Sei sulla pagina 1di 131

Topology

Stanislav Jabuka
Notes for Math 440/640
University of Nevada, Reno

Contents
Chapter 1. Continuity and convergence in Euclidean spaces
1.1. Continuity in Euclidean space
1.2. Open and closed subsets of Euclidean space
1.3. Continuity and convergence in terms of open and closed sets
1.4. Some properties of open and closed sets in Euclidean space
1.5. Exercises

5
5
8
9
12
13

Chapter 2. Topological spaces


2.1. Denition of a topological space
2.2. Examples of topological spaces
2.3. Properties of open and closed sets
2.4. Bases and subbases of a topology
2.5. Exercises

15
15
16
25
28
33

Chapter 3. Continuous functions and convergent sequences


3.1. Continuous functions
3.2. Convergent sequences
3.3. Uniform convergence of functions
3.4. Space lling curves
3.5. Exercises

37
37
45
50
51
56

Chapter 4. Separation axioms


4.1. Degrees of separation
4.2. Examples
4.3. Topological invariants
4.4. A rst application of topological invariants
4.5. The Urysohn Lemma
4.6. Exercises

61
61
63
65
67
70
76

Chapter 5. Product spaces


5.1. Finite products
5.2. Innite products
5.3. Exercises

79
79
85
88

Chapter 6. Compactness
6.1. Denition and rst examples

91
91
3

CONTENTS

6.2.
6.3.
6.4.
6.5.
6.6.

Compactness for metric spaces


Properties of compact spaces
The one point compactication
Flavors of compactness
Exercises

94
98
101
104
105

Chapter 7. Connectedness and Path-Connectedness


7.1. Connected topological spaces
7.2. Path-connectedness
7.3. Local connectivity
7.4. Exercises

109
109
114
119
121

Chapter 8. Quotient spaces


8.1. Quotient spaces: Denitions, properties and rst examples
8.2. Surfaces as quotient spaces of polygons
8.3. Lens spaces
8.4. Seifert bered spaces
8.5. Real and complex projective spaces
8.6. Cylinders, cones and suspensions
8.7. Exercises

123
123
131
134
136
136
136
136

Chapter 9. Foundational theorems


9.1. The Tietze extension theorem
9.2. Metrization theorems
9.3. The Borsuk-Ulam theorem
9.4. The Tychono compactness theorem
9.5. Exercises

139
139
142
142
142
142

Chapter 10. Spaces of functions

143

CHAPTER 1

Continuity and convergence in Euclidean spaces

his chapter is motivational in nature and provides impetus for the denitions
of a topological space X and a continuous function f : X Y between
two topological spaces X and Y , given in Chapters 2 and 3 respectively.
These denitions are straightforward but upon rst encounter seem ad hoc
and arbitrary. To convey a sense of where they are coming from, in Section 1.1 we
rst examine the familiar setting of continuity of functions f : Rn Rm on Euclidean
spaces in terms of its customary denition in the , language from analysis. In Section
1.2 we introduce two families of special subsets of Rn , the open and closed subsets. The
discussion of this chapter culminates in Section 1.3 where we recast continuity in terms
of open and closed subsets of Rn . The main results of this chapter are Theorems 1.3.1
and 1.3.2. The remaining Section 1.4 scrutinizes some properties of open and closed
subsets of Rn .
1.1. Continuity in Euclidean space
Let R denote the set of real numbers. We dene intervals in R as any of the subsets
of R of the form
[a, b] = {x R | a x b},
[a, b = {x R | a x < b},
a, b] = {x R | a < x b},
a, b = {x R | a < x < b}.
The rst of these are referred to as segments or closed intervals, the last in turn are
called open intervals.
The n-dimensional Euclidean space Rn is dened as the set of ordered n-tuples of
real numbers:
Rn = {(x1 , ..., xn ) | xi R },
where, as usual, the symbol R denotes the set of real numbers. We will adopt the
convention of denoting the n-tuple (x1 , ..., xn ) simply by x, the n-tuple (y1 , ..., yn ) by
y, etc. Two elements x = (x1 , ..., xn ) and y = (y1 , ..., yn ) from Rn can be added to each
other and each can be multiplied by a real number R in the familiar ways:
x + y = (x1 + y1 , ..., xn + yn )

and

x = ( x1 , ..., xn ),

endowing Rn with the structure of a real n-dimensional vector space.


5

1. CONTINUITY AND CONVERGENCE IN EUCLIDEAN SPACES

Definition 1.1.1. For a real number p 1 and for p = , we dene the functions
dp : Rn Rn [0, as

1
(|x1 y1 |p + ... + |xn yn |p ) p
; if p =
(1.1)
dp (x, y) =

max {|x1 y1 |, ..., |xn yn |}


; if p = .
When p = 2 we call the resulting function d2 the Euclidean distance function. We will
refer to the functions dp as metrics on Rn , something we will justify in Exercise 2.5.6.
For example
d1 (x, y) = |x1 y1 |+ +|xn yn |

and

3
4

d 3 (x, y) = |x1 y1 | + + |xn yn |


4

3
4

4
3

while d2 takes on the familiar form


d2 (x, y) =

(x1 y1 )2 + + (xn yn )2 .

Definition 1.1.2. Let r > 0 be a real number and let x Rn be a point in


Euclidean space. We dene the open (Euclidean) ball Bx (r) of radius r and with center
x as the subset of Rn given by
Bx (r) = {y Rn | d2 (x, y) < r}.
Note that this denition of Bx (r) used the Euclidean distance function d2 . We
could equally well have used any of the distance functions dp (including the case of
p
p = ) in this denition. Thus, well let Bx (r) denote the open ball with center x and
radius r but taken with respect to the metric dp :
p
Bx (r) = {y Rn | dp (x, y) < r}.
p
Figure 1 illustrates the shapes of B(0,0) (1) R2 for the choices of p = 1, 2, .

1
B(0,0) (1)

2
B(0,0) (1) = B(0,0) (1)

B(0,0) (1)

p
Figure 1. Examples of the open balls B(0,0) (1) R2 for p = 1, 2, .
The dotted lines indicate that the boundaries of these shapes are not
part of the balls.

1.1. CONTINUITY IN EUCLIDEAN SPACE

Definition 1.1.3. A function f : Rn Rm is continuous at the point x Rn if


for every > 0 there exists a > 0 such that d2 (x, y) < implies d2 (f (x), f (y)) < .
Said dierently, f is continuous at x Rn if for every > 0 there exists a > 0 such
that
f (Bx ()) Bf (x) ().
The function f is said to be continuous if it is continuous at every point x Rn . See
Figure 2 for an illustration.

Bf (x) ()
f (x)

Bx ()

Rn

Rm

Figure 2. The function f is continuous at x Rn if for every > 0


there is a > 0 such that the ball of radius centered at x (shaded disk
on the left) maps into the ball of radius > 0 with center f (x) (shaded
disk on the right). The amoeba-like shape inside of Bf (x) () indicates
the image of Bx () under f .
Continuity of functions is closely tied to convergence of sequences, a connection we
briey review next. Recall that a sequence x in Rn is merely a function x : N R.
We denoted x(k) by xk and we write {xk } or simply {xk }k to denote the entire
k=1
sequence (rather than writing x, the name of the function, so as to not confuse it with
our notation x for points in Rn ). To indicate that the sequence {xk }k belongs to Rn ,
we write {xk }k Rn .
Definition 1.1.4. Let {xk }k be a sequence in Rn .
(a) We shall say that {xk }k converges to x Rn if for every > 0 there exists a
positive integer k0 such that for all k k0 the relation xk Bx () holds. In
this case we write limk xk = x or simply lim xk = x.
(b) The sequence {xk }k is called a Cauchy sequence if for every > 0 there is an
integer k0 such that whenever k, m k0 then d2 (xk , xm ) < .

1. CONTINUITY AND CONVERGENCE IN EUCLIDEAN SPACES

The following two theorems are standard results discussed in any introductory
analysis course, their proofs are omitted.
Theorem 1.1.5. Let f : Rn Rm be a function and x Rn be any point. Then f
is continuous at x if and only if for every sequence {xk }k Rn that converges to x in
Rn , the sequence {yk }k with yk = f (xk ), converges to y = f (x) in Rm . Said dierently,
f is continuous at x if and only if it commutes with the limit symbol lim when applied
to any sequence {xk }k with limit x:
f

lim xk = lim f (xk ).

Theorem 1.1.6. A sequence {xk }k R is convergent if and only if it is Cauchy.


1.2. Open and closed subsets of Euclidean space
Given a set X and a subset A X, we dene the complement of A in X as the set
X A = {x X | x A}.
/
Definition 1.2.1. A subset A Rn is called closed if for every convergent sequence
{ai }i A the limit limi ai also lies in A. A subset B Rn is called open if Rn B
is closed.
Example 1.2.2. When n = 1, all closed intervals [a, b] are closed sets while open
intervals a, b are open sets in R. The sets R and are both closed and open while
the intervals [a, b are neither closed nor open. Any nite subset of R is closed.
Example 1.2.3. When n = 2 examples of closed sets are the closed rectangle
[a, b] [c, d], the closed circles {(x, y) R2 | x2 + y 2 r}, the upper half-plane {(x, y)
R2 | y 0}, the graph f = {(x, f (x)) | x R} of a continous function f : R R. The
main examples of open sets are the open balls Bx (r).
The following proposition gives a characterization of open sets independent of the
denition of closed sets.
Proposition 1.2.4. A subset U Rn is open if and only if for every x U there
exists a > 0 such that Bx () U .
Proof. = Let U be an open subset of Rn and let x U be an arbitrary
element. Suppose there were no > 0 for which the inclusion Bx () U were true.
Then, for any positive integer i, the ball Bx ( 1 ) would have to intersect the complement
i
V = Rn U of U . Pick an arbitrary element bi V Bx ( 1 ), one for each i N, thus
i
creating a sequence {bi }i V . The property d2 (x, bi ) < 1 implies that bi converges to
i
x. However, since U is open, V must be closed and so x = lim bi must lie in V . But
clearly x V , creating a contradiction. Therefore, an with the property Bx () U
/
must exist.
= To prove that U is open we need to prove that V = Rn U is closed. Supposed
that this fails. Then there must exists a convergent sequence {bi }i V whose limit

1.3. CONTINUITY AND CONVERGENCE IN TERMS OF OPEN AND CLOSED SETS

x = lim bi lies in U . Pick an > 0 so that Bx () U . Since none of the bi lie in U , we


see that d2 (bi , x) . This however is impossible for a convergent sequence, creating
a contradiction. We are then forced to conclude that V is a closed subset of Rn and
hence that U is open.
Example 1.2.5. The open balls Bx (r) are indeed open set. This follows easily
from the preceding proposition. For if y Bx (r), let = min{d2 (x, y), r d2 (x, y)}. If
y = x then is positive and By () Bx (r). If y = x we may take = r and the same
conclusion holds.
1.3. Continuity and convergence in terms of open and closed sets
The purpose of this section is to prove that the continuity of a function f : Rn Rm
either at a point or globally, can be expressed entirely in terms of open or closed sets.
Recall that the preimage f 1 (V ) of a function f : X Y between two sets X and
Y , and of a subset V Y , is dened as
f 1 (V ) = {x X | f (x) V }.
In particular, f 1 (V ) is a subset of X. We remark that the notation f 1 used above
does not suggest that f has an inverse function. For example, if f : R R is the
function f (x) = x2 , then f 1 ([0, 1]) = [1, 1].
Theorem 1.3.1. Let f : Rn Rm be a function.
(a) The function f is continous at x Rn if and only if for every open set V Rm
containing f (x), there exists an open set U Rn containing x with the property
that f (U ) V .
(b) The function f is continuous if and only if for every open set V Rm the
preimage U = f 1 (V ) of V under f , is an open subset of Rn .
Proof. (a) = Assume that f is continuous at x Rn and let V Rm be
any open set containing f (x). Since V is open, by Proposition 1.2.4, there exists an
> 0 such that Bf (x) () is contained in V . But continuity of f at x then implies the
existence of a > 0 such that f (Bx ()) Bf (x) (). Since Bx () is an open set (see
example 1.2.5) and since Bf (x) () is contained in V , by taking U = Bx (), we see that
f (U ) V , as claimed. See Figure 3 for an illustration.
= Assume that f has the property that for every open subset V Rm that
contains f (x) there is an open subset U Rn containing x such that f (U ) V .
Wed like to show that then f is continuous at x. Pick an arbitrary > 0 and let
V = Bf (x) (). Note that V is an open set containing f (x) and so we are guaranteed
the existence of an open set U containing x such that f (U ) Bf (x) (). But since U is
open, there must exist a > 0 such that Bx () U (by Proposition 1.2.4). Thus we
found a such that f (Bx ()) Bf (x) () and so f is continuous at x, see Figure 4.
(b) = Suppose f is continuous and let V Rm be any open set. Wed like
to show that U = f 1 (V ) is also an open set. Let x U be any point. Since f is

10

1. CONTINUITY AND CONVERGENCE IN EUCLIDEAN SPACES

V
Bf (x) ()

Bx ()
x
Rn

Rm

Figure 3. This picture illustrates the portion (a) = of the proof of


Theorem 1.3.1. V is an arbitrary open set containing f (x) and > 0
is chosen so that Bf (x) () (shaded disk on the right) is contained in V .
By continuity of f at x, there is a > 0 so that Bx () (shaded disks
on the left) maps into B(f (x) () under f . The image of Bx () under f is
indicated by the smaller amoeba-like shape inside of Bf (x) ().
continuous at x, part (a) of the theorem shows that there exists an open set Ux with
f (Ux ) V . In particular, Ux U . But since Ux is open and x Ux , there must exist
a > 0 so that Bx () Ux and so Bx () U . According to proposition 1.2.4, this
shows that U is open.
= Suppose that f has the property that f 1 (V ) is an open set whenever V is an
open set. Wed then like to show that f must be continuous. By denition, this means
that we must show that f is continuous at each point x Rn . Using part (a) of the
theorem, demonstrating the continuity of f at x is equivalent to showing that for each
open V Rm containing f (x), there exists an open set U Rn containing x and such
that f (U ) V . But our working assumption allows us to simply take U = f 1 (V ),
thus completing the proof of the theorem.
The importance of the preceding theorem is that it provides a denition of continuity that only relies on the notion of open sets. This observation will serve as the basis
for the denition of a topological space (in Chapter 2) and the generalization of continuity to such settings. The next theorem testies that one can also dene continuity
in terms of closed sets only.
Theorem 1.3.2. A function f : Rn Rm is continuous if and only if f 1 (B) is a
closed subset of Rn whenever B is a closed subset of Rm .

