Download as pdf or txt
Download as pdf or txt
You are on page 1of 66

CHAPTER 1

Topology

BY

Dr. Dibyendu De

Associate Professor
Department of Mathematics
University Of Kalyani
West Bengal, India
E-mail : dibyendude@gmail.com

1
Module-1: Introduction to definition
of Topological spaces

To get a proper notion of continuity in abstract case one needs a notion of nearness,
so that one can say about continuous transformation.

The idea is that if one geometric object can be continuously transformed into another,
then the two objects are to be viewed topologically equivalent.
For example, an open interval of length one can be stretched to an open interval of length
two, a circle and a square are topologically equivalent, one can be continuously trans-
formed into another by radial projection.
On the other hand, a closed interval is topologically distinct from a circle or square. In
fact, when a point is removed from a circle it remains is still connected but if a point
lying between 0 and 1 remover from a closed interval it produces two different pieces.

The term used to describe two geometric objects are topologically equivalent is home-
omorphism.
Thus a circle and a square are homeomorphic. Concretely, a radial projection from a
circle C to a square S with the same center point, produces a homeomorphism between
C and S.
One of the basic problems of Topology is to determine when two given geometric objects
are non homeomorphic, for example whether the letters M, N, H, B are non homeomor-
phic.
This can be quite difficult in general. Our first goal will be to define exactly what the
‘geometric objects’ are that one studies in Topology. These are called topological spaces.

2
Roughly speaking continuous function from one topological space to another is that
which preserve nearness.
Beyond ε − δ definition a nice definition of a continuous function in terms of open
sets:

Definition 1. A function f : R → R is continuous if for each open set O in R the inverse


image f −1 (O) = {x ∈ R : f (x) ∈ O} is also an open set.

To see that this definition corresponds to the intuitive notion of continuity, let us
examine, what would happen if this condition fails. There would then be an open set O
for which f −1 (O) is not open. This means there would be a point x0 ∈ f −1 (O) for which
there is no interval (a, b) containing x0 and contained in f −1 (O).

This is equivalent to saying there would be points x arbitrarily close to x0 those are
in the complement of f −1 (O). For x to be in the complement of f −1 (O) means that f (x)
is not in O. On the other hand, x0 was in f −1 (O) so f (x0 ) is in O. Since O was assumed
to be open, there is an interval (c, d) about f (x0 ) that is contained in O. The points
f (x), those are not in O are therefore not in (c, d), so they remain at least a fixed positive
distance from f (x0 ). A reasonable interpretation of discontinuity of f at x0 would be
that there are points x arbitrarily close to x0 for which f (x) stays at least a fixed positive
distance say ε away from f (x0 ). Let O be the open set (f (x0 ) − ε, f (x0 ) + ε). Then
f −1 (O) contains x0 but it does not contain any points x for which f (x) is not in O, and
we are assuming there are such points x arbitrarily close to x0 , so f −1 (O) is not open
since it does not contain all points in some interval (a, b) about x0 .

In trying to find a satisfactory definition of a topological space we shall have two aims
in mind.

1. The definition should be general enough to allow a wide range of different structures
as spaces. We would like to consider a finite, discrete set of points as a space, or equally
a whole uncountable continuum of points such as the real line; our nice geometrical sur-
faces should qualify under the definition, and also sets of functions such as the set of
continuous complex-valued functions defined on the unit circle in the complex plane.

3
2. The definition will be so that, we would be able to perform simple constructions with
our spaces, such as taking the cartesian product of two spaces, or identifying some of the
points of a space in order to form a new one. For example in the construction of Mobius
strip we take a rectangular piece of rubber sheets and then identify two opposite sides in
different direction.

The definition of a space should contain enough information so that we can define the
notion of continuity for functions between spaces. It is really this second consideration
which leads to the abstract definition given below.

Definition 2. A topological space is a non empty set X together with a collection O of


subsets of X, called open sets, such that:
(i) Both ∅ and X are in O .
(ii) The union of any collection of sets in O is in O.
(iii) The intersection of any finite collection of sets in O is in O.
The collection O of open sets is called a topology on X.

The most known example of topological space is R with usual open sets as the open
sets of the topology of R.

Example 1. Let X be a nonempty set and O = P(X). Then it is easy to observe that
(X, O) is a topological space called discrete space also denote by Xd .

Example 2. Let X be a nonempty set and O = {∅, X}. Then it is topological space
called indiscrete space.

Example 3. Let X = {a, b, c} and O = {∅, X, {a}, {b}, {a, b}}. Then observed that O is
a topology on X, so that (X, O) is a topological space.

Example 4. Let X be an infinite set and let τcf = {A ⊂ X : X \ A is finite}. Then


(X, τcf ) is a topological space, usually called cofinite space.

4
Proof. ∅ ∈ τcf by default and X ∈ τcf is obvious.
Now let A1 , A2 ∈ τcf . Then

X \ (A1 ∩ A2 ) = (X \ A1 ) ∪ (X \ A2 )

is finite as both is finite and therefore A1 ∩ A2 ∈ τcf .


τcf is closed under arbitrary union is obvious.

Example 5. Let X be an uncountable set and let τco = {A ⊂ X : X\A is countable or finite}.
Then (X, τco ) is a topological space, usually called cocountable space.

Example 6. Let X be a plane. Let O consist of ∅, X and all open disks with centre at
the origin. Then clearly O is topology on X.

Next we define an interesting topology. First we define the following notion.

Definition 3. Let X = {0, 1, 2, ..., n, .....} = N ∪ {0}. For A ⊂ X, let |A ∩ [1, n]| denotes
the cardinality of the set A ∩ [1, n].
T
Theorem 1. Define τ = {A : 0 ∈
/ A or 0 ∈ A and lim |A [1,n]|
= 1}. Then τ is a
n→∞ n

topology on X.

Proof. Clearly ∅ ∈ τ .
Observe that lim |X∩[1,n]| = 1 so that X ∈ τ . If A, B ∈ τ , that is lim |A∩[1,n]| =1
n→∞ n n→∞ n
c c
and lim |B∩[1,n]| = 1. Then lim |A ∩[1,n]| = 0 and lim |B ∩[1,n]| = 0. This implies
n→∞ n n→∞ n n→∞ n
c c
that lim |(A ∪Bn)∩[1,n]| = 0. Since X \(A ∩ B) = (X \ A)∪(X \ B) we have the desired
n→∞
result. Hence τ is closed under finite intersection. That τ is closed under arbitrary union
is easy.

As elementary real analysis we declare a set in a topological space X to be closed if


it is complement of an open set.

Theorem 2. Let X be a topological Then the following conditions hold:


(1) ∅, X are closed sets.
(2) Arbitrary intersection of closed sets is closed
(3) Finite union of closed sets is closed.

5
Proof. (1) is obvious.
(2) Let {Fα : α ∈ Γ} be a collection of closed sets. Then by definition {X \ Fα : α ∈ Γ}
is a collection of open sets. Using elementary set theory the result follows.

Like elementary real analysis we can introduce the notions of Interior, limit points,
closure, boundary points for arbitrary topological spaces.

Definition 4. Let A be a subset of a topological space X and x ∈ X. Then x is said to


be a limit point of A if every open set containing x meets A at a point other that x. If x
is not a limit point of A then it is an isolated point of A. Hence any isolated point of X
is just an open set. A point x ∈ A is said to be an interior point of A if there exists an
open B such that x ∈ B ⊂ A.

Definition 5. Let A be a subset of a topological space X. A point x ∈ A is said to be an


interior point of A if there exists a open O in X such that x ∈ O ⊂ A.

The set of all limit points of a set A is called derived set of A, By the closure of A
we mean the set A along with its limit points, and denoted by A. The set of all interior
points of A called interior of A and denoted by A◦ .

From elementary analysis we already know that closure of Q in R is whole R and


interior of Q is emptyset.

Question 1. What will be the interior, closure of the following sets : (a) {(x, y)|1 <
x2 + y 2 ≤ 2}
(b) R2 with both axes removed

Proposition 1. Let X be a topological space, and A, B be subsets of X. Then prove the


followings:
(1) If A ⊂ B then A ⊂ B,
(2)A ∪ B = A ∪ B,
(3)(A ∩ B)◦ = A◦ ∩ B ◦ .

6
Proof. (1) follows easily. For part (2) let x ∈ A ∪ B and U be a neighborhood of x. Then
U ∩ (A ∪ B) 6= ∅. This means that either U ∩ A 6= ∅ or U ∩ B 6= ∅. This implies that
x ∈ A ∪ B, so that A ∪ B ⊆ A ∪ B.
Conversely let x ∈ A ∪ B and U be a neighborhood of x. Then either U ∩ A 6= ∅ or
U ∩ B 6= ∅, that is U ∩ (A ∪ B) 6= ∅ so that x ∈ A ∪ B. This gives the reverse inequality.

Theorem 3. Let X be a topological and A ⊂ X. Then the followings hold:


(1) A◦ is the largest open set contained in A.
(2) A is the smallest closed set containing A.
(3) A is open if and only if A = A◦ .
(4) A is closed if and only if A = A.

Proof. (1) If x ∈ A◦ then there exists some open Bx such that x ∈ Bx ⊂ A. We have
Bx ⊂ A◦ . In fact for each y ∈ Bx , Bx is an open set contained in A and containing y.
Now let O be an open set contained in A. Then for each x ∈ O there exists some open
Bx such that x ∈ Bx ⊂ A, i.e. each element of O is an interior point of A so that O ⊂ A◦ .

(2) Let x ∈ X \ A. Then there exists an open set O containing x which misses A.
We claim that O also misses A. If not so then there exists some y ∈ A \ A such that
y ∈ O. This implies that O meets A, which is a contradiction. Therefore A is closed. To
prove that A is the smallest closed set containing A, let K be a closed set containing A.
If possible let there exists some y ∈ A \ K. Then there exists an open set O containing
y, which misses A, which contradicts the fact that y ∈ A.

Example 7. Consider the set R2 \ {(x, sin x1 ) : x > 0}.

The graph of so called sin curve is is displayed in pic-


ture. Any point on the sin curve cann’t be a limit point. So the closure of the given set
is itself. Now choose a point on the set {0} × [−1, −1]. Each point on this set is a limit
point of the set {(x, sin x1 ) : x > 0} and not an interior point of the set in question. Hence
the interior of the set in question is R2 \ ({(x, sin x1 ) : x > 0} ∪ {0} × [−1, −1]).

7
We can define convergency of sequence in topological spaces analogues to metric space.

Definition 6. Let X be a topological space and x ∈ X. A sequence (xn )∞


n=1 is said

to converge at x if for any neighborhood Nx of x there exists some n0 ∈ N such that


(∀n ≥ n0 )(xn ∈ Nx ).

In the following we will examine the property which is responsible for unique limit of
any convergent sequence in Ru .

Definition 7. Let X be a topological space

(1) X is said to be T1 if every finite set is closed in X. (2) X is said to be Hausdorff


or T2 if any two distinct points can be strongly separated by two disjoint open sets.

Proof. Let x ∈ X and y 6= x. Then there exist disjoint open sets U and V containing
x and y respectively. Therefore y is not a limit point of the set {x}. This implies that
every singletoned is closed and therefore every finite set is closed.

