∈ N } is another count-
able neighborhood base of x in which the neighborhoods decrease with
increasing n : N f1 ⊃ Nf2 ⊃ N f3 ⊃ N fn ⊃ · · · Thus in a C1 - space we
can always choose a countable, monotonically decreasing fundamental
neighborhood system of neighborhoods of each of its points. Now we
verify that first countability is a topological property.
Let f : (X, τ ) → (Y, τ 0 ) be a homeomorphism. Let (X, τ ) be a
C1 -space. We verify that each y ∈ Y has a countable fundamental
neighborhood system. Let x ∈ X satisfy f (x) = y. If Mn = f (Nn )
then for each n, Mn are neighborhoods of y. In fact {Mn : n ∈ N} is a
fundamental neighborhood system of y. This proves that (Y, τ 0 ) is C1 -
space. Thus homeomorphic image of a C1 -space (X, τ ) is a C1 - space.
We will verify that first countability is also a hereditary property.
Let (X, τ ) be a C1 -space and let Y be a non-empty subset of X. Let
y ∈ Y. Then y ∈ X. By the C1 - property of (X, τ ), y has a fundamental
neighborhood system {Nn : n ∈ N} in (X, τ ). Then {Nn∗ := Nn ∩Y : n ∈
N} is a fundamental neighborhood system of y in (Y, τy ). This proves
that the subspace (Y, τy ) is also C1 - space. Hence it is a hereditary
property.
Definition 15. A topological space (X, τ ) is second countable (or sat-
isfies the second countability axiom) if the topology has a countable
base.
A second countable space is first countable, for if B is a countable
basis for (X, τ ), then B(x) = {B ∈ B for which x ∈ B} ⊆ B and
hence is countable. Thus, B(x) is a countable base to the complete
neighborhood system N (x) and hence C1 - axiom is satisfied.
The converse is not true. A space may be first countable without
being second countable. Consider the discrete topological structure
on R. This topological space (R, P (R)) has no countable base. But
18
it is C1 - space because for every point x ∈ X the singleton family
{{x}} is a fundamental neighborhood system of x. This shows that
second countability is stronger than first countability. Some examples
of second countable spaces are given below.
• Let X = {a, b, c}. Then τ = {∅, {a, b}, X} is not T0 because the
condition fails for the pair a, b.
19
By the T0 property of (X, τ ) there exists an open set G of (X, τ )
which contains one of the points x1 , x2 avoiding the other. With-
out loss of generality suppose x1 ∈ G and x2 6∈ G. Let f (G) = H.
Then by defining property of the homeomorphism H ∈ τ 0 . y1 ∈ H
but y2 6∈ H. This proves (Y, τ 0 ) is a T0 space.
20
• Discrete topological space (X, P (X)) and metric spaces are ex-
amples of T1 spaces.
• On X = {a, b, c} we define τ = {∅, {a}, {a, b}, {a, c}, X}. (X, τ )
is T0 but not T1 . The pair a, b does not satisfy the requirements
of a T1 space.
21
2.6 Hausdorff Topological Spaces
22
space. Thus if the set X has more than one point then each sequence
has more than one limit. Another such example is X = {a, b, c} and
τ = {∅, {c}, {a, b}, X}. Now define the sequence {an : n ∈ N} by a2k =
a and a2k+1 = b for all k ∈ N. This sequence converges to both b and c.
Following is an example of a sequence that converges to infinitely
many points.
Let X = N. Let Jn = {n, n + 1, n + 2, . . .} and τ = {∅, J1 =
N, J2 , J3 , . . .}. Define the sequence (an ) by an = n for every n. The
sequence converges to every m ∈ N.
We cannot find such examples in Hausdorff spaces because of the
following
The real line is a separable metric space because the set of rational
numbers is a countable dense subset. But R with the co-countable
topology is not a separable topological space. Because if you take a
countable set A then R \ A is a co-countable set. This means R \ A is
open and A is a closed set which gives c(A) = A is a proper subset of
R as R is uncountable.
23
Remark 2.6.2. Being separable is not hereditary.