1.3. CONTINUITY AND CONVERGENCE IN TERMS OF OPEN AND CLOSED SETS

11

f
V = Bf (x) ()

U
Bx ()
x

Rn

Rm

Figure 4. This picture illustrates the portion (a) = of the proof of


Theorem 1.3.1. Given any > 0 we set V = Bf (x) () rendering it an open
set containing f (x). By assumption, there is an open set U (amoeba-like
set on the left) containing x and such that f (U ) V . But since U is
open, there is a > 0 so that Bx () (shaded disk on the left) is contained
in U . The images of U and Bx () are indicated inside of V .
Proof. = Suppose that f is continuous and let V Rm be a closed set. Wed
like to show that A = f 1 (B) is a closed subset of Rn . To see this, note that
f 1 (Rm B) = Rn f 1 (B) = Rn A.
Since Rm B is open and since according to Theorem 1.3.1 f 1 (Rm B) then also
must be open, the above equality of sets shows that Rn A is open. By denition, this
means that A is closed.
= Suppose that f 1 (B) is closed whenever B is closed. Let V Rm be any
open set, let U = f 1 (V ) and set B = Rm V (note that B is closed). Then, on one
hand, A = f 1 (B) is closed while on the other,
f 1 (B) = f 1 (Rm V ) = Rn f 1 (V ) = Rn U.
Thus U must be open and so, according to Theorem 1.3.1, f must be continuous.
We next turn to sequences where, as with continuity, convergence can be expressed
solely in terms of open sets.
Theorem 1.3.3. A sequence {xk }k Rn converges to x Rn if and only if for
every open set U Rn containing x, there exists a natural number k0 such that for all
k k0 we obtain xk U .

12

1. CONTINUITY AND CONVERGENCE IN EUCLIDEAN SPACES

Proof. = Suppose that lim xk = x and that U is an open set containing x.


Then there exists an > 0 such that Bx () U . Find a natural number k0 such that
for all k k0 the points xk lie in Bx (). Clearly then xk U also for all k k0 .
= Suppose now that xk is a sequence in Rn with the property that for every
open set U containing x, there is a natural number k0 such that k k0 implies that
xk U . Given any > 0, we need to demonstrate that there is a k0 so that k k0
implies that xk Bx (). Of course, choosing U = Bx () nishes the proof.
1.4. Some properties of open and closed sets in Euclidean space
In the proof of Theorem 1.4.1 below, we shall use DeMorgans laws (1.2) below.
To state them, let X be any set and let Ui Rn be a family of subsets of X with i
running through some indexing set I. Then the following, easy to prove equalities of
sets hold:
X (iI Ui ) = iI (X Ui )
(1.2)

X (iI Ui ) = iI (X Ui )

Said dierently, these relations imply that


The complement of the intersection is the union of the complements.
The complement of the union is the intersection of the complements.
With these preliminaries in place, we are ready to turn to the main result of this
section.
Theorem 1.4.1. The following are properties of open subsets of Euclidean space:
(a) The sets Rn and are both open and closed sets.
(b) The union of any number of open sets is an open set.
(c) The intersection of any nite number of open sets is an open set.
Proof. (a) Follows directly from the denitions of open and closed subsets of Rn .
(b) Let I be an arbitrary indexing set and for each i I, let Ui Rn be an open
set. Let U = iI Ui . We need to show that U is an open set. Let x U . Then x Ui
for some i I. By Proposition 1.2.4, there must exist a > 0 such that Bx () Ui .
But then Bx () U since Ui U . This shows that for every x U there is a > 0
such that Bx () U . According to Proposition 1.2.4 this means that U is an open set.
(c) Let U1 , ..., Um be a nite family of open sets and let V = m Uj . To see that
j=1
V is open, pick an arbitrary x V . Then x Ui for every j {1, 2, ..., m} and so there
exist numbers j > 0 with the property that Bx (j ) Uj . Let = min{1 , ..., m } and
note that > 0. Clearly then Bx () Uj for every j {1, ..., m} so that Bx () V .
This shows that V is an open set.
Corollary 1.4.2. The intersection of any number of closed subsets of Rn is a
closed set. The union of any nite number of closed subsets of Rn is a closed set.

1.5. EXERCISES

13

Proof. The corollary is a direct consequence of Theorem 1.4.1 and DeMorgans


laws (1.2). Namely, given a family Vi Rn of closed sets, indexed by an indexing set
I, set Ui = Rn Vi . Clearly each Ui is then open and so by part (b) of Theorem 1.4.1,
so is U = iI Ui . But then V = Rn U is closed, whereas by DeMorgans laws, V
equals
V = Rn U = Rn (iI Ui ) = iI (Rn Ui ) = iI (Rn (Rn Vi )) = iI Vi .
The case of nite unions of closed sets follows similarly and is left as an easy exercise.
There are many examples of innite families of open sets whose intersection is not
open, and innite families of closed sets whose union is not closed.
Example 1.4.3. For each i N let Ui = 1 , 1 . Then each set Ui is an open
i i
subset of R, but the intersection Ui = {0} is not open.
i=1
Example 1.4.4. For i N let Vi = 1 , 3
i
their union Vi = 0, 3 is not closed.
i=1

1
i

. Each Vi is a closed subset of R but

1.5. Exercises
1.5.1. Determine if the following subsets of R are open, closed or neither:
(a) [0, 1 1, 2].
(b) Q.
(c) nZ [2n, 2n + 1].
(d) The Cantor set C. (The Cantor set C is obtained from [0, 1] by dividing it into 3
segments of equal length, and removing the middle open interval 1 , 2 . In the
3 3
2
remaining union of two segments [0, 1 ] 3 , 1], each is divided into 3 segments of
3
equal length, and from each the middle open interval is again removed so as to
1
obtain[0, 1 ] [ 2 , 3 ] [ 2 , 7 ] [ 7 , 1]. The Cantor set is obtained by continuing this
9
9
3 9
9
process ad innitum. Thus at each stage, one divides each remaining segment
into 3 segments of equal length, and removes the open middle interval. See
Figure 5 for the rst several iteration of this construction.)
C1 = [0, 1]
1
C2 = [0, 3 ] [ 2 , 1]
3
1
2
8
C3 = [0, 9 ] [ 2 , 1 ] [ 3 , 7 ] [ 9 , 1]
9 3
9
1
2
C4 = [0, 27 ] [ 27 , 1 ] [ 8 , 25 ] [ 26 , 1]
9
9 27
27
1
2 1
C5 = [0, 81 ] [ 81 , 27 ] [ 26 , 79 ] [ 80 , 1]
27 81
81

Figure 5. The rst ve iterations Cn in the construction of the Cantor set.


1.5.2. Proof the claims made in

14

1. CONTINUITY AND CONVERGENCE IN EUCLIDEAN SPACES

(a) Example 1.2.2,


(b) Example 1.2.3.
1.5.3. Let A be a subset of R that is both open and closed. Show that A is either
the empty set or else A = R. Is the same claim true for subsets of Rn ?
1.5.4. For any p [1, {} once can dene the p-open subsets of Rn as the sets
p
U Rn such that for every x U there exists an > 0 with Bx () U . Show that
the 1-open subsets and the -open subsets of Rn are the same as the open subsets of
Rn . More generally, show that the set of p-open subsets is independent of the choice
of p [1, {}.
1.5.5. A point x Rn is said to be an accumulation point of the subset A Rn if
every open set U Rn that contains x, intersects A nontrivially. Show that A Rn
is a closed set if and only if it contains all of its accumulation points.

CHAPTER 2

Topological spaces

ere we formally introduce the notion of a topological space X as a nonempty


set X equipped with a topology T . A topology on X is merely a collection
of subsets of X, subject to axioms motivated by the results of Theorem 1.4.1
for the case of X = Rn . We give many examples of topological spaces in
Section 2.2, examples which we rely on in later chapters to examine various properties of topological spaces. In Section 2.3 we introduce the important notions of the
interior, closure and boundary of a subset of a topological space, and verify some of
their properties. The nal section of this chapter introduces bases and subbases of a
topology, and denes notions such as rst and second countability and separability of a
topological space.
2.1. Denition of a topological space
Definition 2.1.1. A topology T on a set X is a collection of subsets of X subject
to the following three rules, called the Axioms of a topology:
1. The empty set and all of X belong to T .
2. Let I be an arbitrary indexing set and for each i I, pick a set Ui T . Then
we require that iI Ui also belong to T . Said dierently, T must be closed
under taking arbitrary unions.
3. For elements U1 , ..., Un T , the set U1 ... Un must also belong to T . Said
dierently, T must be closed under taking nite intersections.
A topological space is a pair (X, T ) consisting of a set X and a topology T on X. A
subset U of X is called an open set if U T and a subset V X is called closed if
X V is open. A neighborhood of a point x X is any open set U X that contains
x.
This denition is central to the remainder of the book and so, before moving on
to consider examples, we rst pause to elucidate its various aspects. The choice of the
three axioms of a topology should not be too surprising given the results from Chapter
1. Specically, they are modeled on the three properties of open subsets of Euclidean
space proved in Theorem 1.4.1. Keeping in mind that open subsets of Rn were used
to recast the denition of continuity (Theorem 1.3.1), the attentive reader will have
little diculty in guessing what the denition of a continuous function between two
topological spaces should be (for an answer, see Denition 3.1.1).
15

16

2. TOPOLOGICAL SPACES

Given a topological space (X, T ) we shall often simply write X when the topology
T is understood from context, and refer to X as a topological space. On the other
hand, when several topological spaces are involved in a discussion, we may label the
topology T by TX to indicate that it belongs with X. For example, we shall write
(X, TX ) or (Y, TY ) to label topological spaces.
The notion of open and closed subset of a topological space X shall be crucial to
all subsequent chapters. Whether or not a given subset A X is open or not, depends
on the choice of a topology T on X. As we shall see in the examples below, a set X
admits many dierent topologies and a subset A X may be open with respect to
some but not with respect to other topologies. The most common misconception about
open and closed sets among novices, is the notion that they form a dichotomy.
Remark 2.1.2. In a topological space X, the notions of open subsetand closed
subsetdo not form a dichotomy. That is, the failure of a subset A X to be open
does not typically imply that A is closed. Conversely, the failure of A to be closed does
not typically render it open. Typically, a topological space has numerous subsets that
are neither open nor closed, but can also have subsets that are both open and closed.
The empty set and all of X are examples of the latter.
Definition 2.1.3. Given a set X and two topologies T1 and T2 on X, we say that
T1 is ner than T2 or, equivalently, that T2 is coarser than T1 , if T2 T1 . We shall
write T2 T1 to denote this relation between the two topologies. As usual, we shall
write T2 < T1 to indicate that T2 T1 and T2 = T1 .
The relation gives the set of all topologies on X a partial ordering in that
T1 T2 and T2 T1 imply T1 = T2 . Likewise, T1 T2 and T2 T3 imply T1 T3 .
However, given two topologies T1 and T2 on X, neither of T1 T2 or T2 T1 has to
necessarily hold.
Before exploring properties of open and closed subsets of a topological space X,
we turn to examine several examples of topological spaces. We recommend that the
reader not skip this next section, it will be used as a testing ground for many of the
concepts touched upon in later chapters.
2.2. Examples of topological spaces
This section is devoted to exploring some of the many examples of topological
spaces. Example 2.2.4 below shows that every set X can be equipped with a topology,
exhibiting that topological spaces are indeed very common animals in the jungle of
mathematics. Examples 2.2.1 2.2.3 discussing the Euclidean, the subspace and the
metric topology, are particularly relevant as they make frequent appearances throughout the text.
Example 2.2.1. The Euclidean topology Work from Chapter 1 shows that the
Euclidean space Rn becomes a topological space when equipped with the topology TEu ,
henceforth referred to as the Euclidean topology, dened as
TEu = {U Rn | x U r > 0 such that Bx (r) U }.

2.2. EXAMPLES OF TOPOLOGICAL SPACES

17

Recall that Bx (r) denotes the open ball {y Rn | d2 (x, y) < r} from Denition 1.1.1.
With this denition of TEu , the verication of the three axioms of topology is provided
courtesy of Theorem 1.4.1.
Example 2.2.2. The relative or subspace topology. Let (X, T ) be a given
topological space and let A be a subset of X. Then A automatically inherits the
structure of a topological space from X by equipping it with the relative topology or
subspace topology TA dened as:
TA = {U A | U T }
We call the topological space (A, TA ) a subspace of X. Saying that A is a subspace of
Xmeans that we have given the subset A of X the subspace topology.
It is quite straightforward to verify that TA satises the axioms of a topology on A:
1. Since , X T then A = and X A = A belong to TA .
2. Let Ui TA , i I be given and set U = iI Ui . For each Ui there exists a
Vi T such that Ui = Vi A. But then U = V A where V = iI Vi T
showing that U TA .
3. Let U1 , ..., Un TA and nd sets V1 , ..., Vn T such that Ui = Vi A. Then the
set U = U1 ... Un equals V A where V = V1 ... Vn T and is therefore
contained in TA .
The subspace topology gives us immediately a myriad of examples of topological
spaces by applying it to various subsets of (Rn , TEu ) from the previous example. For
instance, each of
the n-sphere
the graph of f : Rn Rm
the 2-dimensional torus

S n = {x Rn+1 | d2 (x, 0) = 1} Rn+1 ,


f = {(x, f (x)) Rn Rm | x Rn } Rn+m ,
T 2 = {x R4 | x2 + x2 = 1 and x2 + x2 = 1} R4 ,
4
3
2
1

becomes a topological space with the relative Euclidean topology. In the denition of
T 2 , the symbol x was used to denote the ordered quadruple (x1 , x2 , x3 , x4 ).
Example 2.2.3. Metric spaces. A metric space is a pair (X, d) consisting of a
non-empty set X and a function d : X X [0, , referred to as the metric on X,
that is subject to the next three axioms of a metric:
1. d(x, y) = 0 if and only if x = y.
2. Symmetry: d(x, y) = d(y, x) for all x, y X.
3. Triangle inequality: d(x, z) d(x, y) + d(y, z) for all x, y, z X.
When the metric d is understood from context, we will call X itself a metric space. In
analogy to the case of the Euclidean metric d2 on Rn , here too we can dene what we
shall again call the open ball with center x U and radius r > 0 as
Bx (r) = {y X | d(x, y) < r}.
Every metric space (X, d) comes equipped with a natural choice of topology Td , called
the metric topology, dened by
Td = {U X | p U r > 0 such that Bp (r) U }.