Note that cofinite space is not Hausdorff. In fact if x and y are two distinct points U
and V are two disjoint opensets containing x and y respectively, then X \ U and X \ U
are finite. But then (X \ U ) ∪ (X \ V ) = X \ (U ∩ V ). But the left hand side being finite
and the right hand side being whole X as U ∩ V = ∅ we get a contradiction.
One can observe that for the cofinite space every sequence converges to every point.
In this position we like to introduce the new two definitions. Here we prove that Haus-
dorfness is responsible for unique limit of any convergent sequence.

Proposition 2. In a Hausdorff space any sequence can have at most one limit.

Proof. Let x ∈ X and y 6= x. Then there exist disjoint open sets U and V containing
x and y respectively. Now if a sequence (xn ) converges to x then it is not possible to
converge to y, as except finitely many points, all the points of the sequence are in U .

8
CHAPTER 1

Topology

BY

Dr. Dibyendu De

Associate Professor
Department of Mathematics
University Of Kalyani
West Bengal, India
E-mail : dibyendude@gmail.com

1
Module-2: Base of Topological
spaces

Since open sets are defined in terms of open intervals many arguments with open sets
in R reduce to looking at what happens with open intervals. A similar statement holds
for Rn with open balls in place of open intervals. In each case arbitrary open sets are
unions of the special open sets given by open intervals, and balls. Generalizing this idea
we introduce the following definition.

Definition 1. Let X be a nonempty set. A collection B is said to be a basis for some


topology on X if B satisfies the following two properties:
(1) Every point x ∈ X lies in some set B ∈ B .
(2) For each pair of sets B1 , B2 in B and each point x ∈ B1 ∩ B2 there exists a set
B3 in B with x ∈ B3 ⊂ B1 ∩ B2 .

The topology generated by a basis B, generally denoted by τ (B) can be defined as


follows :
A subset O ⊂ X is to be declared as open if for any x ∈ O there exists some B ∈ B
such that x ∈ B ⊂ O.

It needs to prove that this τ (B) is in fact a topology.

Theorem 1. Let X be a non empty set and B be a basis. Let τ (B) be the collection
defined as follows: U ∈ τ (B) if for each x ∈ U there exists some B ∈ B with the property
x ∈ B ⊂ U . Then τ (B) is a topology on X.

Proof. Vacuously ∅ ∈ τ (B) and X ∈ τ (B) is obvious.


That τ (B) is closed under arbitrary union is also clear.
Finite intersection property follows from point (2) of definition of basis.

2
Most known example is that R with usual topology has the following two bases:
1. {(a, b) : a, b ∈ R}
2. {(a, b) : a, b ∈ Q}.
Another way of describing the topology generated by a basis is given in the following
lemma:

Lemma 1. Let X be a set; let B be a basis on X. Then τ (B), the topology generated by
B equals the collection of all unions of elements of B.

Proof. Since τ (B) is a topology generated by B, all possible union of members of B is a


subset of τ (B). Conversely, given U ∈ τ (B) choose for each x ∈ U an element Bx of B
S
such that x ∈ Bx ⊂ U . Then U = x∈U Bx , so U equals a union of elements of B.

Observation 1. The above theorem shows that a topology on a set is all possible unions
of members of base.

In the above Theorem we mentioned how to from topology from a basis. The following
is one way of obtaining a basis for a given topology. We shall use it frequently.

Proposition 1. Let B is a collection of open sets of a topological space X satisfying


that for each open set U of X and each x in U , there is an element in B ∈ B such that
x ∈ B ⊂ U . Then B is base for the topological space X.

Proof. The first condition, is obvious. For the second condition, suppose B1 and B2 are
elements of B and x ∈ B1 ∩ B2 . Since B1 ∩ B2 is an open set there exists some B3 ∈ B
such that x ∈ B3 ⊂ B1 ∩ B2 .
The topology generated by B equals to the topology of X, is left as an exercise.

We have already seen that both the following sets generates the usual topology on R.
1. {(a, b) : a, b ∈ R}
2. {(a, b) : a, b ∈ Q}. But when topologies are given by bases, it is useful to have a
criterion in terms of the bases for determining whether one topology is finer than another.

Proposition 2. Let B and B1 for the topologies τ1 and τ2 on a set X. Then the following
conditions are equivalent.

3
(1) τ2 is finer than τ1 .
(2) For each x ∈ X and each basis element B ∈ B containing x, there is a basis element
B2 ∈ B2 such that x ∈ B2 ⊂ B1 .

Proof. 1) ⇒ 2). Given an element U of τ1 , we wish to show that U is in τ2 also. Let


x ∈ U . Since B1 generates τ1 there is an element B ∈ B such that x ∈ B ⊂ U . Condition
2) tells us there exists an element B2 ∈ B2 such that x ∈ B2 ⊂ B1 . Then x ∈ B2 ⊂ U ,so
U is in τ2 by definition.

2) ⇒ 1). We are given x ∈ X and B ∈ B, with x ∈ B. Now B belongs to τ1 by


definition and τ1 ⊂ τ2 . by condition 1); therefore, B ∈ τ2 Since τ2 is generated by B2
there is an element B2 ∈ B2 such that x ∈ B2 ⊂ B1 .

Example 1. 1. One of the most beautiful example of the above Theorem is that the set
{(a, b) : a, b ∈ R} and {(a, b) : a, b ∈ Q} both generates the same topology R. It happens
as Q is dense in R.
2. Another beautiful application of the above Theorem is that open rectangles in R2 and
open discs in R2 generates the same topology on R2 . In fact if we take an open disk D
and a point x in D then we can inscribed a rectangle in D containing x. Similarly the
other. Therefore the topologies generated by them are equivalent.

Example 2. Sorjenfrey line is an extremely important topological space. This is alter-


natively known as lower limit topology.

Consider the collection B = {[a, b) : a, b ∈ R} of subsets of R. Then it is a basis for


some topology on R. In fact
a. For any r ∈ R, r ∈ [r, r + 1),
b. If [a, b) and [c, d) be two members of B and r ∈ [a, b)∩[c, d) then clearly the intersection
is in B.

4
Therefore the collection B generates some topology on R, which is known as lower
limit topology and denoted by Rl .

An interesting point to observe that, Rl has a basis each of whose member is closed
as well as open. In fact each set [a, b) is also closed in Rl . In fact if x 6∈ [a, b) then either
x < a or x ≥ b. In the first case we can choose r ∈ (x, a) sot that [x, a) misses [a, b) and
in the second case [x, x + 1) misses [a, b). Therefore x is not a limit point of [a, b) and
hence [a, b) is a closed set.

Another important observation about Rl is that each open interval in Rl is open, that
means that the topology of Rl is larger than the usual topology.

Example 3. Show that the collection

B = {[x, y) : x < y and x, y ∈ Q}

is a basis that generates a topology different from the lower limit topology on R.

Example 4. Consider the set K = { n1 : n ∈ N} and the subsets of the form (a, b) \ K.
The collection

B1 = {(a, b) ⊂ R : a, b ∈ R} ∪ {(a, b) \ K ⊂ R : a, b ∈ R}

is a basis for a topology on R: The topology it generates is known as the K-topology on R:


Clearly, K-topology is finer than the usual topology. Note that there is no neighbourhood
of 0 in the usual topology which is contained in (−1, 1) \ K ∈ B. This shows that the
usual topology is not finer than K-topology. The same argument shows that the lower
limit topology is not finer than K-topology. Consider next the nbd [2, 3) of 2 in Rl . Then
there is no nbd of 2 in the K-topology which is contained in [2, 3). Thus we conclude that
the K-topology and the lower limit topology are not comparable.

A partially ordered set is a pair (X, ≤), where X is a set and ≤ is a relation on X such
that: (i) x ≤ x for all x; (ii) if x ≤ y and y ≤ z, then x ≤ z; (iii) if x ≤ y and y ≤ x, then
x = y. Let us call a partially ordered set linearly ordered if whenever x and y are in X,
either x ≤ y or y ≤ x. (For example, X = R or any of its subsets is linearly ordered.) (a)
If X is a partially ordered set and S is the collection of all sets having the form of either

5
{y : y ≤ x and y 6= x} or {y : x ≤ y and y 6= x}, we can show that S is a subbase for a
topology on X. This is called the order topology on X. (b) Show that if (X, ≤) is linearly
ordered, then the order topology satisfies the Hausdorff property. Can you find another
condition on the ordering such that the order topology has the Hausdorff property? (c)
When a and b are elements of a partially ordered space, let (a, b) = {x ∈ X : a < x < b}.
If (X, ≤) is linearly ordered, show that (a, b) = {x ∈ X : a < x < b} is a base of the
order topology.

Definition 2. A neighborhood of a point x in a topological space X is any set A ⊂ X


that contains an open set O containing x. Dually x is said to be an interior point of
A; that is, A is a neighbourhood of x if and only if x ∈ A◦ . The collection Ax of all
neighbourhoods of x is the neighbourhood system of x.

Proposition 3. The neighbourhood system Ax at x in a topological space X has the


following properties:
(a) Ax 6= ∅, for all x ∈ X
(b) if A ∈ Ax then x ∈ A
(c) if A1 , A2 ∈ Ax then A1 ∩ A2 ∈ Ax
(d) if A ∈ Ax then there is a B ∈ Ax such that A ∈ Ay for each y ∈ B
(e) if A ∈ Ax and A ⊂ B then B ∈ Ax .
O ⊂ X is open if and only if O contains a neighbourhood of each of its points. Show
that this conditions generates a unique topology on X.

Proof. Let τ be the collection of all such sets in X. Vacuously ∅ ∈ τ and X ∈ τ is


obvious. Let A1 , A2 ∈ τ and x ∈ A1 ∩ A2 . Then there exists O1 , O2 ∈ Ax such that
x ∈ Oi ⊂ Ai for all i. So by condition (c) A1 , A2 ∈ τ . Arbitrary union of membered of τ
is in τ is obvious. Hence τ is a topology.

As an example of this property consider the set of open balls in a metric space.
For a ∈ Z and K ∈ N, we define a + KZ = {a + Kn : n ∈ Z}. We say that A ⊂ Z is
open if for each point a ∈ A, there is a number K ∈ N such that {a + Kn : n ∈ Z} ⊂ A.
In other words, a subset of Z is open if each of its points is contained in an arithmetic
progression belonging to the set. Obviously the sets {a + Kn : n ∈ Z} are open. So the
arithmetic progressions form a basis for the topology. Surprisingly, this basic sets are

6
closed also because the complement of a + KZ = {a + Kn : n ∈ Z} is the union of other
arithmetic progressions with the same difference.

There are many proofs that there exist an infinity of primes. Using this topology we
can prove that the infiniteness of primes.

Proposition 4. The number of primes is infinite.

Proof. Since the sets {i, i + d, i + 2d, ........}, i = 1, 2, ...., d are open, pairwise disjoint
and cover the whole N, it follows that each of them is closed. In particular, for each
prime number p the set {p, 2p, 3p, .......} is closed. All together, the set of the form
{p, 2p, 3p, ......} cover N − {1}. Hence if the set of prime numbers were finite, then the
set {1} would be open. However, it is not a union of arithmetic progressions.

Proposition 5. Let X be a set. Given any family S = {Sα : α ∈ I} with X ⊂ ∪α Sα .


of subsets of X, there always exist a unique smallest topology generated by S. S is called
a subbasis for τ (S), the topology generated by S.