24
Proof. • Suppose that X is regular and suppose that the point x
and the neighborhood U of x are given. Let B = X \ U ; then
B is a closed set. By hypothesis, there exist disjoint open sets V
and W containing x and B respectively. The set c(V ) is disjoint
from B since if y ∈ B the set W is a neighborhood of y disjoint
from V. Therefore c(V ) ⊆ U as desired.
To prove the converse, suppose the point x and the closed set B
not containing x are given. Let U = X \B. By hypothesis there is
a neighborhood V of x such that c(V ) ⊆ U. The open sets V and
X \ c(V ) are disjoint open sets containing x and B respectively.
Thus X is regular.
2.8 Glossary
In this chapter, you have learnt the following:
2.9 Bibliography
25
2.10 Let Us Sum Up
A topological system is a generalization of the classical number sys-
tem R. But in such a wild generalization many important properties of
the number system are lost. We regain these properties by introducing
a number of additional axioms. There are many types of such axioms.
One such type of axioms are separation axioms. They are denoted by
T0 , T1 , T2 , . . . . They are about enclosing a pair of points in disjoint open
subsets, disjoint closed subsets etc. Of these we study T0 , T1 and T2 .
These properties are both topological and hereditary.
A T0 space separates a pair of points from the space by a single
open set. A space (X, τ ) is T0 if and only if distinct singleton sets have
distinct closures. A space is T1 if given any pair of distinct points there
are open sets each one of them containing one point but not the other.
A topological space is T1 if and only if singleton sets are closed. In a
T1 space if p is a limit point of a set A then every neighborhood of p
contains infinitely many points of A.
A T2 space is also called a Hausdorff space. It separates distinct
points by disjoint open sets. In a Hausdorff space a convergent sequence
has unique limit. Regular and normal spaces satisfy stronger separation
axioms. Countability axioms are other type of auxiliary axioms. We
study first and second countability axioms. Both these axioms are
topological and hereditary. A metric space is called separable if it
has a countable dense subset. A second countable topological space is
separable.
26
1. Show that being T1 is a topological and hereditary property.
10. Let X be the set of all irrational numbers with usual metric d. Is
(X, d) separable?
27
28
Chapter 3
Compactness
Chapter Structure
3.1 Introduction
3.2 Objectives
3.3 General Definition of Compactness
3.4 Continuity and Compactness
3.5 Finite Products and Compactsness
3.6 Local Compactness and one-point Compactification
3.7 Lindelof Topological Spaces
3.8 Glossary
3.9 Bibliography
3.10 Let Us Sum Up
3.11 References for Further Reading
3.12 Chapter End Exercises
3.1 Introduction
Compactness is a word we use daily in our life, to indicate that
objects occupy less space. Even in mathematics the word has the same
sense in Rn namely that compact objects are those which are closed
and bounded. However, in an arbitrary topological space such a nice
formulation may fail, as we shall see from examples.
In an arbitrary topological space the correct notion of compactness
is defined in terms of open covers. Using this defintion, we prove that
compactness is preserved under continuous functions and under finite
products. We then move on to the notion of local compactness and also
study spaces which are not too far away from being compact: namely
spaces which admit a one-point compactification.
We also study a notion similar to compactness: namely the notion
of a Lindelöf space.
29
3.2 Objectives
After going through this chapter you will know:
• General definition of compactness.
• Continuity preserves compactness.
• A finite product of compact spaces is compact.
• Definition of local compactness.
• Construction of one point compactification.
• Lindelöf Spaces.
Definition 22. (Open cover) Let X be a topological space and {Uα }α∈Λ
be a family of open subsets of X. Then {Uα }α∈Λ is said to be an open
cover of X if X ⊂ ∪α∈Λ Uα . (Note that this condition is same as saying
X = ∪α∈Λ Uα , as Uα ⊂ X, for all α ∈ Λ.)
30
From the above definition it is easy to check that {Uα }α∈Λ is an
open cover of X if and only if X = ∪α∈Λ Uα , for open subsets Uα of X.
Let us see some examples of open covers:
1. Let X = {1, 2, 3} with the discrete topology. Then {1}, {2, 3} is
an open cover of X.