18

2. TOPOLOGICAL SPACES

The reader will have noticed that this denition agrees with the denition of the Euclidean topology from Example 2.2.1. Indeed, the premier examples of metric spaces
are (Rn , dp ) with p [1, {} (with dp as given in Denition 1.1.1). The verication of the axioms of a metric for dp is deferred to Exercise 2.5.6. The fact that Td is
indeed a topology on X follows exactly as in proof of Theorem 1.4.1 for the Euclidean
space. Indeed, the proof of the latter never uses the fact that it deals with (Rn , d2 )
explicitly.
We shall encounter more about metric spaces in later chapters and so for now
we limit ourselves to only one additional example. Consider the segment [a, b] R
and let X = C 0 ([a, b], R) be the set of continuous functions f : [a, b] R. Dene
d : X X [0, as
b

|f (t) g(t)| dt.

d(f, g) =
a

Then (X, d) is a metric space and thus becomes a topological space. The second
and third axiom of a metric are readily veried (and their verication does not use
continuity of the functions involved):
b

|f (t) g(t)| dt =

d(f, g) =
a

|g(t) f (t)| dt = d(g, f )


a

|f (t) h(t)| dt

d(f, h) =
a
b

|(f (t) g(t)) + (g(t) h(t))| dt

=
a
b

|f (t) g(t)| + |g(t) h(t)| dt


a
b

|f (t) g(t)| dt +

=
a

|g(t) h(t)| dt
a

= d(f, g) + d(g, h)
The demonstration of axiom 1 of a metric is left for Exercise 2.5.1.
Example 2.2.4. The discrete and indiscrete topologies. Every set X always
admits topologies. Namely, every set can be made into a topological space by choosing
either the indiscrete topology Tindis (also referred to as the trivial topology) or the
discrete topology Tdis dened as
Tindis = {, X},
Tdis = {A | A X}.
Thus Tindis only contains the empty set and all of X and is therefore the smallest
possible topology on X (according to the rst axiom of a topology) while Tdis equals the
entire power set of X and is consequently the largest topology on X. In the notation of

2.2. EXAMPLES OF TOPOLOGICAL SPACES

19

Denition 2.1.3, any topology T on X satises the double inequality Tindis T Tdis .
The axioms of a topology are trivially true for Tindis and Tdis .
These two extreme topologies on X do not lead to interesting topological properties.
For example, as we shall see in Chapter 3, every function on (X, Tdis ) is continuous.
Fertile ground for topological exploration lies with those topologies that live between
these two extremes.
Example 2.2.5. Topologies on nite sets. On nite sets, topologies are by
necessity also nite and can be listed by simply listing their elements. For instance, if
X = {1, 2, 3, 4, 5}, then each of the following is a topology on X (Exercise 2.5.2):
(a) T1 = {, X, {1, 2}, {3, 4, 5}}
(b) T2 = {, X, {1}, {2}, {1, 2}}
(c) T3 = {, X, {1, 2, 3}, {2, 3, 4}, {2, 3}, {1, 2, 3, 4}}
(d) T4 = {, X, {1}, {2}, {3}, {1, 2}, {2, 3}, {1, 3}, {1, 2, 3}}
Example 2.2.6. Included point topology. Let X be any non-empty set and let
p X be an arbitrary point. We dene the included point topology on X as
Tp = {U X | U is the empty set, or p U }.
To see that (X, Tp ) is a topological space, we need to verify the axioms of topology for
Tp from Denition 2.1.1.
1. Clearly Tp be denition. Also, X Tp since p X.
2. Let Ui Tp with i I where I is any indexing set and let U = iI Ui . If each
Ui is the empty set then so is U and is therefore contained in Tp . If at least one
set Ui is not empty, then p Ui and thus p U showing again that U Tp . So
in either case U must belong to Tp .
3. Let U1 , ..., Un Tp and let V = n Ui . If even one of U1 , ..., Un is empty then
i=1
V is empty as well and thus a member of Tp . If none of U1 , ..., Un is empty then
they each must contain p and therefore V must contain p as well. So in this
case V is also in Tp .
Example 2.2.7. The excluded point topology. Let X again be any non-empty
set and, as in the previous example, pick an arbitrary point p X. The excluded point
topology T p on X is then dened to be
T p = {U X | U equals X, or p (X U )}.
Lets verify the axioms of a topology:
1. X belongs to T p be denition and belongs to T p since p .
/
2. Let Ui T p , i I and set U = iI Ui . If at least one Ui equals X then U = X
and so U T p . If none of the Ui equals X then no Ui can contain p and so p
cannot be contained in U either. Thus, in this case too, we get U T p .
3. Take U1 , ..., Un T p and let V = U1 ... Un . If all of the set Ui happen to
equal X then so does V and is thus automatically contained in T p . Conversely,
if there is at least one Ui not equal to X then that particular Ui cannot contain

20

2. TOPOLOGICAL SPACES

p and consequently neither can V . Thus, in this case too, V is again an element
of T p .
See Exercise 2.5.3 for a generalization of the included/excluded point topology.
Definition 2.2.8. Let X be any non-empty set. A partition P on X is a collection
of subsets of X such that
1. If A, B P are two distinct elements of P then A B = .
2. The elements of P cover all of X: AP A = X.
Thus a partition P is a way of dividing all of X into mutually disjoint set. The partition
P = {X} consisting of only X shall be referred to as the trivial partition.
Examples of partitions abound. For instance, if we take X = R, then
P1 = {[a, a + 1 | a Z},
P2 = {Q, R Q},
(2.1)

P3 = {x + Z | x , + 1]},

are all examples of partitions.


Example 2.2.9. Partition topology Let X be a non-empty set and let P =
{Ui | i I} be a partition on X. We then dene the partition topology TP as
TP = {jJ Uj | J I}.
Thus elements of TP are obtained by taking unions of set from P. To see that this is
a topology we check the three axioms of a topology.
1. Choosing J = and J = I renders the set jJ Uj equal to the empty set and
all of X respectively.
2. Let Jk , k K be a family of subsets of I giving rise to the sets Vk = jJk Uj
from TP and let V = kK Vk . Rewriting this denition of V we see that
V = jL Uj

with

L = kK Jk I

Thus, of course, V belongs to TP .


3. This case proceeds in complete analogy with the previous point. Let V1 =
jJ1 Uj , ..., Vn = jJn Uj and set W = n Vk . But then
k=1
W = jL Uj

with

L = n Jk I
k=1

we see again that W lies in P.


For instance, let us pick the partion P1 from (2.1) on X = R. Examples of open
sets in the associated partition topology TP1 are intervals of the form [a, b as well as
[a, and , b with a, b Z. However 0, 1 is not an open set in this topology
(verify this!).
Example 2.2.10. Finite complement topology. On a non-empty set X we
dene the nite complement topology Tf c as
Tf c = {U X | U =

or

X U is a nite set }.

2.2. EXAMPLES OF TOPOLOGICAL SPACES

21

If X is itself a nite set then the nite complement topology agrees with the discrete
topology (from Example 2.2.4) but if X is innite, Tf c and Tdis are rather dierent
topologies. We leave the verication of the axioms of topology for Tf c for Exercise
2.5.8.
Example 2.2.11. The countable complement topology. The countable complement topology Tcc on a set X is dene as
Tcc = {U X | U =

or

X U is a countable set }.

Note that if X is itself countable then Tcc agrees with the discrete topology Tdis on X.
However, for example on X = R, the two topologies are dierent. The axioms of a
topology for this example are addressed in Exercise 2.5.8.
Example 2.2.12. The Fort topology. This example is obtained by combining
the excluded point topology with the nite complement topology from above. Assume
that X is an innite set and let p X be an arbitrary point. We then dene the Fort
topology TF ,p as
TF ,p = {U X | Either X U is nite or p U }.
/
We turn to the verication of the axioms of a topology.
1. Since p and since X X is a nite set, we nd that , X TF ,p .
/
2. Let Ui TF ,p , i I be a family of sets in this topology and set U = iI Ui .
To see that U too must belong to TF ,p we must consider two cases separately.
Firstly, suppose that none of the sets Ui contains p. In this case U doesnt
contain p either and so U TF ,p . In the second case, suppose that at least one
Ui contains p (and therefore U must also contain p). This particular Ui then
has to have nite complement X Ui . Since X U X Ui (this follows
for example from DeMorgans laws (1.2)) we see that X U is also nite and
therefore that U TF ,p .
3. Left as an exercise (Exercise 2.5.9).
For instance, choosing X = R and p = 0, the closed sets of the Fort topology TF ,0
are all nite subsets of R and all subsets of R that contain 0.
Example 2.2.13. The order topology In this example we suppose that the set
X is equipped with an ordering , a notion which we briey review before dening
the associated order topology T .
Recall than a relation r on X is simply a subset of X X. It is customary to write
xry if (x, y) r. A typical choice for the name of a relation is a relation symbol,
such as , , , etc. For example, if we called a relation then we would write
x y to indicate that (x, y) belongs to this relation.
A (total) ordering on X is a relation on X subject to the conditions
1. is reexive: x x for any x X.
2. is antisymmetric: If x y and y x then x = y.
3. is transitive: If x y and y z then x z.

22

2. TOPOLOGICAL SPACES

4. satises the trichotomy law: For all x, y X, either x = y or x < y or


y < x.
We shall write x < y to mean that x y but x = y.
Given an ordering on a set X with at least two elements, let us consider the
following special subset of X
Lx = {y X | y < x}

and

Rx = {z X | x < z}.

In terms of these we dene the order topology T associated to the ordering as


T = {U X | U is obtained by taking arbitrary unions of
nite intersections of the sets Lx and Ry with x, y X}
Examples of elements from T are Lx Ry which we shall denote by x, y and refer
to as the open intervals of the order topology. Note that x, y is an empty set unless
there exists an element z X with x < z < y.
Since the set T is closed under nite intersection and arbitrary unions (by denition
of T ), axioms (2) and (3) of a topology (denition 2.1.1) are trivially true. To see
that the empty set and X belong to T we proceed as follows. Given two distinct
elements x, y (recall that we assumed that X has at least two elements), we have either
x < y or y < x (according to trichotomy axiom above). Suppose that y < x, then
x, y = showing that T . On the other hand, for these same x, y X we obtain
X = Lx Ry (Exercise 2.5.11) and so X T .
As an illustration of the order topology, we consider the lexicographic ordering on
Rn . Let denote the standard ordering of the real numbers R and extend it to an
ordering on Rn by the following rule: (x1 , ..., xn ) < (y1 , ..., yn ) if
x1 < y1
x1 = y1 and x2 < y2
x1 = y1 , x2 = y2 and x3 < y3
.
.
.

or,
or,
or,

x1 = y1 , x2 = y2 , ..., xn2 = yn2 and xn1 < yn1


x1 = y1 , x2 = y2 , ..., xn2 = yn2 , xn1 = yn1 and xn < yn .

or,

The thus obtained ordering is called the lexicographic ordering on Rn . When n = 1,


this topology equals the Euclidean topology on R (from Example 2.2.1). However,
when n 2 the resulting order topology on Rn is quite dierent from its Euclidean
counterpart. For example, when n = 2, consider the interval (0, 0), (1, 0) R2 given
by
(0, 0), (1, 0) = {(x, y) R2 | 0 < x < 1} {(0, y) R2 | y > 0} {(1, y) R2 | y < 1}.
Figure 1 shows this interval drawn in the Euclidean plane. Note that the interval
(0, 0), (1, 0) does not belong to the Euclidean topology on R2 .

2.2. EXAMPLES OF TOPOLOGICAL SPACES

23

Figure 1. The shaded region in R2 represents the interval (0, 0), (1, 0)
in the order topology on R2 associated to the lexicographic ordering. The
dotted lines are not part of the region while the full lines are. The region
extends innitely vertically in both directions.
Example 2.2.14. The lower and upper limit topologies on R. The lower-limit
topology Tll on R is dened as
Tll = {U X | U is obtained by taking unions of nite intersections of
sets [a, b with a, b R}.
The denition of Tll shows that it is closed under nite intersections and arbitrary
unions while the empty set and R belong to Tll because, for example, = [0, 1 [3, 5
and R = aZ [a, a + 1 . Thus, Tll is indeed a topology.
The upper limit topology Tul on R is dened analogously by replacing the sets [a, b
in the denition of Tll by the sets a, b].
Notice that the open intervals a, b belong both to Tll and to Tul since, for example,

a, b =

a+
n

1
,b
n

Tll .

The starting value of n in the above union is chosen large enough so that a + 1/n < b.
From this observation it is not hard to show that TEu Tll and TEu Tul . However,
the set [0, 1 belong to Tll but not to TEu showing that TEu = Tll . A similar observation
applies to the upper limit topology.
Example 2.2.15. The topologists sine curve. This example and the two subsequent ones, do not dene new types of topologies, but rather look at subspaces of
(R2 , TEu ) (see Examples 2.2.1 and 2.2.2) with certain special properties that shall be
explored in later chapters.

24

2. TOPOLOGICAL SPACES

We dene the topologists sine curve to be the subspace X R2 given by


X = {(x, sin(1/x)) | x 0, 1]} ({0} [0, 1]).
Thus, the topologists sine curve is the union of the graph of sin(1/x) over 0, 1] union
the closed segment [1, 1] on the y-axis. This space is illustrated in Figure 2a.

1
(a)

...