Proof. Let τ (S) be the intersection of all topologies containing S; such topology exists,
since P(X) is one such. So clearly τ (S) is a topology. It evidently satisfies the re-
quirements of ‘unique’ and ‘smallest’. To verify the members of τ (S) are as described,
S
note that since S ⊂ τ (S) so τ (S) must contain all the sets listed. Conversely, since
T α
distributes over , the sets listed actually do form a topology containing S, and which
therefore contains τ (S).

Remark 1. The construction of a topology from a subbasis loses some control over the
open sets; they build up from the finite intersections of the Sα ’s rather than from the Sα
themselves.

Example 5. Every basis for a topology is a subbasis.


(a)S = {(a, ∞), (−∞, b) : a, b ∈ R} is a subbasis for the usual topology on R. We can
restrict the numbers a in these intervals to be rational or irrational, and we still have a
subbasis that generates the usual topology.
(b) S = {[a, ∞), (−∞, b) : a, b ∈ R} is a subbasis for Rl
(c) S = {R − {a} : a ∈ R} is a subbasis for the cofinite topology on R.

7
CHAPTER 1

Topology

BY

Dr. Dibyendu De

Associate Professor
Department of Mathematics
University Of Kalyani
West Bengal, India
E-mail : dibyendude@gmail.com

1
Module-3: New spaces from old one

In this chapter we deal with constructing new spaces from old ones. This topic which
seems simple enough at first glance, but turns out to be a source of many examples of
topological spaces.
Given a topology O on a space X and a subset A ⊂ X, we would like to use the
topology on X to define a topology OA on A. This can be done just taking the trace of
open sets in X on A. Define a set O ⊂ A to be in OA if there exists an open set O0 in O
such that O = A ∩ O0 . Let us first prove that this is in fact a topology.

Proposition 1. Prove that OA is in fact a topology on A.

Proof. OA contains ∅ and A as

∅ = A ∩ ∅ and A = A ∩ X.

The fact that it is closed under finite intersection and arbitrary union follows from
the following equations

(U1 ∩ A) ∩ (U2 ∩ A) ∩ . . . ∩ (Un ∩ A) = (U1 ∩ U2 ∩ . . . ∩ Un ) ∩ A,

!
[ [
(Uα ∩ A) = Uα ∩ A.
α∈J α∈J

The topology OA on A is called the subspace topology, and A with this topology is
called a subspace of X. For example, if we take X to be R2 with its usual topology, then
every subset of R2 becomes a topological space. In particular, geometric figures such
as circles and polygons can now be viewed as topological spaces. Likewise, geometric
figures in R3 such as spheres and polyhedra become topological spaces, with the subspace
topology from the usual topology on R3 .

2
In case the space X is a metric space, any subset A ⊂ X becomes a metric space
by restricting the metric X × X −→ R to A × A , since the three defining properties of
ametric obviously still hold for the restricted distance function. The following Proposition
gives some strong evidence that the subspace topology is a natural topology to use on
subsets.

Proposition 2. The metric topology on a subset A of a metric space X is the same as


the subspace topology.

Proof. Observe first that for a ball Br (x) in X, the intersection A ∩ Br (x) consists of all
points in A of distance less than r from x, so this is a ball in A regarded as a metric
space in itself. For a collection of such balls Br (x) we have

A ∩ (∪α Brα (xα )) = ∪α (A ∩ Brα (xα )) .

The left side of this equation is a typical open set in A with the subspace topology,
and the right side is a typical open set in the metric topology, so the two topologies
coincide.

For a subspace A ⊂ X, a subset of A which is open or closed in A need not be open


or closed in X. However, we have the following fact, the easy proof is left to the readers.

Lemma 1. For a subspace Y ⊂ X which is open (resp. closed) in X, a subset A ⊂ Y is


open in Y (resp. closed) if and only if it is open (resp. closed) in X.

Question 1. Give an example of a topological space to establish that open or closed for
the subspace Y in the above lemma can not be removed.

Consider the closed interval [0, 1] in Ru . (0, 1] is open in [0, 1] in the subspace topology
but not open in Ru .
The following Theorem show that closures behave nicely with respect to subspaces:

Theorem 1. Given a space X, a subspace Y , and a subset A ⊂ Y , then the closure of A


in the space Y is the intersection of the closure of A in X with Y . That is a point y ∈ Y
is a limit point of A in Y (i.e. using the subspace topology on Y ) if and only if y is a
limit point of A in X.

3
Proof. For a point y ∈ Y to be a limit point of A in X means that every open set O in
X that contains y meets A. Since A ⊂ Y , this is equivalent to O ∩ Y meeting A, or in
other words, that every open set in Y containing y meets A.

Question 2. Show by an example that analogues statement of the above theorem is not
true for interiors.

Exercises
(1) Prove that subspae topology on Z induced from R is discrete.

(2) Prove that subspace topology on Zn induced from Rn is discrete.

(3) Show that if Y is a subspace of X, and A is a subset of Y , then the topology on A


inherits as a subspace of Y is the same as the topology it inherits as a subspace of X.

Theorem 2. Let X be a linearly ordered topological space and let Y be a subset of X that
is convex in X. Then the order topology on Y is the same as the topology on Y inherits
as a subspace of X.

Question 3. Give an example to show that convexity can not be removed in the above
Theorem.

Consider Z in R.
If L is a straight line in the plane, describe the topology L inherits as a subspace
of Rl × R and as a subspace of Rl × Rl . In each case it is a familiar topology.

So in the first case The topol-


ogy is lower limit topology and in the second case the topology is discrete.

Circle is an important subspace of R2 .

4
If we take trace of any open disk we get a an open arc, on the circle which are open sets
of the circle. Mathematically it can be expressed as {e2πiθ : 0 < θ ≤ 1} and denoted by
S 1.

Sphere is an important subspace of R3 .


If we take trace of any open solid ball we get a an open disk, on the sphere which are
open sets of the sphere. Sphere generally denoted by denoted by S 2 .

This is a figure of an in-


teresting topological space torous, which we shall discuss elaborately latter.

Figure 8 is another interesting topological space, which has more importance in alge-
braic topology.
Assigning a standard sort of topology to the cartesian product of spaces is a tool of
constructing new topological spaces from old. For example, the Euclidean plane is the
product space of the real numbers (with usual topology) with itself and Euclidean n-
space is the product of the real numbers n times.

5
Our objective now is to define a topology on the cartesian product of topological
spaces in some natural and useful way.
If X and Y are topological spaces, we can define a topology on X × Y by saying that
a basis consists of the subsets U × V as U ranges over open sets in X and V ranges over
open sets in Y .

In the next Theorem we shall show that it is sufficient to consider basis elements also.

Theorem 3. Let X and Y be topological spaces with bases B and C. Then the collection

B × C = {U × V : U ∈ B, V ∈ C}

is a basis for the product topology on X × Y .

Proof. The first condition for basis is obvious.


Now let U × V and M × N be in B × C and let (x, y) ∈ (U × V ) ∩ (M × N ). Then there
exists A ∈ B and B ∈ C such that x ∈ A ⊂ U ∩ M and y ∈ B ⊂ V ∩ N . Therefore
(x, y) ∈ A × B ⊂ (U × V ) ∩ (M × N ). Hence B × C is a basis.

More generally one can define the productX = X1 × . . . × Xn to consist of all ordered
n-tuples (x1 , . . . , xn ) with xi ∈ Xi for each i. A basis for the product topology on
X = X1 × . . . × Xn consists of all products U1 × . . . × Un as each Ui ranges over open sets
in Xi , or just over a basis for the topology on Xi .
Thus Rn with its usual topology is also describable as the product of n copies of R,
with basis the open ‘boxes’ (a1 , b1 ) × . . . × (an , bn ).

Definition 1. The map πi : X → Xi defined by πi (x1 , . . . , xn ) = xi is called i-th projec-


tion map.

It is obvious that projection maps are always surjective.

6
CHAPTER 1

Topology

BY

Dr. Dibyendu De

Associate Professor
Department of Mathematics
University Of Kalyani
West Bengal, India
E-mail : dibyendude@gmail.com

1
Module-4: Introduction to
Continuity

Recall that a function f : R → R is continuous if f −1 (O) is open in R for each open set
O in R. We take this as formal definition of continuous function for arbitrary topological
spaces.

Definition 1. A function f : X → Y between two spaces X and Y is continuous if


f −1 (O) is open in X for each open set O in Y .

Definition of continuous function may be described using closed sets as well, which
can be shown from the following Prposition.

Proposition 1. A function f : X → Y is continuous if and only if f −1 (C) is closed in


X for each closed set C in Y .

Proof. It follows from the fact X \ f −1 (K) = f −1 (Y \ K).

Example 1. Let X = R with discrete topology τ1 and Y = R with usual topology τ2 . If


we define f : X → Y by f (x) = x then f is continuous. But if we take a map g : Y → X
defined by g(y) = y, is g is not continuous

The following proposition shows that continuous maps behave well under formation
of closure.

Theorem 1. Let f : X → Y be a function, where X and Y be given topological spaces.


Then the following conditions are equivalent.
(a) f : X → Y is continuous,
(b) For any subset A ⊂ X, f (c`X A) ⊂ c`Y f (A),
(c) For each x in X and each neighborhood U of f (x) there is a neighborhood V of x such
that f (V ) ⊂ U .

2
Proof. (a) ⇔ (b) If x is a limit point of A then f (x) is a limit point of f (A) (verify).
Hence f (c`X A) ⊂ c`Y f (A).
Conversely let K be a closed set in Y and A = f −1 (K). Thenf (c`X A) ⊂ c`Y f (A) =
c`Y f ((f −1 (K)). This implies that f (c`X A) ⊂ c`Y K = K, so that A ⊂ c`X A ⊂ f −1 (K) =
A. Hence c`X A = f −1 (K), that is f −1 (K) is closed and therefore f is continuous.

(a) ⇔ (c) The necessary part is obvious. Let V be a non empty open set of Y and
choose x ∈ f −1 (V ). This implies that f (x) ∈ V . Then by the given hypothesis there
exists a neighborhood U of x such that f (U ) ⊂ V , so that x ∈ U ⊂ f −1 (V ). Hence x is
an interior point and hence f is continuous.

Question 2. Let τ1 and τ2 be two topology on X. Prove that the identity map id :
(X, τ1 ) → (X, τ2 ) is contnuous if and only if τ2 ⊂ τ1 .

Question 3. Let f : X → Y be a continuous map. Find out whether or not it is


continuous with respect to
(a) a finer topology on X and the same topology on Y .
(b) a coarser topology on X and the same topology on Y .
(c) a finer topology on Y and the same topology on X.
(d) a coarser topology on Y and the same topology on X.

The following lemma will be useful in our course.

Lemma 1. Let (Xα , τα )α∈I be a collection of topological spaces, indexed by the set I,
and we have maps fα : X → Xα , α ∈ I. Then there is a smallest topology τ on X for
which the maps fα are continuous.

Proof. Clearly discrete topology makes all fα continuous. A topology σ on X makes all
−1 (U ) : U ∈ τ , α ∈ I}.
the fα continuous if and only if it contains A = {fα α

The following fact again shows that information about base is sufficient for any topo-
logical spaces, whose easy proof is left as exercise.

Lemma 2. Given a function f : X → Y and a basis B for Y , then f is continuous if


and only if f −1 (B) is open in X for each B ∈ B.