31
Example 1. Every finite set is a compact set. (Since any open cover
of this finite set will require atmost finitely many open sets to cover it.)
2. Show that every subset of the real line R with the finite comple-
ment topology is compact.
(b) For the other way, start with an open covering of Y, say {Vα,Y }α∈Λ .
Hence there exists a family of open sets {Vα,X }α∈Λ of X such that
Vα,Y = Vα,X ∩ Y, i.e., Vα,Y ⊂ Vα,X . Then clearly we get:
32
Thus, we get a covering of Y by open subsets of X. The hypoth-
esis now gives a finite subcollection covering Y, i.e. there exist
α1 , . . . , αn such that Y ⊂ Vα1 ,X ∪ Vα2 ,X · · · ∪ Vαn ,X . Taking inter-
section of both sides with Y, we get that
33
Vy1 , . . . , Vyn . Hence the open set V := Vy1 ∪ · · · ∪ Vyn contains C. Let
U := Uy1 ∩ · · · ∩ Uyn . Clearly U being a finite intersection of open sets
is open in X. Also, U ∩ V = ∅. (For if there exists r ∈ U ∩ V, then
r ∈ Vyi for some 1 ≤ i ≤ n and r ∈ Uyi for all 1 ≤ i ≤ n implies that
r ∈ Uyi ∩ Vyi , which is not possible as Uyi ∩ Vyi = ∅.)
Check Your Progress
1. Show from first principles (i.e., using the definition of compact-
ness) that R with the usual topology is not compact. (Hint: It is
enough to produce an open cover which does not admit a finite
subcover.)
2. Show that closed subsets of compact sets are compact. (Hint:
Imitate the proof in Lemma 3.4.3.)
34
Lemma 3.5.1. Let X, Y be topological spaces. Then the projection
maps p1 : X × Y → X and p2 : X × Y → Y are continuous in the
product topology. Let x0 ∈ X. Then the slice x0 × Y is homeomorphic
to Y. In particular, if Y is compact, then the slice x0 × Y is a compact
subset of X × Y, where X × Y is given the product topology.
35
Remark 3.5.1. One has a more general theorem which says that an
arbitrary product of compact spaces is compact too. This theorem is
known as Tychonoff’s theorem. The proof of this is beyond the scope
of these notes, but it is a very useful theorem.
We use the Tube lemma now to prove that the product of two
compact spaces is again compact. The result for finitely many compact
spaces then follows by induction.
Theorem 3.5.1. Let X, Y be compact topological spaces. Then their
Cartesian product X × Y (with the product topology) is also compact.
Proof. Let R be an open cover of X × Y. Given x0 ∈ X, the slice x0 × Y
is compact by Lemma 3.5.1. Hence, there exist finitely many open
subsets, say U1 , . . . , Un of R which cover x0 ×Y. Then, N := U1 ∪· · ·∪Un
is an open set containing the slice x0 ×Y. Hence, by Tube Lemma, there
exists a tube say Wx0 ×Y around x0 ×Y, contained in N. Clearly, Wx0 ×Y
has a finite subcover, as Wx0 × Y ⊂ N = U1 ∪ · · · ∪ Un .
Now for each x ∈ X, repeat the procedure above to get a tube
Wx × Y containing the slice x × Y. Now observe that {Wx |x ∈ X} is
an open cover of X, as Wx are open in X by construction. As X is
compact, there exist x1 , . . . , xn ∈ X such that X ⊂ Wx1 ∪ · · · ∪ Wxn .
Hence,
X × Y ⊂ (Wx1 × Y ) ∪ · · · ∪ (Wxn × Y ).
As each of Wxi × Y has a finite cover, so does X × Y. Hence, X × Y is
compact.
Check Your Progress
1. Let f be a continuous real-valued function on [a, b]. Prove that
the graph of f , i.e., the set {(x, f (x)) | x ∈ [a, b]} is a compact
subset of R2 .
2. State true or false with correct justification (if false, give a counter-
example): An arbitrary union of compact sets is compact.
36
3.6 Local compactness and one-point
compactification
There is another more useful notion of compactness called local
compactness. This is a weaker notion than compactness. We shall see
examples of spaces which are locally compact but not compact.