.
.
.
(b)

(c)

Figure 2. (a) The topologists sine curve. (b) The innite broom. (c)
The Hawaiian earrings.
Example 2.2.16. The innite broom. Let In R2 be the closed straight-line
segment joining the origin (0, 0) to the point (1, 1/n) for n N. The innite broom is
the subspace X of (Rn , TEu ) dened by (see Figure 2b)
X = ( In ) ([0, 1] {0}).
n=1

2.3. PROPERTIES OF OPEN AND CLOSED SETS

25

Example 2.2.17. Hawaiian earrings. For n N let Cn R2 be the circle with


1
center (1/n, 0) and with radius rn = n . Note that all o these circles pass through the
origin (0, 0).The Hawaiian earrings is the subspace X of (R2 , TEu ) given as the union
of these circles (see Figure 2c):
1
1
X = Cn = {(x, y) R2 | (x n )2 + y 2 = ( n )2 }.
n=1
n=1

2.3. Properties of open and closed sets


This section discusses some general properties of open and closed sets of a topological space (X, T ). Recall that a subset U X is called open if U T , a subset A X
is called closed if X A T . While typically a random subset Y X is neither open
nor closed, we shall see presently it can be approximatedby both kinds of sets.

Definition 2.3.1. Let A X be an arbitrary subset of X. Dene the interior A

(or Int(A)) and the closure A (or Cl(A)) of A as

A = union of all open sets contained in A,

A = intersection of all closed sets containing A.



The boundary or frontier A of A is dened as A = A A.
Lemma 2.3.2. Let A be a subset of the topological space X. Then

(a) The interior A of A is an open set and it is the largest open set contained in

A. The equality A = A holds if and only if A is open.

(b) The closure A of A is a closed set and it is the smallest closed set containing

A. The equality A = A holds if and only if A is closed.

(c) A point x X belongs to A if and only if every neighborhood of x intersects A.


Proof. These claims follow readily from the denition of interior and closure of a
set.

(a) Since A is obtained as a union of open sets (those contained in A) it is an


open set. If U is any other open set containing A, then U is one of the sets from the

union which denes A and is thus contained in A, showing that the interior of A is the

largest open set contained in A. If A = A then clearly A must be open since A is open.
Conversely, if A is open then it is clearly the largest open set containing A and thus

by necessity equal to A.

(b) By denition, A is an intersection of closed set and must therefore be closed.


If B is any closed set containing A, then B occurs in the intersection of sets dening

A and hence A B. This shows that A is the smallest closed set containing A. If

A = A then A is closed, since A is. If A is closed then A itself is the smallest closed

set containing A and is thereby equal to A.

(c) Suppose rstly that x A and let U be a neighborhood of x. If we had


U would be a closed set containing A and would properly be
A U = then A

contained in A, an impossibility according to part (b) of the lemma. Thus we must


have A U = .

26

2. TOPOLOGICAL SPACES

Conversely, suppose that x X is a point such that U A = for every neigh


borhood U of x. If we had x A, we could take U = X A which would give an
/

immediate contradiction. Thus we must have x A.


Denition 2.3.1 shows that the inclusions

A A A,

hold for every subset A of X while Lemma 2.3.2 shows that A is always an open set
is always a closed set. It is in this sense that A can be approximated by both
and A
an open set (its interior) and a closed set (its closure). How good an approximation of

A is given by the sets A and A depends heavily on the topology on X.


Example 2.3.3. Consider the set R and its subset A = 0, 1 . Find the interior

A, the closure A and the boundary A of A with respect to the following choices of
topologies on R:
1. The Euclidean topology (Example 2.2.1). With this toplogy A itself is open
and so, by Lemma 2.3.2, the interior of A equals A itself. The closure of A is

a closed set cointaining A. We guess that A = [0, 1]. Indeed, [0, 1] is closed
and contains A and the only smaller subsets cotaining A are [0, 1 , 0, 1] and A
itself, neither of which is closed. Thus

A = A = 0, 1 ,
A = [0, 1],
A = {0, 1},
a result which conrms our Euclidean intuition.
2. The included point topology with p = 0 (Example 2.2.6). Since p A we see
/
= A. On the other hand, the only open set
that A is in fact closed so that A

from Tp that is contained in A is the empty set, thus A = . Therefore,

A = ,

A = A = 0, 1 ,

A = A = 0, 1 .

3. The included point topology with p = 1/2 (Example 2.2.6). Since p A we see

that A = A. However, since p A, the only closed set containing A is all of R

showing that A = R. We arrive at

A = A = 0, 1 ,
A = R,
A = , 0] [1, .
4. The nite complement topology (Example 2.2.10). In this topology, closed set

are nite subsets of R and all of R. This makes is clear that A = R. No subset
= . In summary
of A has nite complement showing that A

A = ,
A = R,
A = R.
The next lemma provides an alternative denition of the boundary A.
Lemma 2.3.4. Let X be a topological space and let A be a subset of X. Then

(a) X A = X A.

(b) Int(X A) = X A.
X A.
(c) A = A

2.3. PROPERTIES OF OPEN AND CLOSED SETS

27

Proof. (a) By denition, the closure of X A is the intersection of all closed sets
containing X A:
X A = BB B

with

B = {B X | B is closed and X A B}.

Let C be the collection of subset of X gotten from B by taking complements of elements


from B:
C = {C X | X C B}.
The key observation now is that C C if and only if C is open (since X C must be
closed) and C is contained in A (since X C contains X A). Therefore, by denition

A = CC C.
An applicatioin of DeMorgans laws 1.2 yields the desired result:

X A = BB B = CC (X C) = X CC C = X A.
(b) This part follows in much the same way as part (a). Namely, let now B be the
family of closed set containing A and C be the family of open sets contained in X A.
As before, there is a bijective correspondence B C by sending B to X B. Using
again DeMorgans law nishes the proof:

Int(X A) = CC C = BB (X B) = X BB B = X A.
(c) This follows easily from part (a) of the present theorem:

A = A A = A (X A) = A X A.

Definition 2.3.5. Let X be a topological space. A subset A of X is called dense

in X if A = X. The space X is called separable if there exists a countable dense subset


A of X.
As many topological spaces X are uncountable as sets, the notion of separability
provides a measure of how big X is as a topological space. A dense subset A X
has the property that it intersects every open set of X (Corollary 2.3.7). Thus, if one
can nd a countable dense subset of X (i.e. if X is separable), one should think of
X as having relatively few open sets and accordingly as being a relatively small
topological space.
Example 2.3.6. The subset A = 0, 1 of R is dense with respect to either the
particular point topology Tp with p = 1/2 or with respect to the nite complement
topology Tf c (Example 2.3.3).
Corollary 2.3.7. A subset A of the topological space (X, T ) is dense if and only
if A U = for all U T other than U = .
Proof. This is a direct consequence of part (c) of Lemma 2.3.2.

28

2. TOPOLOGICAL SPACES

Example 2.3.8. The set Q of rational number is dense in (R, TEu ) since it intersects
every open interval a, b and every open set is a union of open intervals. More generally,
by the same principle, Qn is dense in (Rn , TEu ). Consequently, since Qn is a countable
set for each n 0, (Rn , TEu ) is a separable topological space for all n 0.
Example 2.3.9. Consider the discrete topology Tdis on R (Example 2.2.4). In this

topology every subset A R is open and closed so that A = A for every A R.


Therefore the only dense subset of (R, Tdis ) is R itself. We infer that (R, Tdis ) is not
separable.
2.4. Bases and subbases of a topology
The reader familiar with linear algebra will recall that a basis for a real nite
dimensional vector space V is a set of vectors {e1 , ..., en } V such that every vector
v V can be written uniquely as a linear combination v = 1 e1 + .. + n en with
i R. Thus, while V is typically an innite set, we can capture the totality of
its vectors with the nite set {e1 , ..., en } by relying on the vector space operations of
vector additionand scalar multiplication.
As a topology T on a set X is a collection of subsets of X, it comes equipped with
two operations among its elements, namely those of taking unions and taking intersections. It is thus conceivable, in analogy with the vector space basis, that there are
subsets of T which generateall of T by means of taking unions and/or intersections
of its elements. This is indeed that case and we shall consider both subsets B T
which generateall of T by means of only taking unions of elements from B, and
subsets S T which will generate T by relying on both unions and intersection. The
rst of these cases will lead the notion of a basis for (X, T ), the closest analogy to a
vector space basis. The second will lead to the notion of a subbasis, one that is without
analogue in the world of vector spaces.
Definition 2.4.1. Let (X, T ) be a topological space and let B and S be subsets
of T such that
(a) Every set U T is a union of sets from B.
(b) Every set U T is a union of nite intersections of sets from S.
Then B is called a basis for the topology T while S is called a subbasis for the topology
T . We also say that T is generated by B (by mean of taking unions of elements from
B) or that T is generated by S (by means of taking unions of nite intersections of
elements from S.
If B and S are given by B = {Ui T | i I} and S = {Vj T | j J } with I
and J being two indexing sets, then properties (a) and (b) from Denition 2.4.1 mean
that every set U T can we written as
(a) U = iIU Ui , for some subset IU of I.
(b) U = LU (jJ Vj ), for some family of indices LU and for nite families of
subsets J of J with LU .

2.4. BASES AND SUBBASES OF A TOPOLOGY

29

The indexing set LU from (b) above, is of course allowed to be innite. While we will
typically start out with a topological space (X, T ) and then seek bases and subbases
for T , it is meaningful to ask about going the other way. That is, given a set X and
two collections B, S of subsets of X, under what conditions are B and S a basis and a
subbasis for some topology T . The next two lemmas address this question.
Lemma 2.4.2. A collection B of subsets of a set X is a basis for some topology T
if and only if B has the properties
(a) X = BB B.
(b) For every B1 , B2 B and every point x B1 B2 , there exists an element
B3 B such that x B3 B1 B2 .
If these two conditions are met, the topology T generated by B consists of all possible
unions of elements from B:
T = {U X | U is a union of sets from B}.
Proof. One implication of the lemma is immediate. Namely, if B is a basis for the
topology T then, since X T , it must be that X = BB B. Likewise, if B1 , B2 B,
then B1 B2 is an open set and therefore a union of elements from B.
Lets turn to the other implication. We assume that B is a collection of subsets of
X subject to conditions (a) and (b) from the lemma and let T be the set
T = {U X | U is a union of elements from B}.
It is then clear that X T (by condition (a)) and T (the empty set is the empty
union of any sets). By denition of T , it is obviously closed under unions since unions of
unions are again just unions. To see that T is closed under nite intersection it suces
to show that it is closed under twofold intersections. Thus, let U1 , U2 T , we need to
show that U1 U2 is a union of elements from B. Let x U1 U2 be any point. By
denition of T , there must exist elements B1 , B2 B with x Bi Ui , i = 1, 2. But
then by property (b), there is an element Bx B such that x Bx B1 B2 U1 U2 .
But then clearly
U1 U2 =
Bx ,
xU1 U2

and so U1 U2 T . This shows that T is indeed a topology on X. Moreover, the


denition of T shows that B is a basis for T , as claimed.
Lemma 2.4.3. A collection S of subsets of a set X is a subbasis for a topology T
on X, if and only if X = SS S. In this case, the topology T is obtained as
T = {U X | U is a union of nite intersections of sets from S}.
Proof. If S is a subbasis for a topology on X, then the condition X = SS S is
clearly satised.
Conversely, suppose that X = SS S and let T be dened as in the statement of
the lemma. Then T and X T by the stated condition on S. The fact that T is
closed under nite intersections and arbitrary unions, follows from the denition of T ,

30

2. TOPOLOGICAL SPACES

showing that T is a topology. Likewise, the fact S is a subbasis of T also follows from
the denition.
Lemmas 2.4.2 and 2.4.3 give us ways to dene topologies on a set X by either
picking a basis or a subbasis rst and letting them generate the topology. Here are a
few examples.
Example 2.4.4. The sets
B1 = { a, b | a, b R, a < b},
S1 = { , b , a, | a, b R},
are a basis and subbasis for (R, TEu ) (recall that TEu denotes the Euclidean topology
on R). Similarly, by relying on the density of the rational numbers Q in R, one can
show that
B2 = { a, b | a, b Q, a < b},
S2 = { , b , a, | a, b Q},
are also a basis and a subbasis for (R, TEu ). The key dierence between the two
examples is that the sets B2 and S2 are countable sets while both of B1 and S1 are
uncountable.
Example 2.4.5. Consider the included point topology Tp on R (Example 2.2.6).
The sets {p} and {x, p}, for any x R, are open sets. However, neither {p} nor {x, p}
can be obtained as a union of other nonempty open sets and must therefore be part of
every basis B. Accordingly, every basis B for (R, Tp ) has uncountably many elements.
The smallest possible basis for (R, Tp ) is
B = {{p}, {x, p} | x R {p}}.
Example 2.4.6. Let (X, ) be an ordered set and consider the order topology T
on X (Example 2.2.13). Then a basis and a subbasis for (X, T ) are given by
B = { a, b | a, b X, a < b}

and

S = {La , Rb | a, b X}.

In the case of X = R and with being the usual ordering of real numbers, these two
sets agree with B1 and S1 from Example 2.4.4.
Example 2.4.7. Consider the set R and let S be the collection of subsets of R
given by
S = {x + Q+ , y + Q | x, y R},
where Q+ and Q are the sets of the positive and of the negative rational numbers
respectively. According to Lemma 2.4.3, S is a subbasis for a topology TS on R. An
example of an open set in this topology is 1, 1 Q.
Just as vector spaces are divided into nite dimensional and innite dimensional
examples according to whether or not they possess a nite basis or not, so too topological spaces can be group into two distinct categories. The rst attempt to dene the

2.4. BASES AND SUBBASES OF A TOPOLOGY

31

analogue of a nite dimensional vector space for topological spaces, might be to demand the existence of nite basis for the topology. However, since a topology generated
by a nite basis is by necessity nite, this denition would exclude most interesting
examples (for instance, no Euclidean space (Rn , TEu ) with n 1 admits a nite basis).
Instead, we will divide topological spaces into two categories according to whether or
not they possess a countable basis or not.
Definition 2.4.8. A topological space (X, T ) is called second countable if it possesses a countable basis B.
We should think of a second countable topological space as being small and of
one that isnt second countable, as being large. Another measure of size for a
topological space we encountered previously was that of being separable (Denition
2.3.5). The next lemma explicates the relation between the two.
Lemma 2.4.9. A second countable topological space is separable. The converse is
false in general (Example 2.4.10).
Proof. Let B = {Ui | i N} be a countable basis for the second countable topological space (X, T ). Without loss of generality we can assume that Ui = for any
i N. Let ai Ui be any element and let A = {ai | i N} X. Then A is a countable

set and we claim that A = X. For if U X is any nonempty open set then there exists
some j N with Uj U . But then A U is nonempty (as it contains aj ) showing that
A is dense according to Corollary 2.3.7.
Example 2.4.4 shows that the Euclidean space (R, TEu ) is second countable while
Example 2.4.5 shows that (R, Tp ) is not second countable.
Example 2.4.10. We just saw that the set of real numbers R with the included
point topology Tp isnt a second countable space. On the other hand, let A = {p} R

and notice that A = R (since closed sets in this topology are sets either not containing
p or else all of R). Thus (R, Tp ) is separable.
We nish this section by considering a local version of the notion of second countability.
Definition 2.4.11. Let (X, T ) be a topological space and let x X be a point in
X. A neighborhood basis around x is a collection Bx of neighborhoods of x such that
for every neighborhood U of x there is an element V Bx with x V U . We say
that (X, T ) is rst countable if every point x X possesses a countable neighborhood
basis.
It should be clear that a second countable space is automatically rst countable.
The converse is false as the next example demonstrates.