3
Proof. Let O be an open set of Y and B be a basis for the topology of Y . Then O = ∪α Bα ,
where Bα ∈ B. Hence f −1 (O) = f −1 (∪α Bα ) = ∪α f −1 Bα . Since each Bα is open in Y
and open sets are closed under union we have that f −1 (O) is also open.

The following lemma shows that continuous maps behave well under composition.

Lemma 3. If f : X → Y and g : Y → Z are continuous, then their composition


g ◦ f : X → Z is also continuous.

Proof. To prove that it is sufficient to observe that for any open set U in Z, (g ◦f )−1 (U ) =
f −1 (g −1 (U )).

Lemma 4. Inclusion maps are always continuous functions.

In fact if U is open in X, A is a subspace of X, then i−1


A (U ) = U ∩ A, which is open

in A by definition of the subspace topology.

Lemma 5. If f : X → Y is continuous and A is a subspace of X, then the restriction


f/A of f to A is continuous as a function A → Y .

Proof. Obvious.

Theorem 4. If X is a topological space then C(X), the set of all real valued continuous
functions form a lattice ordered ring.

Proof. If f, g ∈ C(X), then f + g and f · g defined by

(f + g)(x) = f (x) + g(x) for all x ∈ X and f · g(x) = f (x) · g(x) for all x ∈ X

are continuous. Hence C(X) becomes a ring under pointwise addition and multiplication.

f ≤ g iff f (x) ≤ g(x) for all x ∈ X makes C(X) a lattice ordered ring.

Theorem 5. If X = [0, 1] then C(X) becomes a complete normed linear space.

Proof. That C(X) is a vectorspace follows from the above Theorem. Now if we define

kf k = sup{f (x) : x ∈ X}

then it can be easily proved that k · k is a norm on C(X), which is complete.

4
Theorem 6. Let X = A ∪ B, where A and B are closed sets in X. Let f : A → Y and
g : B → Y be continuous. If f (x) = g(x) for all x ∈ A ∩ B, then there exists a continuous
function h : X → Y such that h(x) = f (x) for all x ∈ A and h(x) = g(x) for all x ∈ B.

Proof. Let C be a closed subset of Y . Then h−1 (C) = f −1 (C) ∪ g −1 (C). Since f is
continuous, f −1 (C) is closed in A and, therefore, closed in X. Similarly, g −1 (C) is closed
in B and therefore closed in X. Their union h−1 (C) is thus closed in X.

We introduce the following definition in contrast to continuous mapping.

Definition 2. A function f : X → Y between topological spaces is said to be open if for


any open (resp. closed) set A, in X, f (A) is open (resp. closed) in Y .

Example 2. (a) Give an example of a continuous function which is neither open nor
closed.
(b) Give an example of a open map which is not continuous.
(c) Give an example of a closed map which is not continuous.

Example 3. Continuous but neither open nor closed The mapping p : [0, 1) → S 1 defined
by p(x) = e2πix is a continuous bijection, which is neither open nor closed.

Example 4. Open but not continuous The mapping 1 : R → Rl is an example of a open


map which is not continuous.

Example 5. Closed but not continuous The mapping 1 : R → Rl is an example of a


closed map which is not continuous.

Example 6. The graph of a function f : R → R is defined to be the subset {(x, f (x)) :


x ∈ R} ⊂ R2 . Show that the graph of a continuous function is a closed set in the standard
topology on R2 .

Proof.

5
Example 7. Let (X, τ ) be a topological space and (Y, d) be a metric space. If f, g : X → Y
are continuous then the set
{x ∈ X : f (x) = g(x)} is closed. To see this, let E={x ∈ X : f (x) = g(x)}. We show
that the complement of E is open.
Let y ∈ X − E. Then f (y) 6= g(y). We can find open sets U and V such that
f (y) ∈ U ,g(y) ∈ V and U V = φ. Now f −1 (U ) is open, as is g −1 (V ), so y ∈
T

f −1 (U ) g −1 (V ) ∈ τ . But f −1 (U ) g −1 (V ) ⊂ X − E. Thus X − E is open.


T T

Note that in the above example if Y is taken Hausdorff space then the result also
remains valid.

Question 7. Let (X, τ ) be a topological space and (Y, d) be a metric space. If f, g : X →


Y are continuous and f (x) = g(x) for all x ∈ A, where A is dense in X, show that
f (x) = g(x) for all x ∈ X. What happen if we replace the mtric space’ by (Y, σ) by a a
Hausdorff space?

Proof. If possible let there exists a point x ∈ X − A such that f (x) 6= g(x). Then there
exist disjoint open sets Uf (x) and Vg(x) containing f (x) and g(x) respectively. Then there
exist an open set U containing x in X such that f (U ) ⊂ Uf (x) and f (U ) ⊂ Ug(x) . But
this gives us a contradiction as f and g agree on A and U meets A non trivially.

6
CHAPTER 1

Topology

BY

Dr. Dibyendu De

Associate Professor
Department of Mathematics
University Of Kalyani
West Bengal, India
E-mail : dibyendude@gmail.com

1
Module-5: Homeomorphism

In studying group theory, metric spaces we have observed structure preserving map-
pings such as isomorphism, isometry. In this module we will discuss structure preserving
mappings of topological spaces.

Definition 1. A continuous map f : X → Y between topological spaces is said to be a


homeomorphism if there exists a continuous map g : Y → X such that g ◦ f = 1X and
f ◦ g = 1Y .

Clearly, here g is actually f −1 . So if O is an open set of X then the inverse image of


O under f −1 is the same as the image of O under the map f . The same thing happens
in case of closed sets. So we can define a homeomorphism in the following way.

Theorem 1. A mapping f : X → Y between two topological spaces is a homeomorphism


if and only if it is continuous and open or closed.

Proof. Since f is a homeomorphism there exists g : Y → X such that g ◦ f = 1X and


f ◦ g = 1Y . This means that f (U ) = g −1 (U ) which is open.

Example 1. A function f : R → R defined by f (x) = ax + b, where a 6= 0 is a


homeomorphism.
f (y)−b
Proof. g(y) = a
is the inverse mapping. Continuity is clear.

Example 2. Now we can observe that any two same type of intervals are homeomorphic.
Without loss of generality let us consider open intervals (0, 1) and (3, 4). The mapping
f : (0, 1) → (3, 5), defined by f (x) = ax + b gives a homeomorphism. We have just to
choose a, b suitable real numbers.

2
Example 3. This example shows that R and (0, ∞) are homeomorphic.

Proof. Let f : R → (0, ∞) defined by f (x) = ex . The following diagram clearly shows
that f is a homeomorphism.

Example 4. This example shows that unit square and unit circle are homeomorphic.

Proof. The following diagram clearly shows the homeomorphism.

Example 5. Consider B n ⊂ Rn be the open unit ball, and if we define a map f : B n → Rn


by
x
f (x) = .
1 − |x|
Then this gives a homeomorphism from B n to Rn . An easy computation shows that
g : Rn → B n defined by
x
g(x) =
1 + |x|
is the continuous inverse of f .

Example 6. Next we present an example of a homeomorphism between sphere S 2 and


cube C = {(x, y, z) : max{x, y, z} = 1}. First we define : C → S 2 by the mapping
(x, y, z)
f ((x, y, z)) = p .
x2 + y 2 + z 2
Now g : S 2 → C defined by
(x, y, z)
g((x, y, z)) =
max{|x|, |y|, |z|}
gives the inverse of f .

3
Example 7. The mapping p : [0, 1) → S 1 defined by p(x) = e2πix is a continuous
bijection, which is not a homeomorphism.

Let S n denote the n-dimensional sphere {x ∈ Rn+1 : kxk = 1} taken with the subspace
topology. We claim that removing a single point from S n gives a space homeomorphic
to Rn . Which point we remove is irrelevant because we can rotate any point of S n into
any other; for convenience we choose to remove the point N = (0, 0, . . . , 0, 1). Now the
set of points of Rn+1 which have zero as their final coordinate, when given the induced

topology, is clearly homeomorphic to Rn .

We define a function h : S n \ {N } → Rn , called stereographic projection, as follows.


For any x ∈ S n \ {N }, let h(x) be the point of intersection of Rn and the straight line
determined by x and N . Clearly h is bijective. Let O be an open set in Rn , we construct
a new set U in S n whose points are the points of intersection of the straight line segments
which start at N and pass through points of O, except the point N (See the diagram in
the next page). Then O is open in S n . But h−1 (O) is precisely the set U . Therefore
h−1 (O) is open in S n \ {N }. This establishes the continuity of h and a precisely similar
argument deals with h−1 . Therefore h is a homeomorphism.

4
Definition 2. A property say P of a topological space is said to be topological property if
when ever two topological spaces X and Y are homeomorphic and one posses the property
then the other will posses the property.

Example 8. Let X = R and let d be the usual metric on R. Let Y = (0, 1) (the open
interval) and let ρ be the usual metric on (0, 1). Then X and Y are homeomorphic as
topological spaces, but (X, d) is complete and (Y, ρ) is not. So the completeness is not a
topological property.

Theorem 2. Both T1 ness and Hausdorffness are topological property.

Proof. Let X and Y be two topological spaces, f : X → Y be a homeomorphism and X


be T1 . Now if F be a finite set of Y then |f −1 (F )| = |F | and hence f −1 (F ) is closed and
f being closed F = f (f −1 (F )) is also closed. Hence Y is T1 .

Let f : X → Y be a homeomorphism, X be Hausdorff and let x 6= y be two distinct


points in Y . Choose u and v be unique preimages of x and y respectively. Then there exist
disjoint open sets U and V in X containing x and y respectively. Then using openness
of f we get disjoint open sets f (U ) and f (V ) containing x 6= y respectively. This proves
that proved that Y is Hausdorff.

The following result will be needed in the study of manifold. By a disc we shall mean
any space homeomorphic to the closed unit disc D in R2 . If A is a disc, and if h : A → D
is a homeomorphism, then h−1 (S 1 ) is called the boundary of A and is written ∂A.

5
Theorem 3. Any homeomorphism from the boundary of a disc to itself can be extended
to a homeomorphism of the whole disc.

Proof. Let A be a disc and choose a homeomorphism h : A → D. Given a homeomor-


phism g : ∂A → ∂A we can easily extend hgh−1 : S 1 → S 1 to a homeomorphism of all
 
−1 x
of D as follows. Send 0 to 0, and if x ∈ D \ {0} send x to the point kxkhgh kxk
.
In other words extend conically. If we call this extension f , then h−1 f h extends g to a
homeomorphism of all of A as required.

We have already discussed about finite product topology. One can observe easily that
each projection map is continuous. Let us examine the following example.

Example 9. If we view points in the unit circle S1 in R2 as angles θ, then polar coordi-
nates give a homeomorphism f : S1 ×(0, ∞) → R2 \{0} defined by f (θ, r) = (rcosθ, rsinθ).
This is one-to-one and onto since each point in R2 , other than the origin has unique polar
coordinates (θ, r). To see that f is a homeomorphism, just observe that it takes a basic
open set U × V , (where U is an open interval (θ0 , θ1 ) and V is an open interval (r0 , r1 ))
to an open polar rectangle and such rectangles form a basis for the topology on R2 \ {0},
as a subspace of R2 . By restricting f to a product S1 × [a, b] for 0 < a < b we obtain
a homeomorphism from this product to a closed annulus in R2 , the region between two
concentric circles.