37
compact space! Thus by adding one point suitably you are sometimes
able to get a compact space out of your old space. Such a process is
called as one point compactification.
(a) U , with U open in X. (We call these open sets of Type I.)
Proposition 3.6.1. Let Y be the set defined above with the open sets
of the two types. Then, Y is a topological space.
38
1(c) Let U1 , U2 be both open sets of Type II i.e, there exist compact
sets C1 , C2 ∈ X such that U1 = Y \ C1 and U2 = Y \ C2 . Hence,
U1 ∪ U2 = Y \ (C1 ∩ C2 ). As a finite intersection of compact sets
is again compact, we observe that U1 ∪ U2 is an open set of Type
II.
1(a) Let {Uα }α∈Λ be a family of open sets of Type I. Then clearly,
∪α∈Λ Uα is again an open set of Type I.
1(b) Let {Uα }α∈Λ be a family of open sets of Type I and {Uβ }β∈Λ0 be
a family of open subsets of Type II.
We will prove that (∪α∈Λ Uα )∪(∪β∈Λ0 Uβ ) is again an open set in Y.
For this observes that (∪α∈Λ Uα ) ∪ (∪β∈Λ0 Uβ ) equals U ∪ (Y \ C) =
Y \ (C \ (U ∩ C)). Now note that U ∩ C is open in C and hence,
C \ (U ∩ C) is closed in C and hence compact, as C is compact.
Thus, (∪α∈Λ Uα ) ∪ (∪β∈Λ0 Uβ ) is a set of Type II and hence open.
1(c) Let {Uβ }β∈Λ0 be a family of open sets of Type II. Then ∪β∈Λ0 Uβ
is again an open set of Type II as this equals ∪β∈Λ0 Y \ Cβ , for
compact subsets Cβ of X. This union now equals Y \ ∩β∈Λ0 Cβ .
Now note each Cβ is a compact subset of the Hausdorff space X
and hence is closed in X. Hence, ∩β∈Λ0 Cβ is a closed subset of X.
Moreover, for a fixed β0 ∈ Λ, ∩β∈Λ0 Cβ ⊂ Cβ0 and hence we have
a closed subset of a compact set and hence ∩β∈Λ0 Cβ is a compact
subset of X. This implies that we get that ∪β∈Λ0 Uβ is an open set
of Type II.
39
Example 4. One-point compactification of (0, 2π) is homeomorphic
with the circle S 1 .
Define a map f : (0, 2π) ∪ {∞} → S 1 by: f (θ) = (cos(θ), sin(θ)),
for all x ∈ (0, 2π) and f (∞) = (1, 0). Clearly, f is a bijection.
We now prove that f is continuous. Let V ⊂ S 1 be open. Then,
either (1, 0) 6∈ V or (1, 0) ∈ V. If (1, 0) 6∈ V, then V is an arc of S 1
not containing (1, 0) and hence f −1 (V ) is of the form (θ1 , θ2 ) for some
θ1 < θ2 ∈ (0, 2π). In this f −1 (V ) is an open subset of (0, 1).
If (1, 0) ∈ V, then there exists an arc around (1, 0) contained in V.
Thus, this arc consists of some points in the first quadrant and some
points in the fourth quadrant, which are determined by angles θ1 < θ2
with θ1 > 0 and θ2 < 2π. Thus, in this case f −1 (V ) = {∞} ∪ (0, θ1 ) ∪
(θ2 , 2π). Clearly, then f −1 (V ) = Y \ C, where C is the compact subset
[θ1 , θ2 ] of (0, 1). This finishes the proof that f is continuous.
Let Y = (0, 2π) ∪ {∞}. We will now prove that the direct image of
an open subset of Y under f is again open. Similar to the case above,
it is easy to check that if U is an open subset of (0, 2π), then f (U ) is
open in S 1 . Let U be an open subset of (0, 2π) ∪ {∞} containing ∞.