32

2. TOPOLOGICAL SPACES

Example 2.4.12. Consider the space (R, Tp ) where Tp is the included point topology
(example 2.2.6) and let x R be any point. A neighborhood basis Bx for x is given by

; if x = p
{{p}}
Bx =
{{x, p}}
; if x = p
Thus (R, Tp ) is rst countable but it isnt second countable according to Example 2.4.5.
Example 2.4.13. The real number line R with the nite complement topology
Tf c (Example 2.2.10) is not rst countable. For suppose that Bp = {Ui | i N} were a
countable neighborhood basis at some point p R, with Ui = R{ai , . . . , ai i } for some
1
n
ai , . . . , ai i R {p} and some ni N. Additionally, let A = iN {ai , . . . , ai i } and
n
n
1
1
note that A is a countable set, and hence that R A is innite. For any neighborhood
V = R {b1 , ..., bn } of p there would have to be some i N with Ui V , that is
with {b1 , . . . , bn } {ai , . . . , ai i } A. A contradiction is obtained by simply choosing
1
n
elements bj from R (A {p}), showing that p has no countable neighborhood basis.
Compare to Exercise 2.5.22.
Example 2.4.14. A metric space (X, d) equipped with the metric topology Td
(Example 2.2.3) is always rst countable. Namely, given a point x X we can dene
Bx as
Bx = {Bx (r) | r Q+ }.
where, as before, Q+ is the set of positive rational number. Clearly Bx is a countable
set and if U is any neighborhood of x, then there must exist some real number r > 0
such that Bx (r) U . Taking any r 0, r Q gives an element Bx (r ) Bx with
x Bx (r ) Bx (r).
We saw that second countability and separability are generally two distinct measures for the size of a topology (Lemma 2.4.9). However, for metric spaces, these two
notions agree. The reason for this lies in the rst countability.
Proposition 2.4.15. A metric space (X, d) equipped with the metric topology Td
is separable precisely when it is second countable.
Proof. Let A = {ai | i N} be a countable dense subset of X and for each ai A
let Bi = {Bai (r) | r Q+ } be a countable neighborhood basis for ai . Let B = Bi .
i=1
Since B is a countable union of countable sets, it is itself a countable set. To see that B
is a basis for (X, T ), let U X be an open set and let x U be any point. There has to
exist a rational number r > 0 such that Bx (r) U (be denition of Td , Example 2.2.3).
Consider the open set Bx (r/2). By Corollary 2.3.7 the intersection A Bx (r/2) has to
be nonempty and so without loss of generality we can suppose that a1 A Bx (r/2).
But then x Ba1 (r/2) since a1 Bx (r/2), and additionally Ba1 (r/2) Bx (r) for if
y Ba1 (r/2) then d(y, x) d(y, a1 ) + d(a1 , x) < r/2 + r/2 = r. Since Ba1 (r/2) B,
we have shown that for every point x U there is a set Ux B with x Ux U .
Therefore U = xU Ux showing that B is a basis.

2.5. EXERCISES

33

Remark 2.4.16. It is not true in general that a separable and rst countable space
is second countable (Exercise 2.5.24).
Definition 2.4.17. We say that a topological space (X, T ) is metrizable if there is
a metric on X whose associated metric topology agrees with T .
Proposition 2.4.15 and the observation from Example 2.4.14 can be used to nd
rst examples of non-metrizable topologies (see Section 9.2 for more on metrizability).
For instance, the space (R, Tp ) is not metrizable as according to Example 2.4.10 it is
separable but not second countable (note however that it is a rst countable space
according to Example 2.4.12). Similarly, (R, Tf c ) is not metrizable as it is not rst
countable according to Example 2.4.13.
Theorem 2.4.18. Let (X, TX ) be a topological space and let Y X be a subspace
of X. If X is either second countable or rst countable, then so is Y . Separability of
X does not in general imply separability of Y .
Proof. Suppose that X is second countable and let B = {Ui | i N} be a basis
for the topology on X. Then BY = {Ui Y | i N} is a countable basis for the relative
topology on Y . To see this, let V Y be an open set in Y and let U be an open set in
X such that V = U Y . Since B is a basis for the topology on X, then U = iM Ui
for some subset M N. Thus V = iM (Ui Y ) proving the claim.
The case of rst countability follows analogously. For a topological space showing
that separability is not necessarily inherited by a subspace, see Example 2.4.19.
Example 2.4.19. Consider the included point topology Tp on X = R. This is a
separable space with a dense subset given by {p}. Let Y = R {p} be given the
subspace topology. Since the closed sets in X are those not containing p and X itself,

it follows that the closed subsets of Y are all subsets of Y . Accordingly, A = A for
any A Y . Thus the only dense subset of Y is Y itself. Since Y is not countable, it
follows that it is not separable.
2.5. Exercises
2.5.1. Verify axiom 1 for the metric d(f, g) =

b
a

|f (t) g(t)| dt from Example 2.2.3.

2.5.2. Verify that each of T1 , T2 , T3 , T4 from Example 2.2.5 is a topology on X =


{1, 2, 3, 4, 5}.
2.5.3. This exercise generalizes Examples 2.2.6 and 2.2.7. Let X be a topological
space and A X a subset of X.
(a) Dene TA as the collection of subset of X given by
TA = {U X | U is the empty set, or A U }.
Show that TA is a topology on X (called the Included subset topology).

34

2. TOPOLOGICAL SPACES

(b) Dene T A as the collection of subset of X given by


T A = {U X | U equals X, or A (X U )}.
Show that T A is a topology on X (called the Excluded subset topology).
2.5.4. Let X be a set and p1 , p2 X two distinct points. Dene the collection Tpp2
1
of subsets of X as
Tpp2 = {U X | p1 U, or p2 (X U )}.
1
Show that Tpp2 denes a topology on X (called the Included-excluded points topology).
1
2.5.5. On R dene the set T of subsets of R as
T = {U X | U is the empty set, or R U is a nite union of closed intervals}.
Show that T denes a topology on R (called the Compact complement topology).
2.5.6. Show that the functions dp from Denition 1.1.1 are each a metric for any
choice of p [1, {}.
2.5.7. Two metrics d and d on a set X are called equivalent if there exist positive
real numbers c1 , c2 such that
c1 d(x, y) d (x, y) c2 d(x, y),

for all x, y X.

(a) Show that the notion of equivalence between metrics on a set X is an equivalence
relation.
(b) Show that equivalent metrics induce the same metric topology, that is show
that if d and d are equivalent metrics, then Td = Td (Example 2.2.3).
(c) Show that for any pair p, p [1, {}, the two metrics dp and dp (Denition 1.1.1) are equivalent metrics on Rn . Conclude that all of the metrics dp ,
p [1, {} induce the Euclidean topology on Rn (Hint: Prove that for
any p [1, , the double inequality
1
dp (x, y) d (x, y) dp (x, y)
p
n
holds for all x, y Rn , and use parts (a) and (b) of the exercise.).
2.5.8. Show that the nite complement topology Tf c and the countable complement topology Tcc from Examples 2.2.10 and 2.2.11, satisfy the axioms of a topology
(Denition 2.1.1).
2.5.9. Verify axiom 3 of a topology (Denition 2.1.1) for the Fort topology TF ,p
from Example 2.2.12.
2.5.10. Show that X belongs to the order topology T from Example 2.2.13, by
showing that X = Lx Ry for a pair of elements x, y X with y < x.
2.5.11. For the topological space (X, T ), determine the interior, closure and boundary of the given subset A X.

2.5. EXERCISES

35

(a) X = R with the Euclidean topology TEu and A = Q.


(b) X = R with the Fort topology TF ,p (Example 2.2.12) with p = 0 and A = [0, 1].
(c) X = R2 with the lexicographic order topology T (Example 2.2.13) and A being
the unit square A = [0, 1] [0, 1].
(d) X = R with the lower limit topology Tll (Example 2.2.14) and with A = 0, 1 .
2.5.12. Are the irrational numbers dense in R equipped with the
(a) Euclidean topology TEu ?
(b) Countable complement topology Tcc (Example 2.2.11)?
(c) Fort topology TF ,p (Example 2.2.12)? Does the choice of p R matter?
2.5.13. Let X be a set and let T1 , T2 be two topologies on X and assume that
T1 T2 (Denition 2.1.3). For a subset A of X, write Inti (A), Cli (A) and i (A) for
the interior, closure and boundary of A with respect to Ti . Show that
(a) Int1 (A) Int2 (A).
(b) Cl1 (A) Cl2 (A).
2.5.14. For a topological space X and subsets A, B X, show the following relations: (a) Cl(A B) = Cl(A) Cl(B).
(b) Int(A) Int(B) Int(A B).
Cl(A B) Cl(A) Cl(B).
Int(A) Int(B) = Int(A B).
Cl(Cl(A)) = Cl(A).
Int(Int(A)) = Int(A).
Show by example that the inclusions from the second point of (a) and rst point
of (b) may be proper inclusions.
2.5.15. Let X be a topological space and let Z Y X be subsets. Let TY X
and TZX be the relative topologies on Y and Z respectively, induced by the topology
on X. Furthermore, let TZY be the relative topology on Z induced by the topology
TY X on Y . Show that TZY = TZX .
2.5.16. Let X be a topological space and Y X a subspace. Show that a subset
A Y is closed in Y if and only if there exists a closed subset B X of X such that
A=BY.
2.5.17. Let X be a topological space and let A, B X be two subspaces of X with
A B.
(a) Show that if A is open in B and B is open in X then A is also open in X.
(b) Show that if A is closed in B and B is closed in X then A is also closed in X.
(c) Find an example of A, B and X with A open in B but not in X. Similarly,
nd an example with A closed in B but not in X.
2.5.18. Let X be a topological space and Y X a subspace. Write ClX (A) and
ClY (A) for the closure of A in X and Y respectively. Similarly, write IntX (A) and
IntY (A) for the interior of A in X and Y respectively. Show that for a subset A of Y ,
the following inclusions hold:
(a) ClY (A) ClX (A).

36

2. TOPOLOGICAL SPACES

(b) IntX (A) IntY (A).


Show by example that both inclusions may be proper.
2.5.19. Let X be a set and let T1 , T2 be two topologies on X with T1 T2 . Show
that if (X, T1 ) is separable, then so is (X, T2 ). Conclude that (R, Tll ) and (R, Tul )
(Example 2.2.14) are separable.
2.5.20. Show that Q with the discrete topology (Example 2.2.4) is second countable.
Show on the other hand that R with the discrete topology is not second countable.
2.5.21. Let X be a set and P = {Ui X | i I} a partition of X. Show that X
equipped with the partition topology TP (Example 2.2.9) is second countable if and
only if the indexing set I is countable.
2.5.22. Show that (R, Tcc ) from Example 2.2.11 is neither rst nor second countable.
2.5.23. Show that (R, Tf c ) from Example 2.2.10 is separable.
2.5.24. Show that (R, Tll ) from Example 2.2.14, is separable and rst countable but
not second countable.
2.5.25. Show that any open subset U R with respect to the lower limit topology
Tll (Example 2.2.14), is a countable union of intervals [a, b with a, b R, a < b.

CHAPTER 3

Continuous functions and convergent sequences

ontinuous functions and convergent sequences are introduced, following the


intuition from Euclidean spaces introduced in Chapter 1 (see specically Theorems 1.3.1 and 1.3.3). Many examples of continuous functions are presented,
and some of their basic properties are illuminated in Section 3.1. Section 3.2
focuses on convergent sequences in topological spaces, and heeds special attention to
the question of uniqueness of limit. The connection between continuous functions
and convergent sequences in general topological spaces holds in weakened form, and
Theorem 3.2.11 presents the generalization of Theorem 1.1.5 from the Euclidean case.
Section 3.3 presents only one result - the Uniform convergence theorem(Theorem
3.3.2) - which speaks to when a sequence of continues functions that converges point
wise, denes a continuous limit function. This theorem is used several times in the
remainder of the text, the rst instance of which occurs in Section 3.4. The latter
discusses space lling curves, that is continuous surjections from the closed interval
[0, 1] to the closed cube [0, 1]n for any n N.
3.1. Continuous functions
Definition 3.1.1. Let (X, TX ) and (Y, TY ) be topological spaces and let f : X Y
be a function.
(a) We say that f is continuous at x X if for every neighborhood V of f (x) there
exists a neighborhood U of x such that f (U ) V .
(b) We say that f is continuous if f 1 (V ) TX for every V TY .
We shall use the term map as synonymous with continuous function.
This denition of both local continuity (part (a)) and global continuity (part (b))
are directly motivated by Theorem 1.3.1. In particular, we see that with the choices
of X = Rn and Y = Rm and with both TX and TY being the Euclidean topologies,
Denition 3.1.1 agrees with the usual denition of continuity on Euclidean spaces
familiar from analysis.
Global continuity of a function between Euclidean spaces (rst encountered in Definition 1.1.3) is dened to simply mean local continuity at each point of the domain. In
contrast, Denition 3.1.1 gives separate meaning to both local and global continuity.
Nevertheless, the relation between the two remains the same.
Theorem 3.1.2. A function f : X Y between two topological spaces is continuous if and only if it is continuous at every point x X.
37