Example 10. A product S1 ×[1, 2] is homeomorphic to a cylinder as well as to an annulus.


If we use cylindrical coordinates (r, θ, z) in R3 then a cylinder is specified by taking r to
be a constant 1, letting range over the circle S1 , and restricting z to an interval [1, 2].

Consider the spaces Con-

sider [0, 1] and R. This following Theorem shows that we cann’t hope a Theorem like
Cantor Bernstein Theorem.

6
CHAPTER 1

Topology

BY

Dr. Dibyendu De

Associate Professor
Department of Mathematics
University Of Kalyani
West Bengal, India
E-mail : dibyendude@gmail.com

1
Module-6: Product topology

We have already known how to form the finite product of topological spaces. Now we
wish to generalize this for arbitrary product.

Definition 1. Let {Xi : i ∈ I} be an arbitrary collection of nonempty sets. Then the


Q S
product of these sets is defined by i∈I Xi = {x : I → i∈N Xi such that x(i) ∈ Xi }.

Wer want to mention here that the Axiom of Choice is precisely the statement that
Q
if Xi 6= ∅ for all i, then i∈I Xi 6= ∅. Idea behind it is that if we can make well ordered
the set I then arbitrary product of non empty sets is non empty.
Next we want to give a topology on the infinite product which will come from the topol-
ogy of factor spaces. Let us first examine that what happens if we proceed like finite case.

Proposition 1. Let X1 , X2 , . . . , Xn be finitely many topological spaces and X be an an-


other topological space. Then a function f : X → ni=1 Xi is continuous iff πi ◦ f is
Q

continuous for each i.

Proof. If f is continuous then πi ◦ f , being composition of two continuous functions is


continuous.
For the converse let πi ◦ f is continuous for each i and let U be a basic open set in
the product. Then there exists open set Ui in each Xi such that U = ni=1 Ui . So
Q

f −1 ( ni=1 Ui ) = {a ∈ A : f (a) ∈ Ui ∀i} = f −1 ( ni=1 Ui ).


Q T

The above mentioned property is so natural, one should expect that in any successful
generalization of finite product this property will carry over.
Let us try to proceed as the finite case and see what happens. Given topological spaces
{Xn : n ∈ N} let
Y
B={ Ui : Ui is an open set in Xi }.
n∈N

2
Then it can be easily shown that the collection B is a basis. One might think to consider
this collection as a basis for infinite product. But this definition will not be a successful
generalization as can be seen from the following example.

Example 1. Let us consider R with usual topology and consider the topology constructed
as above on RN . Consider the mapping f : R → RN . If the above definition is a successful
generalization then f : R → RN defined by f (x) = (x, x, . . .) should be continuous as the
finite case. But if we consider the open set
1 1 1 1
U = (−1, 1) × (− , ) × . . . (− , ) . . . .
2 2 n n
Then f −1 (U ) = 1 1
T
n∈N (− n , n ) = {0}, which is not open. Hence f is not continuous

This example shows that the straight forward generalization from finite product to
infinite product is not a successful generalization.

Definition 2. If I is a nonempty set and for each i in I we have a topological space Xi ,


then for each i ∈ I, we define the the following collection

Si = {πi−1 (Ui ) : i ∈ I}
S
and S = i∈I Si .
Q
It can be easily verified that S is a subbase for some topology on i∈I Xi . The topol-
Q
ogy generated by S will be called product topology on i∈I Xi .

If we view in the coordinate wise pattern, a typical subbasic element will be of the
form
Y
Ui , where Ui = Xi for all i except a single coordinate .
i∈I

As we know that we get basis from subbasis taking all possible finite intersection, here
a typical basis element will be of the form i∈F πi−1 (Ui ) where F is a finite subset of I
T

and Ui is an open set in Xi . Coordinate wise a typical subbasic element will be of the
form
Y
Ui , where Ui = Xi for all except finitely many i’s .
i∈I

As usual taking basic open sets in the factor spaces in place of open sets also give rise to
the product topology.

3
The following Theorem is a very useful one.
Q
Proposition 2. Let {Xi : i ∈ I} be a collection of topological spaces and i∈I Xi be
product space.
Then projection πi onto Xi is continuous and open map for each i.

Proof. Write basic open sets of the product in coordinate form. The proof will be clear.

The projection maps need not be closed. In fact The projection map π : R2 → R is
not closed map. Consider {(x, x1 ) : x > 0}.

Proposition 3. Let {Xi : i ∈ I} be a collection of topological spaces indexed by I; let


Q
Ai ⊂ Xi for each i. If i∈I Xi is given the product topology, then
Y Y
Ai = Ai .
i∈I i∈I
Q Q
Proof. Let x = (xi )i∈I be a point in i∈I Ai and i∈I Ui be a basis element for the prod-
uct topology containing x. Since xi ∈ Ai , we can choose a point yi ∈ Ui ∩ Ai for each i.
Q Q
Then y = (yi )i∈I belongs to both U and i∈I Ai . Therefore x ∈ i∈I Ai .

Q
For the converse let x = (xi )i∈I be a point in i∈I Ai . We claim that for any index
i ∈ I xi ∈ Ai . Let Vj be an arbitrary open set of Xj containing xj . Since πj−1 (Vj ) is
Q Q
open in i∈I Xi in product topology, it contains a point y = (yi )i∈I in i∈I Ai . Then yj
belongs to Vj ∩ Aj . It follows that xj ∈ Aj .

The following Theorem shows that product topology is the successful generalization
of finite product topology.
Q
Theorem 1. Let f : X → a∈I Xa be given by the equation f (x) = (fa (x))a∈I . Where
Q
fa : X → Xa for each a. Let a∈I Xa have the product topology. Then the function f is
continuous if and only if each function fa is continuous.
Q
Proof. Let f : X → a∈I Xa be continuous. The function fb is in fact equal with the
composition map πb ◦ f . So being the composite of two continuous functions, it is con-
tinuous.

4
Converse follows by the equality f −1 (πb−1 (Ub )) = fb−1 (Ub ).

Definition 3. Let X be a set and {Xi | i ∈ I} be a collection of topological spaces with


fi : X → Xi , for each i ∈ I. The weak topology induced on X by the collection {fi | i ∈ I}
of functions is the smallest topology on X making each fi continuous.

Proposition 4. Let {Xi : i ∈ I} be a collection of topological spaces. If Z is a topological


space and g : Z → X is a function, then g is continuous if and only if fi ◦ g : Z → Xi is
continuous for every i in I.

Proof. If g is continuous, then fi ◦ g is the composition of two continuous functions and


is therefore continuous. Assume each fi ◦ g is continuous, and let W be the topology of
Z. To show that g is continuous, we need only show that g −1 (S) ∈ W for every S in S,
where S is the topology of X. But if i ∈ I and G ∈ Ti , where Ti is the topology of Xi ,
then g −1 ◦ fi−1 = (fi ◦ g)−1 (G), and this belongs to W since fi ◦ g is continuous.

Q
Theorem 2. The product topology is the weak topology on Xi for which each projection
πj is continuous.

Proof. If τ is any topology on the product in which each projection is continuous, then
for each j, if Gj is open in Xj , so that πj−1 (Gj ) ∈ τ . Consequently, the members of
a subbasis for the product topology all belong to τ and hence the product topology is
contained in τ .

Q
Theorem 3. Let Aα be a subspace of Xα for each α ∈ J. Then α∈J Aα is a subspace
Q
of α∈J Xα .

Q Q
Proof. Let us consider a basic open set α∈J Uα of α∈J Aα . Then there exists a finite
set {α1 , α2 , . . . , αk } ⊂ J such that for all α ∈ J \{α1 , α2 , . . . , αk } we have Uα = Aα . Since
each Aα is a subspace of Xα there exist open sets Oα in Xα such that Uα = Oα ∩Aα for each
Q
α ∈ J. Let us consider the set α∈J Oα where Oα = Xα for all α ∈ J \ {α1 , α2 , . . . , αk }.
Q T Q  Q
Then we can see that α∈J Oα α∈J Aα = α∈J Uα . This proves the result.

We end this section with the following definition, which will be required in future.

5
Definition 4. Let {fα : X → Xα : α ∈ I} be a collection of continuous maps. Then the
Q
evaluation map e : X → Xα induced by the collection {fα : X → Xα } is defined by
[e(x)]α = fα (x) for each x ∈ X.

{fα : X → Xα : α ∈ I} is said to separate points in X if whenever x 6= y in X,


then for some α ∈ I, fα (x) 6= fα (y). We will use this technique in proving Uryshon
Metrization Theorem in future.

6
CHAPTER 2

Countability axioms

BY

Dr. Dibyendu De

Associate Professor
Department of Mathematics
University Of Kalyani
West Bengal, India
E-mail : dibyendude@gmail.com

1
Module-1: Metrizable spaces

One of the most important topological spaces are metrizable topological spaces. Topol-
ogy of such spaces has to be defined in terms of a metric on the set. Topologies given in
this way lie at the heart of modern analysis. In this section, we shall define the metric
topology and shall give a number of examples.

Definition 1. Let X be a nonempty set, A metric on X is a function d : X × X :→ R


which satisfies the following properties:
1) d(x, y) ≥ 0 for all x, y ∈ X, and d(x, y) = 0 iff x = y,
2) d(x, y) = d(y, x) for all x, y ∈ X,
3) d(x, y) ≤ d(x, z) + d(z, y) for all x, y, z ∈ X.

Given a metric d on X, and r > 0 we define Bd (x, r) = {y ∈ X : d(x, y) < r}.

Proposition 1. If d is a metric on the set X, then the collection of all r-balls Bd (x, r),
for x ∈ X and r > 0, is a basis for a topology on X, called the metric topology induced
by d.

Proof. The first condition for a basis is trivial, since x ∈ Bd (x, r) for any r > 0.
First we claim that if y is a point of the basis element Bd (x, r), then there is a basis element
Bd (y, s) centered at y that is contained in Bd (x, r). Define s to be the positive number
r − d(x, y). Then Bd (y, s) ⊂ Bd (x, r). In fact if z ∈ Bd (y, s), then d(y, z) < r − d(x, y)
from which we conclude that

d(x, z) ≤ d(x, y) + d(y, z) < r.

Now we shall be able to prove the second condition for a basis, let B1 and B2 be two
basis elements and let y ∈ B1 ∩ B2 . We have just shown that, we can choose positive
numbers r1 and r2 so that B(y, r1 ) ⊂ B1 and B(y, r2 ) ⊂ B2 . Letting 0 < r < min{r1 , r2 }
we conclude that B(y, r) ⊂ B1 ∩ B2 .

2
Now we can introduce the following definition.

Definition 2. If X is a topological space, X is said to be metrizable if there exists a


metric d on the set X that induces the topology of X. A metric space is a metrizable
space X together with a specific metric d that gives the topology of X.

In the following discussions we shall provide various examples of metrizable spaces.


A very important topological space is discrete space generated by discrete metric.

Definition 3. Let X be a nonempty set and d be a metric defined on X as follows :

d(x, y) = 1 iff x 6= y and d(x, y) = 0 iff x = y.

d is called discrete metric.

Then B(x, r) = {x} if r < 1 and B(x, r) = X if r ≥ 1. Therefore in the topological


space generated by the discrete metric d every single-toned set is open and therefore
discrete space.

Boundedness is not defined in arbitrary topological space. But it can be defined in


metrizable spaces. So the natural question is that if two metrizable spaces are homeo-
morphic and one is bounded then wheather the other is also bounded. In the following
we shall show that this is not the case in general.