Then, U is of the form Y \ C, for some compact subset C of X, then
f (U ) = f (Y ) \ f (C). As C is compact and we have proved that f is
continuous, we get that f (C) is compact and hence closed in S 1 . Hence,
f (U ) = S 1 \ f (C) is an open subset of S 1 , proving that f is an open
map. This completes the proof that f is a homeomorphism.
40
of X. As X is not compact, C is a proper subset of X and hence the
complement X \ C contains at least one point. This then implies that
every open subset around ∞ intersects X in a point different from ∞,
proving that ∞ is a limit point of X.
Clearly, Y \ X is the single point infinity and it now remains to
prove that Y is compact and Hausdorff. First we prove compactness
of Y . Let R be an open covering of Y. As ∞ should belong to this
collection, R contains an open subset of Type II, say of the form Y \ C,
with C compact in X. Now take all members of R other than Y \ C
and intersect them with X. They form a collection of open sets in X
covering C. Compactness of C implies that there exist finitely many of
them covering C. Take the corresponding finitely many elements of R
alongwith Y \ C to get an open cover of Y.
Let us now check that Y is Hausdorff: let x, y ∈ Y. If both x, y lie
in X, then X Hausdorff implies that there exist disjoint open sets U, V
in X such that x ∈ U and y ∈ V. If x ∈ X and y = ∞, then local
compactness of X allows us to choose a compact set C in X containing
a neighbourhood U of x. Then, U and Y \C are disjoint neighbourhoods
of x and ∞ respectively in Y.
41
of X. As X is Lindelöf, {f −1 (Vα )}α∈Λ admits a countable subcover,
say {f −1 (Vαi )}αi ∈Λ,i∈N . Then it can be checked that (Vαi )αi ∈Λ,i∈N is a
countable subcover of {Vα }α∈Λ for f (X). This proves that f (X) is Lin-
delöf.
Check Your Progress
• Show that if A is a closed subspace of a Lindelöf topological space
X, then A is Lindelöf.
Theorem 3.7.1. For a metric space (X, d) the following are equivalent:
42
(c) =⇒ (a) Let (X, d) be a second countable space. Let {Ui }i∈I
be an open cover of X. Since X has a countable basis, denote this by
{Vn }n∈N , where Vn are open subsets of X. Let S = {n ∈ N | Vn ⊆
Ui for some i ∈ I}. Clearly, S is countable as a subset of a countable
set is countable. Using the fact that {Vn }n∈S is a basis of X, it is easy
to check that {Vn }n∈S is an open cover of X. For each n ∈ S, choose
i(n) ∈ I such that Vn ⊆ Ui(n) , for some i(n) ∈ I. Then, {Ui(n) }n∈S
gives the required countable subcover of X, proving that X is Lindelöf.
(Note here that we did not use the fact that X was a metric space.
Thus this result is more general, but here we will restrict ourselves to
metric spaces.)
In general, Lindelöff property is not hereditary. Here is a counter-
example. Let X be an uncountable set and let x0 ∈ X. Let τ = {A ⊆
X | x0 6∈ A}. It can be checked that τ is indeed a topology on X. It is
also easy to check that X is Lindelöf, in fact X is compact. It can now
be checked that (Y, τY ) with Y = X \ {x0 } with the subspace topology
τY is not Lindelöf.
In general, even a finite product of Lindelöff spaces is not Lindelöff.
Other results regarding these spaces are very technical. We shall not
go into further details about this property, but the interested reader
can look up the references for further reading.
3.8 Glossary
In this chapter, you have learnt the following:
• Open covers and finite subcover of a given cover.
• Tube lemma.
• One-point compactification.
43
• Basic properties of one-point compactification.
3.9 Bibliography
44
2. M. A. Armstrong: Basic Topology, Springer UTM
A set X with a simple order is said to have the least upper bound
property if every non-empty subset X0 of X that is bounded above
has a least upper bound.
Show that every simply ordered set with the least upper bound
property is locally compact.
45
7. Show that if Y is compact, then the projection π1 : X × Y → X
is a closed map i.e., π1 carries closed sets to closed sets.