38

3. CONTINUOUS FUNCTIONS AND CONVERGENT SEQUENCES

Proof. = Suppose that f : X Y is continuous globally and let x be a


point in X. Pick an arbitrary neighborhood V of f (x) in Y , then U = f 1 (V ) is a
neighborhood of x in X with f (U ) V .
= Suppose that f : X Y is continuous at every point x X, let V be an open
subset of Y and set U = f 1 (V ). Wed like to show that U is open. For that purpose,
let x U be an arbitrary point (if U is the empty set then it is automatically open)
and note that V is a neighborhood of f (x). By continuity of f at x, there must exist
a neighborhood Ux of x with f (Ux ) V . This latter relation shows that Ux U and
consequently we obtain U = xU Ux . Being a union of open sets, U is itself open.
Example 3.1.3. Let (X, TX ) = (R, Tf c ) and let f : X X be the function
f (x) = x2 . Let U Tf c be any nonempty open set and assume that U = R
{x1 , . . . , xk , y1 , . . . y } with x1 , . . . , xk < 0 and y1 , . . . , y 0. Then f 1 (U ) = R
{y1 , y1 , ..., y , y } which has nite complement in R. Since in addition f 1 () = ,
we see that f is continuous.
Example 3.1.4. Let (X, TX ) = (R, Tf c ), let (Y, TY ) = (R, Tp ) with p = 0 and let
f : X Y be again the function f (x) = x2 . Then {0, 1} Tp but f 1 ({0, 1}) =
{1, 0, 1} Tf c showing that f is not continuous.
/
Example 3.1.5. Let X be a non-empty set and let P1 and P2 be two partitions on
X and let T1 and T2 be the two associated partition topologies on X (Example 2.2.9).
Let f : X X be the identity function f (x) = x whose domain is equipped with T1
and codomain with T2 . Then f is continuous if and only if every element in P2 is a
union of elements from P1 .
Example 3.1.6. The constant function f : X Y , given by f (x) = p Y for all
x X, is always continuous since f 1 (U ) is either the empty set if p U or all of X
/
if p U .
Example 3.1.7. Let X be equipped with the discrete topology, then any function
f : X Y is continuous. Conversely, if X is given the indiscrete topology, then a
function f : X Y to a Hausdor space Y is continuous if and only if it is constant.
For if f were not constant then we could nd two points a, b X with f (a) = (b). The
Hausdor property guarantees the existence of two open and disjoint sets U, V Y
with f (a) U and f (b) V . But then f 1 (U ) = (since a f 1 (U )) and f 1 (U ) = X
/
(since b f 1 (U )), showing that f is not continuous.
/
Example 3.1.8. Let (X, dX ) and (Y, dY ) be two metric spaces and let TdX and TdY
be the associated metric topologies. Then a function f : X Y is continuous at x X
if and only if for every > 0 there exists a > 0 such that x X and dX (x, x ) <
implies that dY (f (x ), f (x)) < . Thus continuous functions between metric spaces
satisfy the familiar (, )-rule for continuity from analysis. We leave the verication
of this claim as an exercise (Exercise 3.5.7), it follows along the lines of the proof of
Theorem 1.2.1.

3.1. CONTINUOUS FUNCTIONS

39

The next theorem provides alternative denitions of continuity. Part (a) is a generalization of Theorem 1.3.2 from the Euclidean case to general topological spaces.
Theorem 3.1.9. Let f : (X, TX ) (Y, TY ) be a function between two topological
spaces. Then f is continuous if and only if any of the mutually equivalent conditions
below is met:
(a)
(b)
(c)
(d)
(e)

f 1 (B) is a closed subset of X for any closed subset B of Y .

For all subsets B Y , the inclusion f 1 (B) f 1 (B) holds.

For all subsets A X the inclusion f (A) f (A) holds.


1
For all subsets B Y the inclusion f (Int(B)) Int(f 1 (B)) holds.
Given a basis B = {Vi Y | i I} for TY , f 1 (Vi ) is open for every i I.

Proof. We will show that properties (a) and (e) are each equivalent to f being
continuous. We will then prove the implications (a)=(b)=(c)=(a). Showing that
(d) is equivalent to the continuity of f is left as an exercise (Exercise 3.5.6).
(a) Suppose that f is continuous and let B Y be a closed set. Then f 1 (B) =
X f 1 (Y B) is also closed since Y B is open and continuity of f forces f 1 (Y B)
to be open.
Conversely, suppose that f has property (a) and let V Y be any open set. Then
1
f (V ) = X f 1 (Y V ) is open since Y V and f 1 (Y V ) are both closed, the
latter by property (a).

(a)=(b) Let B be any subset of Y . Since B B, we obtain f 1 (B) f 1 (B).

By property (a) of f , the set f 1 (B) is closed but since the set f 1 (B) is the smallest
1

closed set containing f (B) (see Lemma 2.3.2), the inclusion f 1 (B) f 1 (B) is
immediate.
(b)=(c) Take A X to be any subset of X and apply property (b) of f to
B = f (A) to obtain f 1 (f (A)) f 1 (f (A)). Since A f 1 (f (A)) we also get

A f 1 (f (A)). Applying f to the inclusion A f 1 (f (A)) yields the desired result.

(c)=(a) Let B be a closed subset of Y and set A = f 1 (B). Pick a point x A.

Then, according to property (c) we must have f (x) f (A) = B = B showing that
= A and thus that A is closed.
x A. This implies that A
(e) The necessity of property (e) for a continuous function is obvious. Suppose then
that f possesses property (e) and let V Y be an open set. Let J I be such that
V = jJ Vj . Then f 1 (V ) = f 1 (jJ Vj ) = jJ f 1 (Vj ) showing that f 1 (V ) is a
union of open sets and therefore open.
The reader may have noticed that parts (b), (c) and (d) of Theorem 3.1.9 express
continuity in terms of an image/preimage under f , combined with a choice of taking the
interior/closure of a set, see Table 3.1. Of the four possible combinations resulting in
this manner, one is noticeably absent, namely the one involving taking images under f
combined with taking interiors of set. The next example demonstrates that this fourth
possibility cannot be added to Theorem 3.1.9.

40

3. CONTINUOUS FUNCTIONS AND CONVERGENT SEQUENCES

Preimage Image
Closure
(b)
(c)
Interior
(d)
?
Table 1. Conditions (b), (c) and (d) from Theorem 3.1.9 each involve
a combination of taking an image/preimage under f and taking the closure/interior of a set.
Example 3.1.10. Consider the function f : R2 R given by f (x, y) = x where
each of R2 and R are equipped with the Euclidean topology. Then f is clearly continuous, but, taking A = {(x, 0) R2 | x R}, we nd that
Int(f (A) = Int(R) = R

while

f (Int(A)) = f () = .

Thus the inclusion Int(f (A) f (Int(A)) fails in general for continuous functions.
On the other hand, consider the function g : R R given by g(x) = 0 and assume
that both copies of R come with the Euclidean topology. Pick B = R, then
g(Int(B)) = g(R) = {0}

while

Int(g(B)) = Int({0}) = .

We see that the inclusion g(Int(B)) Int(g(B)) also fails in general for continuous
functions (see however part (a) of Proposition 3.1.22 below).
We next single out some simple functions that are always continuous. The reader
will no doubt recognize familiar properties of continuous functions from the Euclidean
case.
Proposition 3.1.11. Let X, Y , Z be topological spaces and let f : X Y and
g : Y Z be continuous functions. Additionally, let A X be any subspace of X.
(a) The inclusion function : A X is continuous. In particular, the identity
function id: X X is always continuous.
(b) The composition function g f : X Z is continuous.
(c) The restriction function f |A : A Y is continuous.
(d) Let Ui X, i I, be a collection of open subsets of X such that X = iI Ui
and let h : X Y be a function with h|Ui : Ui Y continuous for each i I.
Then h is continuous.
Proof. (a) For an open subset U X, the preimage 1 (U ) equals U A and is
therefore open in A with respect to its relative topology.
(b) Let W be an open subset of Z. Then V = g 1 (W ) is an open subset of Y and
thus U = f 1 (V ) must be open in X. Since U = f 1 (g 1 (W ) = (g f )1 (W ), the
claim follows.
(c) This follows from parts (a) and (b) since f |A = f where : A X is the
inclusion map.
(d) Set hi = h|Ui and let V Y be an open set. Then h1 (V ) = iI h1 (V ) and,
i
since each h1 (V ) must be open in Ui (in its relative topology), there must be open
i

3.1. CONTINUOUS FUNCTIONS

41

subsets Wi X with h1 (V ) = Ui Wi . As both Ui and Wi are open in X then so is


i
Ui Wi showing that h1 (V ) is a union of open sets and therefore open.
The following simple lemma will prove a useful tool in subsequent sections.
Lemma 3.1.12. Let (X, TX ) be a topological space and A, B X two subspaces with
X = A B. Assume that either both A and B are open or that both are closed. Let
(Y, TY ) be another topological space and let f : A Y and g : B Y be continuous
functions which agree on A B, that is assume that f (x) = g(x) for every x A B.
Then the function h : X Y dened by
h(x) =

f (x)
g(x)

;
;

x A,
x B,

is continuous.
Proof. Let V Y be any subset of Y . Then
h1 (V ) = (h1 (V ) A) (h1 (V ) B) = f 1 (V ) g 1 (V ).
If A, B are both open, choose V to be open also. Continuity of f and g shows the sets
f 1 (V ) and g 1 (V ) to be open subsets of A and B respectively. Accordingly they are
also open in X since A and B are open subsets of X (part (a) of Exercise 2.5.17). Thus
h1 (V ) is a union of two open sets and hence open, demonstrating the continuity of h.
If A, B are both closed, pick V Y also closed. The same reasoning as in the
previous case (and by relying on part (b) of Exercise 2.5.17) shows that h1 (V ) is a
closed set. Continuity of h now follows from part (a) of Theorem 3.1.9.
Definition 3.1.13. A function f : X Y between topological spaces is called
a homeomorphism if f is a continuous bijection with a continuous inverse function
f 1 : Y X. We say that two topological spaces X and Y are homeomorphic, and
write X Y , if there exists at least one homeomorphism f : X Y .
=
A function f : X Y is called a local homeomorphism if it is surjective and every
point x X has a neighborhood U such that f |U : U f (U ) is a homeomorphism,
where U and f (U ) come equipped with their relative topologies inherited from X and
Y respectively. Two spaces are called locally homeomorphic if there exists at least one
local homeomorphism between them.
Remark 3.1.14. The relation of being homeomorphic to between topological
spaces is an equivalence relation:
1. Reexivity: X X. A homeomorphism from X to X is given by the identity
=
map (see part (a) of Proposition 3.1.11).
2. Symmetry: X Y implies Y X. If f : X Y is a homeomorphism, then
=
=
f 1 : Y X is also a homeomorphism.
3. Transitivity: X Y and Y Z imply X Z. If f : X Y and g : Y Z
=
=
=
are homeomorphisms, then g f : X Z is also a homeomorphism (part (b)
of Proposition 3.1.11).

42

3. CONTINUOUS FUNCTIONS AND CONVERGENT SEQUENCES

In a similar vein one can also prove that the relation of being locally homeomorphicis
an equivalence relations among topological spaces (Exercise ??).
From an abstract point of view, that is, from a point of view where we disregard
the labels attached to the points of a set, two homeomorphic spaces (X, TX ) and
(Y, TY ) are indistinguishable. For all intents and purposes, two homeomorphic spaces
can be considered as being identical, a practice that we shall adopt from hereon out,
sometimes without specically saying so. To imprint this point further, we turn to a
simple example.
Example 3.1.15. Let the sets X = {1, 2, 3} and Y = {a, b, c} be given the two
topologies
TX = {, X, {1, 2}}
and
TY = {, Y, {b, c}}.
The reader will have no trouble recognizing that the two spaces (X, TX ) and (Y, TY )
are homeomorphic with f : X Y given by f (1) = c, f (2) = b and f (3) = a being
one possible homeomorphism. Thus, if we disregard that the elements in X are labeled
by 1, 2, 3 and those in Y by a, b, c, the spaces (X, TX ) and (Y, TY ) become identical.
We can think of the homeomorphism f : X Y as a relabeling tool, one which
takes an element x from X and gives it a new label f (x). This relabeling has to be
done with care so as to respect the topologies on X and Y , the relabeling of an open
subset from X has to result in an open subset of Y .
Example 3.1.16. The function f : (R, T ) (R, T ) given by f (x) = x3 is a
homeomorphism for every choice of T {TEu , Tp , T p , Tf c , Tcc , TF ,p }, with the choice of
p being either 0 or 1. Proving this amounts to showing the f and its inverse are
continuous function with respect to the listed topologies. For instance, when T = Tp ,
and V R is a nonempty open set, then p V and so p f 1 (V ) since f (p) = p for
p {1, 0, 1}. Thus f 1 (V ) is open and hence f continuous. The verication of the
other claims is left as an exercise (Exercise 3.5.8).
Example 3.1.17. The function f : (R, TEu ) (R, TEu ) given by f (t) = sinh t =
et ) is a homeomorphism with inverse function f 1 (t) = sinh1 t = ln(x +
+ 1).

1 t
(e
2

x2

Example 3.1.18. Let (X, dX ) and (Y, dY ) be two metric spaces. A function f :
X Y is called an isometry if dY (f (a), f (b)) = dX (a, b, ) for any pair of points
a, b X. Considering X and Y equipped with their associated metric topologies TdX
and TdY respectively, any surjective isometry f : X Y becomes an homeomorphism.
To see this we rst verify that any isometry is continuous. Let x X be any
point and let V be any neighborhood of f (x). Then there exists an r > 0 such that
Bf (x) (r) V . But then Bx (r) is a neighborhood of x with f (Bx (r)) V for if
a Bx (r) then dX (x, a) = r = dY (f (x), f (a)) showing that f (x) Bf (x) (r).
It remains to see that f is a bijection and that its inverse function is also continuous.
These are deferred to Exercise 3.5.11.