Proposition 2. Let X be a metric space with metric d. Define ρ : X × X → R by the


equation ρ(x, y) = min{d(x, y), 1}. Then ρ is a metric that induces the same topology as
d.

Proof. In the proof first we show that ρ is in fact a metric and then we show that these
two metrics generate same topology.
The first two conditions are obvious. To check the triangle inequality if ρ(x, y) or
ρ(y, z) are greater or equal to 1 then ρ(x, z) ≤ ρ(x, y) + ρ(y, z).
So let both ρ(x, y) or ρ(y, z) are strictly less than 1. Then we have

d(x, z) ≤ d(x, y) + d(y, z) = ρ(x, y) + ρ(y, z).

Since ρ(x, z) ≤ d(x, z), the triangle inequality follows for ρ.


Now we note that in any metric space, the collection of r -balls with r < 1 forms a basis
for the metric topology. It follows that d and ρ induce the same topology on X, because
the collections of r-balls with r < 1 under these two metrics are the same collection.

3
Definition 4. Let us define d on Rn as follows :
p
d(x, y) = (x1 − y1 )2 + (x2 − y2 )2 + . . . + (xn − yn )2 .

This is a metric called Euclidean metric

Definition 5. The following metric ρ on Rn is known as square metric :

ρ(x, y) = max{(x1 − y1 ), (x2 − y2 ), . . . , (xn − yn )}.

In the following Theorem we show that two topologies on a set determined by metrics
can be compared in terms of open balls. This is a particular form of a Theorem we
already told for bases.

Lemma 1. Let d and ρ be two metrics on the set X; let τ1 and τ2 be two topologies
induced by d and ρ respectively. Then τ1 is finer than τ2 if and only if for each x in X
and each r > 0, there exists s > 0 such that

Bρ (x, s) ⊂ Bd (x, r).

Theorem 1. The topologies on Rn induced by the Euclidean metric d, the square metric
ρ and the product topology are equivalent.

Proof. Let x, y ∈ Rn . Then the following inequality can be seen easily :


ρ(x, y) = max{(x1 − y1 ), (x2 − y2 ), . . . + (xn − yn )}
p
≤ d(x, y) = (x1 − y1 )2 + (x2 − y2 )2 + . . . + (xn − yn )2 .

Also we have d(x, y) ≤ nρ(x, y).

Therefore from the first inequality we have

Bd (x, r) ⊂ Bρ (x, r)

for all x and r. Similarly, the second inequality shows that

r
Bρ (x, √ ) ⊂ Bd (x, r).
n

for all x and r. It follows that the topologies generated by these two metrics are the
same.
Next we show that the product topology is the same as that generated by the metric ρ.

4
First let B be a basic open set in the product topology, that means product of open
intevals and let
x ∈ B = (a1 , b1 ) × (a1 , b1 ) × . . . × (an , bn ).

Then for i there exist εi such that (xi − ε, xi + ε) ⊂ (ai , bi ). Let us choose ε =
min{ε1 , ε2 , . . . , εn }. Then Bρ (x, ε) ⊂ B. As a result, the ρ-topology is finer than the
product topology.

Conversely, let Bρ (x, ε) be a basis element for the ρ-topology. Given the element
y ∈ Bρ (x, ε), we need to find a basis element B for the product topology such that
y ∈ B ⊂ Bρ (x, ε). We just take B = ni=1 (yi − ε, Yi + ε).
Q

Next we try to investigate whether infinite product of metrizable spaces is metrizable.


We shall see that metrizability is only countable productive. Before proving the result let
us first see what happens if we try to extend the Euclidean metric in infinite case. For
instance, one can attempt to define a metric d on Rω by the equation
" ∞
#1/2
X
d(x, y) = (xi − yi )2 .
n=1

But this equation does not always make sense, for the series in question need not converge.
Similarly, one can attempt to generalize the square metric ρ to Rω by defining

ρ(x, y) = sup{|xi − yi |}.

If we replace the original metric d in a metric space (X, d) by its bounded counterpart
d(x, y) = min{d(x, y), 1}, then the sup metric makes sense; it gives a metric on countable
product, called the uniform metric. In this lecture let us concentrate on Rω .

Definition 6. Let us define a metric ρ on Rω by the equation

ρ(x, y) = sup{d(xn , yn ) : n ∈ ω},

where d is the standard bounded metric on R. It is easy to check that ρ is indeed a metric;
it is called the uniform metric on Rω , and the topology it induces is called the uniform
topology.

The relation between uniform topology and the product topology is the following:

5
Theorem 2. The uniform topology on Rω is finer than the product topology.
Q
Proof. Let n Un be a basic open set in the product topology containing a point x =
(xn )n . Let n1 , n2 , . . . , nn be the indices for which Un 6= R. Then for Then for each i,
choose ε > 0 so that the ε-balls centered at xni in the d metric is contained in Uni . This
we can do because Uni is open in R. Let ε = min{ε1 , ε2 , . . . , εn }; then the ε-ball centered
at x in the ρ metric is contained in n Un . For if z is a point of Rω such that ρ(x, z) < ε,
Q
Q
then d(xn , zn ) < ε for all n, so that z ∈ n Un . It follows that the uniform topology is
finer than the product topology.

Finally we prove that countably infinite product of metrizable spaces is metrizable.


We will stick to the case Rω . The proof for the general metric spaces is quite similar and
left as exercise.

Theorem 3. Let d(a, b) = min{|a − b|, 1} be the standard bounded metric on R. If x and
y are two points of Rω , define
 
d(xi , yi )
ρ(x, y) = sup .
i

Then ρ(x, y) is a metric that induces the product topology on Rω .

Proof. Let us first proof that ρ is a metric. First two conditions are clear. Let us prove
the triangle inequality.

d(xi , zi ) d(xi , yi ) d(yi , zi )


≤ + = ρ(x, y) + ρ(y, z).
i i i

Therefore  
d(xi , zi )
sup ≤ ρ(x, y) + ρ(y, z).
i
Now let U be open in the metric topology and let x ∈ U . We shall find an open set
V in the product topology such that x ∈ V ⊂ U .
Choose an r-ball Bρ (x, r) lying in U . Then choose a positive integer N such that
1
N
< r. Finally, let V = (x1 − r, x1 + r) × (x2 − r, x2 + r) × . . . × (xN − r, xN + r) × R × R . . .
be a basic open set in the product topology. We claim that V ⊂ Bρ (x, r): Given any
y ∈ Rω ,
d(xi , zi ) 1
≤ for i ≥ N.
i N

6
Therefore  
d(x1 , z1 ) d(x2 , z2 ) d(xN , zN ) 1
ρ(x, y) ≤ max , ,..., , .
1 2 N N
If y is in V , this expression is less than r, so that V ⊂ Bρ (x, r) as desired.
Q
Conversely let n Un be a basic open set in the product topology. Let a1 , a2 , . . . , an be
the indices for which Ua 6= R. Given x ∈ U , choose an interval (xi −εi , xi +εi ) in R lying in
Ui for i = a1 , a2 , . . . , an ; choose each εi ≤ 1. Then define ε = min{ εii : i = a1 , a2 , . . . , an }.
We claim that x ∈ Bρ (x, ε) ⊂ U .
Let y be a point of Bρ (x, ε). Then for all i,

d(xi , yi )
≤ ρ(x, y) < ε.
i
εi
Q
Finally i = a1 , a2 , . . . , an , ε ≤ i
so that d(xi , yi ) ≤ εi ≤ 1. Therefore y ∈ n Un .

7
CHAPTER 3

Separation axioms

BY

Dr. Dibyendu De

Associate Professor
Department of Mathematics
University Of Kalyani
West Bengal, India
E-mail : dibyendude@gmail.com

1
Module-2: Separation Axioms,
Normality

We start this module by showing that normality is closed heredity property.

Theorem 1. Closed subspace of a normal spaces is normal.

Proof. Let X be a normal space and Y be a closed subspace of X. Then any two disjoint
closed sets in Y is also closed in X. Then normality of X can be used to prove the
normality of Y .

In the following example we show that normality is not even finitely productive prop-
erty.

Example 1. In a previous example we have observed that Rl is normal. Now we shall


show that Rl 2 is not normal. Which will further produce an example of a regular space
which is not normal, as product of regular spaces is regular and Rl being normal is regular.

If possible let Rl 2 be normal. Let L be the subspace of Rl 2 , consisting of all points of


the form (x, −x). Then L is closed in Rl 2 and L has the discrete topology.

Example 2. Hence every subset A of L, being closed in L, is closed in Rl 2 . Because L\A


is also closed in Rl 2 , this means that for every nonempty proper subset A of L, one can
find disjoint open sets UA and VA containing A and L − A, respectively. Let D = Q × Q.
Then D is dense in Rl 2 . We define a map f that assigns, to each subset of the line L, a

2
subset of the set D, as follows :

f (A) = D ∩ UA ∅ A L
f (∅) = ∅ .
f (L) = D

We claim that f : P(L) → P(D) is injective.

Example 3. Let A be a proper nonempty subset of L. Then f (A) = D ∩ UA is neither


empty as UA is open and D is dense in Rl 2 , nor all of D since D ∩ VA is nonempty. It re-
mains to show that if B is another proper nonempty subset of L, then f (A) 6= f (B). One
of the sets A, B contains a point not in the other; suppose that x ∈ A and x 6∈ B. Then
x ∈ L−B, so that x ∈ UA ∩VB ; since the latter set is open and nonempty, it must contain
points of D. These points belong to UA and not to UB ; therefore, D ∩ UA 6= D ∩ UB , as
desired. Thus f is injective.

Next as D is countable and L has cardinality of the continuum, there exist a bijection
ϕ : P(D) → L. Then ϕ ◦ f : P(L) → L is an injective mapping, which is a contradiction.

Theorem 2. Every ordered space is normal.

Proof. We assert that every interval of the form (x, y] is open in a well-ordered set X. In
fact if X has a largest element and y is that element, (x, y] is just a basis element about
y. If y is not the largest element of X, then (x, y] equals the open set (x, y) , where y
is the immediate successor of y. Now let A and B be disjoint closed sets in X, assume
for the moment that neither A nor B contains the smallest element a0 of X. For each
a ∈ A, there exists a basis element about a disjoint from B it contains some interval of
the form (x, a], as a is not the smallest element of X. Choose, for each a ∈ A, such an
interval (xa , a] disjoint from B.
Similarly, for each b ∈ B, choose an interval (yb , b] disjoint from A. The sets
[ [
U= (xa , a] and V = (yb , b].
a∈A b∈B

are open sets containing A and B, respectively; we claim that they are disjoint. In
fact if z ∈ U ∩ V . Then z ∈ (xa , a] ∩ (yb , b] for some a ∈ A and some b ∈ B. Assume that
a < b. Then if a ≤ yb , the two intervals are disjoint, while if a > yb , we have a ∈ (yb , b],
contrary to the fact that (yb , b] is disjoint from A. A similar contradiction occurs if b < a.

3
Finally, assume that A and B are disjoint closed sets in X and A contains the smallest
element a0 of X. The set {a0 } is both open and closed in X. Then by above there exist
disjoint open sets U and V containing the closed sets A − {a0 } and B, respectively. Then
U ∪ {a0 } and V are disjoint open sets containing A and B, respectively.