46
Chapter 4
Chapter Structure
4.1 Introduction
4.2 Objectives
4.3 Equivalent formulations of Compactness for Metric Spaces
4.4 Compact in Rn iff Closed and Bounded
4.5 Completeness and Completion in Metric Spaces
4.5.1 Compete Metric Spaces
4.5.2 Completion of a Metric Space
4.6 Lebesgue Covering Lemma
4.7 Uniform Continuity Theorem
4.8 Glossary
4.9 Bibliography
4.10 Let Us Sum Up
4.11 References for Further Reading
4.12 Chapter End Exercises
4.1 Introduction
The aim of this unit is to give for metric spaces other equivalent
formulations of compactness: sequential compactness and limit point
compactness. The main theorem here is that a subset E of Rn is com-
pact if and only if E is closed and bounded in Rn . We then study
complete metric spaces and characterize them in terms of compactness
and total boundedness. We briefly study the notion of completion of a
metric space. We then try to understand compactness better using the
notion of Lebesgue covering. This helps us to prove a very important
theorem: every continuous map from a compact metric space to any
47
other metric space is uniformly continuous.
4.2 Objectives
After going through this chapter you will be able to:
• Give various equivalent definitions of compactness for metric spaces.
• Show E ⊂ Rn is compact if and only if it is closed and bounded.
• Define complete metric spaces and completion of a metric space.
• State Lebesgue covering lemma and find Lebesgue number of a cov-
ering.
• Prove uniform continuity theorem.
48
X \ B(x1 , ). (This is so as by assumption, X cannot be covered by a
single ball B(x1 , ).) Having chosen x1 , x2 , . . . , xn continue by induction
to get xn+1 6∈ B(x1 , ) ∪ B(x2 , ) · · · ∪ B(xn , ).
The choice of xn ’s implies that d(xn+1 , xi ) ≥ , for all 1 ≤ i ≤ n.
Hence the sequence (xn ) can have no convergent subsequence.
Theorem 4.3.1. Let (X, d) be a metric space. Then one has the fol-
lowing implications:
49
A by f (n) = xn . Then, N = ∪x ∈ A f −1 (x). If the number of
elements in f −1 (x) had been finite for every x ∈ A, then
we would get that the set of natural numbers is a finite
set, which is a contradiction. Hence, there exists x ∈ A
such that f −1 (x) is an infinite subset of N i.e., there exist
infinitely many n such that f (n) := xn = x. Now it is obvious
that (xn )n∈f −1 (x) , being the constant sequence is a convergent
subsequence of (xn )n∈N .
– Suppose A is infinite. As X is limit point compact by as-
sumption, we have that A has a limit point, say x. We will
now define a subsequence of (xn ) converging to x as follows:
since x is a limit point of A, B(x, 1), the ball around x of
radius 1 contains a point of A other than x. Thus, one can
choose n1 such that xn1 ∈ B(x, 1). Applying inductively the
same argument again, given a positive integer ni−1 we can
choose an index ni > ni−1 such that xni ∈ B(x, 1/i). (The
guarantee that ni > ni−1 is due to the fact that A is infinite.)
The choice of xni now implies that the subsequence xn1 , xn2 , . . .
converges to x in A.
50
In this section, we are going to characterize compact sets in Rn .
Before proving that, let us recall some basic notions related to metrics
on Rn . First recall the definition of a metric:
Definition 31. A metric on a set X is a function d : X × X → R
having the following properties:
• For all x, y ∈ X, d(x, y) ≥ 0 where equality holds if and only if
x = y.
4. Use the fact above to prove that the topologies induced by the
Euclidean metric and the square metric are the same.
Proof. The proof is based on the least upper bound property (lub prop-
erty) of real numbers. If In = [an , bn ], let E be the set of all an . Then
E is non-empty and bounded above by b1 , for example. Let x be the
least upper bound of E, which exists as E is a non-empty bounded
subset of R. Now observe that for all m, n ∈ N an ≤ an+m ≤ bn+m ≤ bm
which then implies that x ≤ bm for each m, x being the least among
51
upper bounds of E. By definition, am ≤ x, for all m ∈ N. Thus, for all
m ∈ N, x ∈ [am , bm ] := Im , proving that x ∈ ∩∞
i=1 In . This proves that
∞
∩i=1 In 6= ∅.