3.1. CONTINUOUS FUNCTIONS

43

Example 3.1.19 (Stereographic projections). Consider Rn and Rn+1 (with n N)


both equipped with their Euclidean topologies and let S n Rn+1 be the subspace
S n = {(x1 , ...xn+1 ) Rn+1 | x2 + x2 + ... + x2 = 1}.
1
2
n+1
We shall refer to S n as the n-dimensional sphere or n-sphere for short. The goal of
this example is to show that S n {p} is homeomorphic to Rn where p S n is an
arbitrarily chosen point. For convenience, we choose p = N = (0, 0, ..., 0, 1) to be the
north poleof S n .
Let N : (S n {N }) Rn be the function, referred to as stereographic projection
from the north pole N , given by
(x1 , ..., xn )
N (x1 , ..., xn+1 ) =
.
1 xn+1
Note the N is well dened since the only point (x1 , ..., xn+1 ) S n with xn+1 = 1
is the north pole N . It is easy to see that N is a bijection with inverse function
1 : Rn (S n {N }) given by
N
1 (y1 , ..., yn ) =
N

(2y1 , 2y2 , ..., 2yn , |y|2 1)


,
|y|2 + 1

2
2
2
where |y|2 stands for y1 + y2 + ... + yn .
n
Let a = (a1 , ..., an+1 ) S {N } be any point and let > 0 be chosen at will. Let
M = 1 an+1 and dene as any positive real number subject to the inequality

M
M 2

,
2 n(1 + M ) 2

< min

Then N (Ba ()) BN (a) () showing that N is continuous at a and therefore continuous everywhere since a was arbitrary. To verify this inclusion, let a = (a1 , ..., an+1 )
Ba () and note that this implies that |ai ai | < for each i = 1, ..., n + 1 and
1 an+1 > M/2. Thus
n

ai
ai

1 an+1 1 an+1

(d2 (N (a), N (a ))) =


i=1
n

=
i=1

|ai (an+1 an+1 ) + (ai ai )(1 an+1 )|2


|1 an+1 |2 |1 an+1 |2

< 4n 2

(1 + M )2
M4

< 2
as needed. Showing that 1 is continuous follows a similar line of argument and is
N
left as an exercise.
The geometric meaning of the function N , one from which its name derives, is
that N (x) is the intersection point of the line through N and x S n {N } with the
equatorial plane Rn {0} Rn+1 as illustrated in Figure 1.

44

3. CONTINUOUS FUNCTIONS AND CONVERGENT SEQUENCES

N
x

N (x)

Figure 1. The geometric interpretation of the stereographic projection


N : (S n {N }) Rn . Given a point x S n {N }, N (x) is obtained
as the intersection point of the line through N and x, with the equatorial
plane Rn {0} Rn+1 .
Similar stereographic projection can be obtained from any point p S n to Rn
where Rn is characterized as those points in Rn+1 that have dot product zero with p,
i.e. Rn {x Rn+1 | x p = 0}. In Exercise 3.5.9 you will be asked to nd formulas
=
describing the stereographic projection from the south pole of S n .
With these couple of examples in hand, we turn to what is one of the main questions
in topology:
Question 3.1.20. Given two topological spaces (X, TX ) and (Y, TY ), are they homeomorphic?
Given Example 3.1.15 and the discussion preceding it, this question can be interpreted as asking whether or not a pair of topological spaces (X, TX ) and (Y, TY ) are
the same topological space, from an abstract point of view. This is a very dicult
question and at best we can hope to answer it for specic categories of topological
spaces. The chief reason for the diculty is that a given topological space (or better
said, a given homeomorphism class of a topological space) can have many dierent
descriptions, and it is often far from clear when two descriptions yield homeomorphic
spaces. Nevertheless, we shall have more to say about Question 3.1.20 as we progress
in our understudying of topology. The rst partial answers we shall be able to provide,
will come after we develop one of the main tools of topology - topological invariantsin Section 4.3.
Definition 3.1.21. Let f : X Y be a map between topological spaces. We say
that f is open if f (U ) is an open subset of Y whenever U is an open subset of X.
Similarly, we say that f is closed if f (A) is closed in Y for every choice of a closed
subset A X.

3.2. CONVERGENT SEQUENCES

45

Note that every homeomorphism is both an open and closed map. Indeed the
condition for a bijection f to be open is the same as the condition as for f 1 to be
continuous. Part (a) of Theorem 3.1.9 implies that a homeomorphism is also a closed
map. Perhaps more surprisingly, local homeomorphisms are also open maps.
Proposition 3.1.22. Let f : X Y be a map between topological spaces.
(a) f is an open map if and only if f (Int(A)) Int(f (A)).

b) f is a closed map if and only if f (A) f (A) for every subset A X.


(c) If f is a local homeomorphism then f is an open map.
Proof. (a) Suppose rst that f is open and let A be any subset of X. Since
Int(A) A we nd that f (Int(A)) f (A). Since f is open, the set f (Int(A)) is
an open subset of f (A). But Int(f (A)) is the largest open subset of f (A) forcing the
inclusion f (Int(A)) Int(f (A)).
Conversely, suppose that f (Int(A)) Int(f (A)) holds for all A X. Pick an open
subset U X, then f (Int(U )) = f (U ) Int(f (U )). On the other hand, the inclusion
Int(f (U )) f (U ) is trivially true showing that f (U ) = Int(f (U )), thus f (U ) is open.
(b) Assume that f is closed and that A X is any set. Then the inclusion

A A implies the inclusion f (A) f (A). Since f (A) is closed and contains f (A),

the inclusion f (A) f (A) follows since the closure of f (A) is the smallest closed set
containing f (A).

On the other hand, suppose that f (A) f (A) for all subsets A of X, wed like to
show that f is closed. Thus, let B X be any closed set, then

f (B) f (B) = f (B) f (B),


showing that f (B) = f (B). Since f (B) is closed, then so is f (B).
(c) Assume that f is a local homeomorphism. To show that f is an open map, let
U X and set V = f (U ). Wed like to show that V is an open subset of Y . Towards
this goal, pick a point x U . Then there exists a neighborhood Ux of x such that
f |Ux : Ux Vx , with Vx = f (Ux ), is a homeomorphism. Without loss of generality
we can assume that Ux U for if not, we simply replace Ux by Ux U . But then
Vx V and therefore V = xU Vx . Being a union of open sets, V is forced to be open
itself.
We will encounter open and closed maps again in subsequent chapters. Open maps
will play an important role in the study of quotient spaces (Chapter 8). Chapter 6 will
give a nice criterion for a map to be closed in terms of compactness of the domain of
the map (Corollary 6.1.10).
3.2. Convergent sequences
In this section we examine convergent sequences in topological spaces, rst in the
their own right and then with regards to their relation to continuous functions. Familiar
properties of sequences from Euclidean spaces no longer hold in this more general
setting, perhaps most striking being the non-uniqueness phenomena for the limit of a

46

3. CONTINUOUS FUNCTIONS AND CONVERGENT SEQUENCES

convergent sequence. A characterization of continuity in terms of sequences is given in


Theorem 3.2.11.
Definition 3.2.1. Let (X, T ) be a topological space and {xk }k X a sequence.
We say that the sequence {xk }k converges to x X, and write limk xk = x or just
lim xk = x, if for every neighborhood U of x there exists a k0 N such that for all
k k0 we obtain xk U .
Note that if one takes (X, TX ) = (Rn , TEu ) then this denition agrees with the
familiar denition of continuity in Euclidean spaces, compare to Theorem 1.3.3.
Example 3.2.2. Let X = R be equipped with the partition topology TP associated
to the partition
P = {[4a 2, 4a + 2 | a Z}.
Then the sequence {xk }k with xk = (1)k converges to any point x [2, 2 since this
is the smallest non-empty open set containing both 1 and 1.
Example 3.2.3. If X is equipped with the indiscrete topology, then any sequence
in X is convergent and its limit is any point in X. Conversely, if X is given the discrete
topology, then a sequence {xk }k is convergent to x X if and only if xk = x for all
suciently large k.
Example 3.2.4. Let {xk }k R be the sequence xk = 1/k. Determine whether or
not {xk }k converges in (R, T ) for the various choices of T below. If it converges, nd
its limits.
(a) T = TEu . In the Euclidean topology, convergence in the sense of Denition 3.2.1
is the same as in the usual sense from analysis (Denition 1.1.4 and Theorem
1.3.3). Thus limk xk = 0.
(b) T = Tp . In the particular point topology Tp , the sequence {xk }k does not
converge at all. To see this, note that any sequence {yk }k converging to x = p
has to equal x or p for all suciently large indices k. This follows from the
fact that {x, p} is a neighborhood of x. On the other hand, any sequence {yk }k
converging to p has to be constant, and equal to p, for all suciently large k
since {p} is a neighborhood of p. Since xk = 1/k is not of these two types of
sequences, it cannot be convergent.
(c) T = T p . Here {xk }k is convergent and lim xk = p. Namely, given any sequence
{yk }k R that converges to y = p, we must have yk = y for all large enough
k since {y} is a neighborhood of y. If {yk }k converges to p then, since the
only neighborhood of p is R, all that needs to be true is that yk R, a trivial
condition. Thus we see that in this example the limit of {xk }k exists and is
unique, though it may not be zero.
(d) T = T {p1 ,...,pn } . The excluded points topology T {p1 ,...,pn } is a variation of the
excluded point topology T p . A subset U R belongs to T {p1 ,...,pn } if either
U = R or else neither of the points p1 , ..., pn lies in U (compare to part (b) of
Exercise 2.5.3). Arguing as in the previous case, we nd that limk xk = pi

3.2. CONVERGENT SEQUENCES

47

for all i = 1, 2, ..., n. Thus, the sequence xk is convergent but has n dierent
limits.
(d) T = Tindis . As we already saw in Example 3.2.3, in the indiscrete topology we
nd that limk xk = x for every x R showing that {xk }k is convergent and
has innitely many limits.
Example 3.2.5. Let X = R be given the nite complement topology Tf c from
Example 2.2.10. Let {xk }k R be a sequence with xk = x whenever k = . Then
{xk }k converges to any point x R since every neighborhood of x contains all but
nitely many elements of {x1 , x2 , x3 , ...}.
Quite to the contrary, if we equip R with the countable complement topology Tcc
instead (Example 2.2.11), then {xk }k is not a convergent sequence anymore. For if
we had lim xk = x we would need xk = x for all suciently large indices k (Exercise
3.5.15). As we assumed that xk = x whenever k = , the latter condition fails.
As the examples above show, in topological spaces limits of sequences may not be
unique and moreover, their number can vary wildly. Thus, in passing from our accustomed setting of Euclidean spaces to the much more general ambience of topological
spaces, we have paid the price of loosing a familiar and desirable property: uniqueness
of the limit of a convergent sequence {xk }k Rn . However, with just a minor sacrice
of generality, we can regain this attribute. To see how, here is rst a denition.
Definition 3.2.6. A topological space X is called Hausdor if every two points
have disjoint neighborhoods. Said dierently, we require that for each pair of points
a, b X, a = b, there exist open sets Ua , Ub X with a Ua , b Ub and Ua Ub = .
Theorem 3.2.7. Let X be a topological space.
(a) If X Hausdor space and xk X is a convergent sequence, then the limit lim xk
is unique.
(b) If X is st countable and has the property that every convergent sequence xk X
has a unique limit, then X is Hausdor.
Proof. (a) Suppose that there are two (or more) limits for {xk }k , say a and
b. Since X is Hausdor, we can nd disjoint neighborhoods Ua and Ub of a and b
respectively. Let ka N be such that xk Ua for all k ka and kb N have the
property that xk Ub for all k kb . Then for all k max{ka , kb } we have that
xk Ua Ub , a contradiction since Ua Ub = .
(b) Given two arbitrary points a, b X with a = b, we need to nd two disjoint
open sets of which one contains a and the other contains b. Let Ba = {Uia X | i N}
and Bb = {Uib X | i N} be countable neighborhood bases at a and b respectively.
We dene new open sets Via and Vib as
a
a
a
Vka = U1 U2 ... Uk

and

b
b
b
Vkb = U1 U2 ... Uk .

These sets are open since they are nite intersections of open sets. Furthermore, notice
that Vja Uia and Vjb Uib for every j i. Clearly, each Via is a neighborhood of a
and each Vib is a neighborhood of b.

48

3. CONTINUOUS FUNCTIONS AND CONVERGENT SEQUENCES

If Vka Vkb = for some k, we are done. So suppose instead that Vka Vkb = for all
k N. Let xk Vka Vkb be an arbitrary point. This yields a sequence in X which we
a
claim converges to a. To see this, let U be any neighborhood of a and nd an Uka Ba ,
a
ka N, such that Uka U . Then Vka U for every k ka , in particular xk U for
every k ka , showing that lim xk = a. Repeating this same argument for b shows also
that lim xk = b. This is a contradiction since by assumption all convergent sequences
in X have a unique limit. We are thus forced to conclude that there is some k N for
which Vka Vkb = , giving us disjoint neighborhoods of a and b, as needed.
The rst countability condition from part (b) of Theorem 3.2.7 is necessary and
cannot be weakened as the next example shows.
Example 3.2.8. Consider X = R equipped with the countable complement topology. We claim that every convergent sequence {xk }k X has a unique limit. Suppose
not, that is suppose that lim xk = a and lim xk = b with a = b. Let Ua be the open set
Ua = R {xi | i N and xi = a}.
Clearly a Ua and so there must be some na N such that xn Ua for all n na .
But then xn = a for all n na since xn Ua {xi | i N} = {a}. A similar argument
shows that for some nb N all xn = b for n nb . But then xn = a and xn = b
for n max{na , nb } which is impossible since a = b. On the other hand, X is not
Hausdor since every two non-empty open sets have nontrivial intersection. Compare
to Exercise 2.5.22.
Recall that a subset A Rn (with the Euclidean topology) is closed if and only
if it contains the limits of all its convergent sequences (this was the original denition
of a closed subset of Rn , Denition 1.2.1). This characterization of closed sets is only
partly true in general topological spaces.
Theorem 3.2.9. Let X be a topological space and A a subset of X.
(a) If A is a closed then it contains the limits of all its convergent sequences.
(b) If X is rst countable and A contains the limits of all of its convergent sequences, then A is closed.
Proof. (a) Let {xk }k A be a sequence with limit x. If x A then x X A
/
which is an open set. By convergence of {xk }k there must be some k0 N such that
xk X A for all k k0 . This is impossible since xk A and A (X A) = .