In the coming discussions we shall see that Linearly ordered topological spaces satisfies
more stronger separation property called completely normal.
One of the useful properties of the set N of positive integers is the fact that each of its
nonempty subsets has a smallest element. This property leads to the following definition.

Definition 1. A set X with an order relation ≤ is said to be well-ordered if every


nonempty subset of X has a smallest element.

Theorem 3. Every nonempty finite ordered set has the order type of a section {1, 2, . . . , n}
of N, so it is well-ordered.

Theorem 4. (Well-ordering theorem). If A is a set, there exists an order relation on A


that is a well-ordered.

Corollary 1. There exists an uncountable well-ordered set.

For any well ordered set X and given α ∈ X, let Sα let us denote the set Sα = {β ∈
X : β < α}. It is called the section of X by α.

Theorem 5. There exists a well-ordered set A having a largest element Ω, such that the
section SΩ of A by Ω is uncountable but every other section of A is countable.

In fact if B be an uncountable well-ordered and S be the well-ordered set {1, 2} × B


in the dictionary order; then some section of C is uncountable. Let Ω be the smallest
element of C for which the section of C by Ω is uncountable. Then let A consist of this
section along with the element Ω.
Note that SΩ is an uncountable well-ordered set every section of which is countable.
Its order type is in fact uniquely determined by this condition. We shall denote the
well-ordered set A = SΩ ∪ {Ω} by the symbol SΩ + 1.
The most useful property of the set SΩ for our purposes is expressed in the following
theorem:

4
Theorem 6. If A is a countable subset of SΩ , then A has an upper bound in SΩ .

With these few properties of ordinal numbers, which will be useful for us for further
discussions, we end this module.

5
CHAPTER 3

Countability axioms

BY

Dr. Dibyendu De

Associate Professor
Department of Mathematics
University Of Kalyani
West Bengal, India
E-mail : dibyendude@gmail.com

1
Module-1: Properties of normal
spaces

We have already proved that Rl × Rl is not normal. In the following we shall provide
another example to show product of normal spaces may not be normal. This example is
taken from J. R. Munkres.

Example 1. Consider the well-ordered set SΩ + 1, in the order topology, and consider
the subset SΩ , in the subspace topology (which is the same as the order topology). Both
spaces are normal, as linearly ordered topological spaces are normal. We wish to prove
that (SΩ + 1) × SΩ is not normal.
First, we consider the space (SΩ + 1) × (SΩ + 1), and its “diagonal” 4 = {(x, x) : x ∈
SΩ + 1}. Because SΩ + 1 is Hausdorff, 4 is closed in (SΩ + 1) × (SΩ + 1). If U and V
are disjoint neighborhoods of x and y, respectively, then U × V is a neighborhood of x × y
that does not intersect 4.
Therefore, in the subspace SΩ × (SΩ + 1), the set

A = 4 ∩ (SΩ × (SΩ + 1)) = 4 − {Ω × Ω}

is closed. Similarly the set B = SΩ × {Ω} is closed in SΩ × (SΩ + 1). The sets A and B
are disjoint.

Assuming there exist disjoint open sets U and V of SΩ × (SΩ + 1) containing A and
B, respectively, we shall derive a contradiction.
Given x ∈ SΩ , consider the vertical slice x × (SΩ + 1). We assert that there is some point
β with x < β < Ω such that (x, β) lies outside U . In fact if U is contained all points
(x, β) for x < β < Ω then the top point (x, Ω) of the slice would be a limit point of U ,

2
which it is not because V is an open set disjoint from U containing this top point
Choose β(x) to be such a point; just to be definite, let β(x) be the smallest element of
S( Ω) such that x < β(x) < Ω and (x, β(x)) lies outside U .
Let us define a sequence of points of S( Ω) as follows: Let x1 be any point of S( Ω). Let
x2 = β(x1 ), and in general,xn+1 = β(xn ). We have

x1 < x2 < . . . ,

because β(x) > x for all x.


The set {xn } is countable and therefore has an upper bound in S( Ω); let b ∈ SΩ be its
least upper bound. Because the sequence is increasing, it must converge to its least upper
bound; thus xn → b. But β(xn ) = xn + 1 so that β(xn ) −→ b also. Then

(xn , β(xn ) −→ (b, b)

in the product space. Now we have a contradiction, for the point (b, b) lies in the set
A, which is contained in the open set U ; and U contains none of the points (xn , β(xn ).

Now we shall enter to some deeper study of normal spaces. Two important Theorem to
note here are Urysohn Lemma ant Tietz extension lemma. The first one gives information
about rich source of continuous functions and the second one tells about extension of
continuous functions. Let us start with following Theorem

Theorem 1. Every regular space with a countable basis is normal.

Proof. Let X be a regular space with a countable basis say B = {Bn ; n ∈ N}. Let A and
B be disjoint closed subsets of X. The idea of the proof is to cover A with some basic open
sets and to cover B with some basic open sets and then to to eliminate the overlapped
parts. Each point x of A has a neighborhood U not intersecting B. Using regularity,
choose a neighborhood V of x whose closure lies in U ; we can do these operations taking
elements from B. By choosing such a basis element for each x in A, we construct a
countable covering of A by open sets whose closures do not intersect B. Since this
covering of A is countable, we can index it with the positive integers; let us denote it by
{Un }. As an immediate application of the above Theorem we can say that real line with
usual topology is normal.

3
Similarly, choose a countable collection {Vn } of open sets covering B, such that each
S S
set Vn is disjoint from A. The sets U = n Un and V = n Vn are open sets containing
A and B. Given n ∈ N, define
! !
[ [
Un0 = Un \ Vn and Vn0 = Vn \ Un .
n n

Then each Un0 and Vn0 are open sets. It is easy to observe that {Un0 } is a cover of A and
{Vn0 } is a cover of B.
The collection {Un0 } is a cover of A because each x in A belongs to Un for some n, and
x belongs to none of the sets V i . Similarly, the collection {Vn0 } covers B. Now let us set
[ [
U0 = Un0 and V 0 = Vn0 .
n n

It remains to show that U 0 and V 0 are disjoint. For if x ∈ U 0 ∩ V 0 then x ∈ Uj0 ∩ Vk0
for some j and k. Suppose that j < k. It follows from the definition of U 0 that x ∈ Uj ,
and since j ≤ k it follows from the definition of Vk0 that x 6∈ U j . A similar contradiction
arises if j ≥ k.

Quite same construction like above Theorem proves the following Theorem.

Theorem 2. Every regular Lindelöf space is normal.

Next we shall introduce another notion of separability. We know that normality is


not heredity property. But there are normal spaces all whose subspaces are normal. This
motivates the following definition.

Definition 1. A space X is said to be completely normal if every subspace of it is normal.

Definition 2. In a space X two sets A and B will be said separable if A ∩ B = ∅ and


A ∩ B = ∅.

Theorem 3. A space is completely normal iff every pair of separated subsets can be
separated by neighborhoods.

Proof. Suppose A and B is a pair of separated subsets of X. Then Y = X − (A ∩ B) is an


open subset of X that contains both A and B. AY ∩ B Y = Y ∩ A ∩ B = ∅. Thus, AY and
B Y can be separated by open neighborhoods in Y . Since Y is open, these neighborhoods
are also open in X .

4
Conversely take a set Y ⊂ X and two disjoint subsets A, B ⊂ Y closed in Y . AX ∩ B =
AX ∩ Y ∩ B = AY ∩ B = ∅ . Similarly, B X ∩ A = ∅. Therefore, A and B can be separated
by neighborhoods in X and their intersections with Y separate A and B in Y .

In the following we shall examine the completely normality of the some spaces.

Proposition 1. Every subspace of a completely normal space is completely normal.

Proof. Let Y be a subspace of a completely normal space X. Then any subspace of Y is


a subspace of the completely normal space X therefore, is normal. This means that Y is
completely normal.

Proposition 2. Every well-ordered set is Completely normal in the order topology.

Proof. If a set is well-ordered then every element has a successor and {(a, x]} is a basis
at x where a < x or a = −∞. Therefore, for a pair of separated sets we can cover
each set with such neighborhoods that do not intersect the other set. Moreover, the
neighborhoods belonging to one set do not intersect the neighborhoods belonging to the
other set.

Proposition 3. Rl is Completely normal.

Proof. Indeed, the proof of the fact that Rl is Completely normal, is extremely similar to
the previous example: both use the fact that there is a basis at x with sets of the form
[x, a). This shows that every point in one closed set has such a neighborhood that does
not intersect the other set. Then coverings by such basis sets are disjoint automatically.
So, we obtain the disjoint open neighborhoods for any pair of sets such that neither one
contains limit points of the other one.

Question 1. Does every metric space Completely normal?

Every subspace is metrizable as well, therefore, normal

Question 2. Does every regular space with a countable basis is completely normal?

Every regular second-countable space is normal. Also every its subspace is also regular
and second-countable. Therefore, every regular second-countable space is completely
normal.

Question 3. The product of two completely normal spaces.

Rl is completely normal, but (Rl )2 is not even normal.

5
CHAPTER 3

Separation Axioms

BY

Dr. Dibyendu De

Associate Professor
Department of Mathematics
University Of Kalyani
West Bengal, India
E-mail : dibyendude@gmail.com

1
Module-4: Urysohn’s Lemma

Now we shall gradually move to Urysohn Metrization Theorem. In this course we


shall require the famous Urysohn’s Lemma. We have already given a version of Urysohn’s
lemma for metric space. But that depends completely on the metric. In this module we
shall present a general version of Urysohn’s lemma.

Theorem 1 (Urysohn’s Lemma). A space X is normal if and only if for any two disjoint
closed sets E and F there exists a continuous functions f : X :→ R such that

f (E) = 0 and f (F ) = 0.

This is a deep Theorem, both from the point of view that its proof, which involves
really original idea, also from the point of view of its application.

Proof. Let X be a normal space. Set V = X \ F , an open set containing E. Then by


normality criteria there exists an open set U 1 such that
2

E ⊂ U 1 ⊂ U 1 ⊂ V.
2 2

By successive application of normality criteria gives that there exist open sets U 1 and U 3
4 4

such that

E ⊂ U 1 ⊂ U 1 ⊂ U 1 ⊂ U 1 ⊂ U 3 ⊂ U 3 ⊂ V.
4 4 2 2 4 4

Continuing this manner for each dyadic rational number r ∈ (0, 1), an open set Ur
such that

Ur ⊂ Us , 0 < r < s < 1,

2
E ⊂ Ur , 0 < r < 1,

Ur ⊂ V, 0 < r < 1.

With this information we shall now go to define the function f .


0 if x ∈ Ur for all r > 0


f (x) =
 sup{r : x 6∈ Ur }

Evidently 0 ≤ f ≤ 1, f = 0 on E and f = 1 on F . It suffices to show that f is


continuous.
Let x ∈ X. For convenience, we assume that 0 < f (x) < 1, the case f (x) = 0 and
f (x) = 1 are not difficult. Let ε > 0. Choose dyadic rational number 0 < r < s < 1 and

f (x) − ε < r < f (x) < f (x) + ε.

Then x 6∈ Ut for dyadic rational numbers between r and f (x), so that x 6∈ U r . On


the otherhand x ∈ Us . Hence W = Us \ U r is an open neighborhood of x. If y ∈ W , then
from the definition of f we see that r ≤ f (x) ≤ s. In particular, |f (x) − f (y)| < ε for
y ∈ W , so that f is continuous at x.