52
4.5 Completeness and completion in
metric spaces
We will study another very helpful notion in this section: complete-
ness. The word “completeness” is used in mathematics in the same
sense as in English: absence of gaps. We will then study the relation
between compactness and completeness by introducing a geometric con-
cept of total boundedness.
In the earlier chapters, we have studied the notion of one-point
compactification, which helps us to get a compact space out of a non-
compact one. Here too, we will “complete” spaces that are not com-
plete: for example the set of rational numbers is not a complete metric
space and its completion gives us the real line. Though the formal def-
inition may look very intimidating, the underlying idea is simple: one
is just trying to fill up the gaps in a space which is not complete.
53
Using Theorem 4.5.2, we show that bounded need not imply totally
bounded.
Example 6. Consider the metric space l2 , consisting of all sequences
(xn ) such that xn < ∞. Then, l2 is a metric with d(x, y) := ||x−y||2 ,
P 2
2 21
where for a sequence s = (sn ) ∈ l2 , ||s||2 = ( ∞
P
i=1 sn ) . For each i ∈ N,
let ei be the sequence all whose terms are zero, except the i-th term
which is one. Let E be the set of all ei as above. Then √ one can
easily check that for j 6= k, √ d(e ,
j ke ) = ||e j − e ||
k 2 = 2. Thus, E
is bounded with diameter 2. However E is not totally bounded√as
e1 , e2 , . . . cannot have any Cauchy subsequence, since d(ej , ek ) = 2,
which remains quite large.
54
Definition 37. Two Cauchy sequences (xn ), (yn ) in (X, d) are said to
be equivalent, if d(xn , yn ) → 0 as n → ∞.
55
Definition 38. Let A be a bounded subset of a metric space (X, d).
Then, diameter of A, denoted by diam(A) is defined to be lub{d(a1 , a2 ) | a1 , a2 ∈
A}.
Examples:
(1) Let X = R with the absolute metric. Let A = [0, 1]. Then,
diam(A) := lub{|x − y| | x, y ∈ A}. Note that for all x, y ∈ A,
|x − y| ≤ 1. It is also easy to check that one is the least upper
bound of A. Hence, diam(A) = 1.
Definition 39. Let R be an open covering of the metric space (X, d).
If there exists δ > 0 such that for each subset of X having diameter
less than δ, there exists an element of R containing it, then δ is called
a Lebesgue number for the covering R.
56
Check Your ProgressThe proof above is a nice illustration of
one of the ways of attacking a problem in mathematics: the existen-
tial approach. The above proof asserts the existence of a δ without
actually telling how one can find or construct it. Such a method of
proof was first developed by one of the influential and universal math-
ematicians of the nineteenth and twentieth centuries, David Hilbert.
(See https : //en.wikipedia.org/wiki/David Hilbert). During those
days, Hilbert was criticized for such a method of proof and in fact he
was told by Gordan (another great mathematician) that this was not
mathematics, but it was theology. Later, however Hilbert’s method of
thought became a very important way for going about a proof, specially
in pure mathematics. Many of the proof in mathematics today are of
this kind and one then often needs to write down algorithms to make
the computations explicit.
Can you find at least two more examples of such existential proofs?
(Hint: Linear Algebra).
Let us now prove that for a metric space (X, d) sequential compact-
ness implies compactness. (See Theorem 4.3.1.)
Theorem 4.6.1. Let (X, d) be a metric space. If (X, d) is sequentially
compact, then (X, d) is compact.
Proof. Let R be an open covering of X. Since X is sequentially com-
pact, by Lemma 4.6.1 R has a Lebesgue number δ. Apply Lemma 4.3.1
with = δ/3, to get a finite covering of X by balls of radius δ/3. Each
of these balls has diameter atmost 2δ/3 < δ so we can choose for each
of these balls an element of R containing it. Thus, we get a finite
subcollection of R that covers X, proving that X is compact.
Remark 4.6.1. The property of completion is slightly different from its
sister property of compactness: compactness is a topological property,
but completion is not a topological concept i.e., it is not invariant
under homeomorphism. For example, there is a homeomorphism f :
( −π , π ) → R with R is complete but ( −π
2 2
, π ) not complete.