Recall that A is a Gδ set in a space X if A is the intersection of a countable collection


of open sets of X. In metric space all closed sets are Gδ sets. But this is not true in
general. In normal space we have the following Theorem.

Theorem 2. Let X be normal space. There exists a continuous function f : X → [0, 1]


such that f (x) = 0 for x ∈ A, and f (x) > 0 for x 6∈ A, if and only if A is a closed Gδ set
in X.

Proof. Suppose there exists a continuous function f : X → [0, 1] such that f (x) = 0
for x ∈ A, and f (x) > 0 for x 6∈ A. Then A = f −1 (0) must be closed. Now A =
T −1 1 1
nf (− n , n ) and each f −1 (− n1 , n1 ) is an open set. Hence A is a Gδ set also.
Conversely let A be a closed Gδ set. Then there exists a sequence (Un ) of open sets such
T
that A = n Un . Then for each n there exists a continuous function fn which vanishes

3
on A and equal to 1 on X − Un . Now take
X |fn |
f= .
n
2n

Then clearly f is continuous and serves our purpose.

Now we are in a position to prove the coveted Urysohn Metrization Theorem.

Theorem 3. Every regular second countable space is metrizable.

Proof. We know that any regular second countable space X is normal so that we can
invoke Urysoh’s lemma. Let B = {Bn : n ∈ N} be a countable base. Then for any
m, n ∈ N if Bm ⊂ Bn there exists some f ∈ C ∗ (X) such that f (Bm ) = 0 and f (Bn ) = 1.
In this way we get a countable collention of functions say {fn : n ∈ N} which does the
above job. We define e : X → RN to be πn (e(x)) = fn (x). We shall prove that e is an
embedding. Continuity follows from the fact that each factor map fn is continuous. If
x 6= y then there exists Bm Bn ∈ B such that x ∈ Bm , y 6∈ Bn and Bm ⊂ Bn . So there
exists some fk such that fk (x) = 0 and fk (y) = 1.
It’s remain to show that e : X → e(X) is open. Let U be open in X, then we need
to show that set e(U ) is open in e(X). Let v ∈ e(U ). We have to find an open set W
of e(X) such that v ∈ W ⊂ e(U ). Let u be the point of U such that e(u) = v. Choose
an index N for which fN (u) > 0 and fN (X − U ) = {0}. Take the open ray (0, +∞) in
−1
R, and let V be the open set πN ((0, +∞)) of RN . We claim that v ∈ W ⊂ e(U ) First,
v ∈ W because πN (u) = πN (e(u)) = fN (u) > 0. Second, W ⊂ e(U ). For if z ∈ W , then
z = e(x) for some x ∈ X, and πN (z) ∈ (0, +∞). Since πN (z) = πN (e(x)) = fN (x), and
fN vanishes outside U the point x must be in U . Then z = e(x) is in e(U ) as desired.
Thus e is an imbedding of X in RN .

Definition 1. If E and F be two disjoint closed sets in a space X and there exists a
continuous functions f : X :→ R such that f (E) = 0 and f (F ) = 0. We say that E and
F can be separated by a continuous function.

The Urysohn lemma says that if every pair of disjoint closed sets in X can be separated
by disjoint open sets, then each such pair can be separated by a continuous function. The
converse is trivial, for if
f (E) = 0 and f (F ) = 0.

4
is the function, then f −1 [0, 13 ) and and f −1 ( 32 , 0] are disjoint open sets containing A and
B, respectively.
This fact leads to the analogue question for regular spaces, that is whether regularity
is equivalent with the fact that points and closed sets can be separated by continuous
functions. Unfortunately this does not hold in general, an example will be provided latter.
This fact leads to the following definition.

Definition 2. A space X is said to be Completely Regular if all finite sets are closed and
for any x ∈ X and a closed sets F not containing x, there exists a continuous functions
f : X :→ R such that

f (x) = 0 and f (F ) = 0.

This is quite clear that Completely regular spaces are regular and normal spaces are
Completely Regular.

Theorem 4. A subspace of a completely regular space is completely regular.

Proof. Let X be a completely regular, let Y be a subspace of X. Let x be a point of Y ,


and let K be a closed set of Y disjoint from x. We choose a closed set in X such that
K = H ∩ Y . Therefore x ∈
/ H. Since X is completely regular, we can choose a continuous
function f : X → R such that f (x) = 1 and f (H) = 0. The restriction of f to Y is
desired continuous function on Y .

Theorem 5. Product of completely regular spaces is completely regular.


Q
Proof. Let X = Xα be a product of completely regular spaces. Let x = (xα ) be a
Q
point of X and let U be a open set of X containing x. Then U = Uα , where each
Uα is open in Xα and Uα = Xα except for finitely many α’s, say, α1 , α2 , . . , αk . Then
for each αi we can choose continuous function fi : X → R such that fαi (xαi ) = 1 and
fαi (Xαi \ Uαi ) = 0. Now let us set φαi = fαi ◦ παi . Then each φαi is continuous from
X to R. If we set φ = φα1 .φα2 . . . .φαk , then it is easy to observe that φ(x) = 1 and
φ(X \ U ) = 0.

5
CHAPTER 3

Separation axioms

BY

Dr. Dibyendu De

Associate Professor
Department of Mathematics
University Of Kalyani
West Bengal, India
E-mail : dibyendude@gmail.com

1
Module-5: Tietze Extension
Theorem

Let X be the closed unit interval and A = (0, 1). If we consider the function by
1
f (x) = x
defined on (0, 1), then f cannot be extended to the closed interval [0, 1]. In fact
any continuous function defined on [0, 1] must be bounded. Tietze extension theorem is
one of the immediate consequence of the Urysohn lemma which deals with the problem
of extending a continuous real-valued function that is defined on a subspace of a space
X to a continuous function defined on all of X. This theorem is important in many of
the applications of topology.

Theorem 1. Let X be a normal space; let C be a closed subspace of X. Any continuous


map of C into R may be extended to a continuous map of all of X into R.

Proof. The idea of the proof is approximation. We shall construct a sequence of continu-
ous functions defined on the entire space X, such that the sequence converges uniformly,
and such that the restriction of each function to C approximates f . Then the limit
function will be continuous, and its restriction to C will equal f .
Let X be a normal space, C a closed subset, and f : C → R a continuous map. Let
us first consider the bounded case, i.e. say |f (x)| ≤ M for all x ∈ C.
M
Let A1 = {x ∈ C : f (x) ≥ 3
} and B1 = {x ∈ C : f (x) ≥ − M3 }. Then A1 and B1
are obviously disjoint, and they are both closed subsets of C. But C is closed in X, and
therefore A1 and B1 must be closed in X. By Urysohn’s lemma we can find a continuous
map g1 : X → [− M3 , M3 ] which takes the value M
3
on A1 and − M3 on B1 and which takes
values in (− M3 , M3 ) on X − (A1 ∪ B1 ). Notice that |f (x) − g1 (x)| ≤ 2M
3
on C.
Now consider the function f − g1 and let A2 consist of those points of C for which
2M
|f (x)−g1 (x)| ≥ 9
, and B2 those points for which |f (x)−g1 (x)| ≤ − 2M
9
. Again applying

2
Uryshon’s lemma a second time we can find a map g2 : X → [− 2M
9
, 2M
9
] which takes the
2M
value 9
on A2 and − 2M
9
on B2 , and values in (− 2M
9
, 2M
9
) on the remaining points of X.
4M
If we compute f (x) − g1 (x) − g2 (x), we see that |f (x) − g1 (x) − g2 (x)| ≤ 9
on C.
n−1 M n−1 M
By repeating this process we can construct a sequence of maps gn : X → [− 2 3n
,2 9
]
which satisfy:
2n M
(a) |f (x) − g1 (x) − g2 (x) − . . . gn (x)| ≤ 3n
on C; and
2n−1 M
(b) |gn (x)| < 3n
on X − C.

We now define

X
g(x) = gn (x)
n=1

for all x in X. The convergency of the infinite series follows from the comparison
theorem of calculus, it converges by comparison with the geometric series

∞  n−1
1X 2
.
3 n=1 3

To show that g is continuous, we must show that the sequence sn of partial sums converges
to g uniformly. This fact follows at once from the “Weierstrass M-test”.

Finally we show that g(x) = f (x) for x ∈ C. g(x) is by definition the limit of the
infinite sequence sn (x) of partial sums. Since

n  n
X 2
|f (x) − gi (x)| = |f (x) − sn (x)| ≤
i=1
3

for all x in C, it follows that sn (x) → f (x) for all x ∈ C. Therefore, f and g agree on C.

If |g(x)| is bounded then |g(x)| = ∞


P P∞ 2n−1
n=1 |gn (x)| ≤ n=1 M 3n = M , and |g(x)| is

strictly less than M on X − C by (b).


Let f : C → R be arbitrary continuous function. Let us choose a homeomorphism h from
the real line to the interval (−1, 1) and consider the composition h ◦ f , which is bounded.
Therefore by the above argument we can extend it to a continuous real-valued function
g on X, all of whose values lie strictly between -1 and 1. So the composition h−1 ◦ g is
well defined, and by construction it extends f over X. This completes the proof.

3
The following corollary can be obtained easily.

Corollary 1. Let X be a normal space; let C be a closed subspace of X. Then Any


continuous map of C into the closed interval [a, b] may be extended to a continuous map
of all of X into [a, b].

Proof. Proof follows from the proof of main Theorem.

Though we have used Urysohn Lemma in the proof Tietz extension Theorem, but if
a space satisfies Tietz extension Theorem, then it satisfies Urysohn Lemma.

Proposition 1. Show that the Tietze extension theorem implies the Urysohn lemma.

Proof. We need to use pasting lemma and then to apply Tietze extension theorem.

Definition 1. A space Y is said to have the universal extension property if for any given
normal space X, any closed subset A of X, and any continuous function f : A → Y ,
there exists an extension of f to a continuous map of X into Y .

Proposition 2. Prove that RJ has the universal extension property.

Proof. Consider a normal space X, a closed subset A of X, and a continuous function


f : A → Y . The for each i ∈ J, fi = πi ◦ f : A → R is continuous and hence has a
continuous extension say gi over X. Then g = (gi )i ∈ RJ is a continuous extension of f
over X.

Definition 2. Let X be a topological space. If Y is a subspace of Z, we say that Y is a


retract of X if there is a continuous map r : X → Y such that r(y) = y for each y ∈ Y .

It is easy to observe that if X is Hausdorff then any retract Y of X is closed.

x
Example 1. S1 is a retract of R2 −{0}. The map r : R2 −{0} → S1 defined by r(x) = kxk

is a retraction.

Example 2. If Y is homeomorphic to a retract of RJ , then Y has the universal extension


property.

Definition 3. Let X be a normal space. Then X is said to be an absolute retract if for


any normal space Y and any closed subspace Y0 of Y , homeomorphic to X, the space Y0
is a retract of Y .

4
Proposition 3. If X has the universal extension property, then X is an absolute retract.

Proof. Let Y be a normal space, Y0 be a closed subspace of Y homeomorphic to X. Let


f : Y0 → X be the homeomorphism. Since X has the universal extension property there
exists a continuous extension say g : Y → X of f : Y0 → X. Then r = f −1 ◦ g is the
required retraction.

You might also like