2 2
We can now record the Lebesgue covering lemma for compact spaces.
The proof of it follows immediately from the facts proved above.
Lemma 4.6.2. Let R be an open covering of the metric space (X, d).
If X is compact, there is a δ > 0 such that for each subset of X having
diameter less than δ, there exists an element of R containing it. Such
a number δ is called a Lebesgue number for the covering R.
Here is an important consequence of completeness:
Theorem 4.6.2. Let (X, d) be a complete metric space. For each n ∈ N
let Fn be a closed and bounded subset of X such that F1 ⊃ F2 ⊃ · · · ⊃
Fn ⊃ Fn+1 ⊃ · · · and diam(Fn ) → 0 as n → ∞. Then ∩∞ n=1 Fn contains
precisely one point.
57
Proof. For each n ∈ N let an be any arbitrary point of Fn . Then as
the Fn ’s form a nested sequence, we get an , an+1 , an+2 , . . . ∈ Fn , i.e.,
an+k ∈ Fn for all k ≥ 0. (Call this property P.)
Since the diameter of Fn ’s tends to zero as n tends to infinity, given
> 0 there exists an integer N ∈ N such that diam(Fn ) < and
aN , aN +1 , aN +2 , . . . all lie in FN . Thus for m, n ≥ N we have d(an , am ) ≤
diam(FN ) < . This proves that {an }∞ n=1 is a Cauchy sequence. Since
X is complete this Cauchy sequence converges to a point, say a in X.
By property P and the fact that a ∈ X, we get that a is a limit point of
Fn , for every n ≥ 1 and for every closed subset Fn . Since for all n ≥ 1,
Fn is closed, we get that a ∈ Fn for all n ≥ 1. This shows that ∩∞ n=1 Fn
is non-empty.
We now prove uniqueness: if there exist a, b ∈ X such that both
a, b ∈ ∩∞n=1 Fn , then there exists a K such that d(a, b) > diam(FK ) for
K sufficiently large. Thus, b cannot lie in ∩∞ n=1 Fn , a contradiction.
58
many subsets of R.) As diamA1 = diamA1 , A1 is a closed subset
of X with diameter less than one and which cannot be covered
by a finite number of sets from R.
Since A1 is itself totally bounded, the same reasoning shows that
there exists A2 ⊂ A1 such that diamA2 < 12 and A2 cannot be
covered by finitely many elements of R. Proceed inductively to
get a nested family of closed subsets of X such that A1 ⊃ A2 ⊃
· · · An ⊃ An+1 · · · with diamAn < n1 and such that no finite
number of sets in R form a covering of any An . By Theorem
4.6.2, ∩∞n=1 An contains precisely one point, say x. Since R is a
covering of X, there is a set G ⊂ R such that x ∈ G. Since G is
open, there exists r > 0 such that B(x, r) ⊂ G. Now if N ∈ N
is such that N1 < r, then diamAN < N1 < r. Since x ∈ AN , we
have AN ⊂ B(x, r) ⊂ G. Thus, G alone covers AN . This is a
contradiction to the fact that finitely many subsets of R cannot
cover AN . This contradiction completes the proof of the theorem.
59
• (Uniformly continuous) Let f : R → R be given by f (x) = 2x.
Then, for every > 0, the choice δ = /2 is such that |x−y| < /2
implies that |f (x) − f (y)| = 2|x − y| < (2)/2 < . Since δ is
independent of the choice of the point x chosen, we conclude that
f is continuous.
60
Check Your Progress
1. Prove that every uniformly continuous function is continuous.
The converse need not hold.
1
2. Show that the function f (x) = 1+x2
for x ∈ R is uniformly con-
tinuous on R.
61
4.8 Glossary
In this chapter, you have learnt the following:
4.9 Bibliography
62
about continuous functions from a compact metric space to another
metric space: these functions are also uniformly continuous.
63
8. Prove that a metric space X is complete, if it contains a dense
subset D such that every Cauchy sequence in D has a limit point
in X.
64