(b) We will show that A is closed by exhibiting that A = A. Suppose this were not

true. Then A A would be nonempty. Let x A A and let Bx = {Ui X | i N}


be a countable basis at x and dene Vj as
Vj = U1 U2 ... Uj .
Note that the sets Vj are open, that the inclusion Vj Ui holds for all j i and that
each Vj contains x. We claim next that each set Vi A must be non-empty. For if not

then A Vj would be a closed set smaller than A and containing A, a contradiction.


Thus we can pick an element xk Vk A. But now {xk }k must converge to x since

3.2. CONVERGENT SEQUENCES

49

if V is any neighborhood of x then there must be an index k0 such that x Uk0 V


and thus xk Vk Uk0 V for all k k0 . Since xk A and lim xk = x we conclude

that x A. Therefor A = A.
That the rst countability condition from part (b) of the preceding theorem cannot
be dropped, is illustrated by the next example.
Example 3.2.10. Let X = R be equipped with the countable complement topology
and let A X be the set A = X {0}. Notice that A is not closed (since X A = {0}
is not open). But A contains the limits of all of its convergent subsequences. To see
this we only need to show that no sequence {xk }k A can converge to 0. This is
easy to see since the set U = X {x1 , x2 , ...} is an open set which contains zero but
no element of the sequence xn . Thus A contains the limits of all of its convergent
sequences. Note that this, in conjunction with Theorem 3.2.9, shows that (R, Tcc )
cannot be rst countable and hence neither second countable.
We conclude this section by examining the relation between continuous functions
and convergent sequences. As the reader may guess by now, with the assumption of
rst countability, the relation between the two is just as in Euclidean space (Theorem
1.3.3).
Theorem 3.2.11. Let f : X Y be a function between two topological spaces.
(a) If f is continuous at x X and if {xk }k X is a convergent sequence with
lim xk = x, then {f (xk )}k Y is also a convergent sequence with lim f (xk ) =
f (x).
(b) If X is rst countable and f has the property that limf (xk ) = f (x) for all
sequences {xk }k X converging to x, then f is continuous at x.
Proof. (a) Suppose that f is continuous at x and let {xk }k X be a convergent
sequence with limit x. To show that {yk }k is a convergent sequence with limit y = f (x),
where yk = f (xk ), let V Y be an arbitrary neighborhood of y. Since f is continuous,
the set U = f 1 (V ) X is a neighborhood of x and by convergence of {xk }k to x,
there is some k0 N such that xk U for all k k0 . But then yk V for all k k0 ,
showing that lim f (xk ) = f (x).
(b) Let V be a neighborhood of f (x) and set U = f 1 (V ). We seek to show that
U contains a neighborhood of x. Suppose that this were not so. In that case, let
Bx = {Ui | i N} be a neighborhood basis around x and dene, as in the proofs of
Theorem 3.2.7 and 3.2.9, the open sets Vj , j N as Vj = U1 Uj . By assumption,
the sets (X U ) Vk are nonempty and so we can pick points xk (X U ) Vk .
The sequence {xk }k is easily seen to converge to x and thus by assumption {f (xk )}k
must converge to f (x). This however is impossible since V is open and f (x) V while
f (xk ) V . Consequently, U must contain a neighborhood Ux of x and hence f is
/
continuous at x.

50

3. CONTINUOUS FUNCTIONS AND CONVERGENT SEQUENCES

3.3. Uniform convergence of functions


This section is largely of a utilitarian nature in that its main result, Theorem 3.3.2,
shall be used as the basis for a number of arguments later in the book. The rst
instance of this is in the next section where Theorem 3.3.2 is used in the construction
of space lling curves. For another instance, see Section 9.1 on the Tietze extension
theorem. The reader may skip this section on a rst pass, and return to it as needed.
The main question we want to address here is the following: Suppose that (X, TX )
is a topological space and gm : X R (where R is given the Euclidean topology) is a
sequence of functions indexed by m N such that for every xed x X, the sequence
{gm (x)}m converges. In such a situation we can dene a new function g : X R by
setting g(x) = limm gm (x). We this in mind we ask: Under what conditions on the
functions gm is the function g continuous?
It is easy to see that even if all gm are continuous, the limiting function g need not
be. For example, taking X = [0, 1] and letting gm (x) = xm , we obtain

g(x) =

0
1

;
;

x [0, 1 ,
x = 1,

which is not continuous. The extra condition needed to ensure continuity of g is one
that compares how gm (x) converges to g(x) for various x X.
Definition 3.3.1. Let (X, TX ) be a topological space and let gm : X R, m N
be a sequence of continuous functions. We say that the sequence gm converges uniformly
to the function g : X R if for every > 0 there exists an index m0 N so that all
m m0 imply the inequality |g(x) gm (x)| < for all choices of x X.
Saying that gm (x) converges to g(x) for every x X means that for every xed
x X and for every > 0, there is some integer m0 (x) N with |gm (x) g(x)| <
for all m m0 (x). The extra condition incorporated in the denition of uniform
convergence of gm to g is that m0 (x) can be chosen to be the same for all x X (i.e.
we can take m0 (x) to be a constant function). With this understood, here is the main
result of this section.
Theorem 3.3.2. If gm : X R, m N is a sequence of continuous functions
converging uniformly to the function g : X R, then g is also continuous.
Proof. Let x0 X be any point and let V R be any neighborhood of g(x0 ).
Find an > 0 so that g(x0 ) , g(x0 ) + V . By the uniform convergence
property, we can nd an index m0 so that for all m m0 and all x X we obtain

|g(x) gm (x)| < 3 . Pick any index m m0 , then by continuity of gm , there must exist

a neighborhood U of x0 such that gm (U ) gm (x0 ) 3 , gm (x0 ) + 3 . Given any x U ,

3.4. SPACE FILLING CURVES

51

we nd
|g(x) g(x0 )| = |g(x) gm (x) + gm (x) gm (x0 ) + gm (x0 ) g(x0 )|
|g(x) gm (x)| + |gm (x) gm (x0 )| + |gm (x0 ) g(x0 )|

< + + = ,
3 3 3
showing that g(U ) g(x0 ) , g(x0 ) + V . From this, continuity of g at x0 follows
and since x0 X was arbitrary, g is continuous.
3.4. Space lling curves
In Section 3.1 we asked one of the main questions in topology, the questions of
nding eective ways to distinguish topological spaces up to homeomorphism (Question
3.1.20). While this question is much too dicult to address in full generality, one can
hope for a partial answer by restricting the choice of topological spaces considered. In
this section we ask such a more specic question, one which we shall return to on a
number of occasions later in the text and one which, unlike Question 3.1.20, we shall
be able to answer (the curious reader may skip ahead to Corollaries 7.2.8 and Theorem
??).
Question 3.4.1. For which values of m, n N can the n-dimensional Euclidean
space (Rn , TEu ) be homeomorphic to the m-dimensional Euclidean space (Rm , TEu )?
The goal of this section is to convince the reader that this question may have an
answer dierent from the expected one of Only if n = m. We shall attempt to do so
by demonstrating that there exist continuous surjective functions from the unit segment
[0, 1] R to the unit n-dimensional cube [0, 1]n Rn , each equipped with the relative
Euclidean topology. Such functions are referred to as space lling curves. Hence, if
a curve can ll out n-dimensional space, albeit only an n-dimensional cube, perhaps
there also exist homeomorphisms from R to Rn opening the possibility for unexpected
answers to Question 3.4.1. Wed like to emphasize that we do not claim that answers
besides m = n exist, we merely point out that one needs to exercise caution before
jumping to conclusions with regards to homeomorphisms between spaces of dierent
dimensions.
While there is a multitude of space lling curves, we shall utilize a particular construction and dene only one such curve in detail. However, many others can be
obtained by varying the pattern of the curve below, some additional patterns are given
in Figure 5.
We will start by constructing a continuous surjection f : [0, 1] [0, 1]2 . This is
done by dening piecewise linear and continuous functions fm : [0, 1] [0, 1]2 for each
m N and then letting f (x) = limm fm (x). Of course, we will have to argue that
this latter limit exists and that the thus obtained function f is indeed continuous and
surjective. The case of surjective maps f : [0, 1] [0, 1]n for n 3 is then easily
obtained from this example.

52

3. CONTINUOUS FUNCTIONS AND CONVERGENT SEQUENCES

To dene the functions fm : [0, 1] [0, 1]2 , m N, we proceed as follows: Start


with m = 1 and divide [0, 1] into four segements Ii1 , i = 1, ..., 4 of equal length. Thus
1
1
I1 = [0, 4 ],

1
1
I2 = [ 4 , 2 ],
4

2
1
I1 = [ 4 , 3 ],
4

1
I1 = [ 3 , 4 ].
4 4

Similarly divide the unit square [0, 1]2 into four congruent squares Ji1 , i = 1, ..., 4 by
dening
1
1
J1 = [0, 2 ] [0, 1 ],
2

1
J2 = [0, 1 ] [ 1 , 1],
2
2

1
J3 = [ 1 , 1] [ 1 , 1],
2
2

1
1
1
J4 = [ 2 , 1] [0, 2 ].

We then let f1 : [0, 1] [0, 1]2 be any continuous function that maps Ii1 into Ji1 ,
1
1
1
1
I1 I2 I3 I4

2
I1

f1

......

2
I16

f2

1
J2

1
J3

1
J1

1
J4

2
J6

2
J7

2
J10

2
J11

2
J5

2
J8

2
J9

2
J12

2
J4

2
J3

2
J14

2
J13

2
J1

2
J2

2
J15

2
J16

Figure 2. The subdivision of [0, 1] and [0, 1]2 for the cases of m =
1, 2. The arrows inside of the two copies of [0, 1]2 indicate our choice
of ordering of the subdivision squares Jim (m = 1, 2) with respect to i.
Similar orderings are to be applied to Jim for m 3.
i = 1, . . . , 4 and whose image in [0, 1]2 looks as in Figure 4. We refer to the image of
f1 as the pattern of the curve to be constructed. Note that Im(f1 ) Ji1 = for every
i = 1, . . . , 4.
To obtain f2 : [0, 1] [0, 1]2 we proceed in much the same manner. This time we
i
divide [0, 1] into the 16 congruent segments Ii2 , i = 1, ..., 16 with Ii2 = [ i1 , 16 ] and
16
likewise, we divide [0, 1]2 into 16 congruent squares Ji2 , i = 1, ..., 16. These subdivision
squares of [0, 1]2 are chosen as
2
J1 = [0, 1 ] [0, 1 ],
4
4

1
2
J2 = [ 4 , 2 ] [0, 1 ],
4
4

1
1
2
J3 = [ 4 , 2 ] [ 4 , 2 ],
4
4

1
2
J4 = [0, 1 ] [ 4 , 2 ],
4
4

3
2
J5 = [0, 1 ] [ 2 , 4 ],
4
4

2
J6 = [0, 1 ] [ 3 , 4 ],
4
4 4

1
3
2
J7 = [ 4 , 2 ] [ 4 , 4 ],
4
4

1
2
2
J8 = [ 4 , 2 ] [ 4 , 3 ],
4
4

3
3
2
3
2
2
2
2
2
J9 = [ 2 , 4 ] [ 2 , 4 ], J10 = [ 4 , 3 ] [ 3 , 4 ], J11 = [ 3 , 4 ] [ 4 , 4 ], J12 = [ 3 , 4 ] [ 4 , 3 ],
4
4
4
4 4
4 4
4
4 4
4
4
2
2
3
4
2
2
2
2
J13 = [ 3 , 4 ] [ 1 , 4 ], J14 = [ 4 , 3 ] [ 1 , 2 ], J15 = [ 2 , 4 ] [0, 1 ], J16 = [ 3 , 4 ] [0, 1 ].
4
4
4
4 4
4
4
4
4

3.4. SPACE FILLING CURVES

53

Figure 2 illustrates our choices of Ii1 , Ii2 , Ji1 and Ji2 for the various indices i. The
function f2 : [0, 1] [0, 1]2 is chosen so that f2 (Ii2 ) Ji2 for each i = 1, ..., 16 and so
that its image looks as in Figure 4. Notice again that our denition of f2 implies that
Im(f2 ) Ji2 = for all i = 1, ..., 16.
From here on we proceed inductively on m to dene fm : [0, 1] [0, 1]2 , m 3.
We do so by subdividing [0, 1] into 4m congruent intervals Iim , i = 1, ..., 4m and by
subdividing [0, 1]2 into 4m congruent squares Jim , i = 1, ..., 4m . The function fm is then
chosen subject to the condition fm (Iim ) Jim for each i = 1, ..., 4m and is otherwise
dened as in Figure 3. The images of the thus constructed functions fm are shown in
Figure 4 for m = 1, 2, 3, 4.
The two key features of the functions fm : [0, 1] [0, 1]2 obvious from their construction, are:
(a) If fm (x) Jim for some i = 1, ..., 4m , then fm+k (x) Jim for all k N.
(3.1) (b) For every m N and for every i = 1, ..., 4m we obtain Im(fm ) Jim = .
For the remainder of this section, we shall refer to these simply as Properties (a) and
(b). The attentive reader will notice that the proofs of the Lemmas 3.4.2, 3.4.3 and
3.4.4 below, will be carried out only by relying on these two properties rather than on
any other specics about the construction of fm .
Lemma 3.4.2. For every x [0, 1] and for every k N we obtain the inequality

2
d2 (fm (x), fm+k (x)) m ,
2
where d2 is the Euclidean metric from (1.1). Consequently, the sequence {fm (x)}m is
convergent for every x [0, 1].
Proof. The stated inequality for d2 (fm (x), fm+k (x)) follows from Property (a) of

2
(3.1) and from the observation that 2m is the length of the diagonal of Jim and is
therefore the largest distance between any two points in Jim . This inequality shows
that the sequence {fm (x)}m is a Cauchy sequence in [0, 1]2 for any choice of x [0, 1]
and is thus convergent by Theorem 1.1.6.
We are now in the position to dene our space lling function f : [0, 1] [0, 1]2 as
(3.2)

f (x) = lim fm (x).


m

2
Applying the limit as k goes to innity to the inequality d2 (fm (x), fm+k (x)) 2m from
Lemma 3.4.2, we obtain the useful inequality

2
(3.3)
d2 (fm (x), f (x)) m
x [0, 1].
2
This step is justied by continuity of the function x d2 (y, x) for any choice of y R2
(Exercise 3.5.10) and by Theorem 1.1.5. Our rst order of business is to show that f
is continuous.

3. CONTINUOUS FUNCTIONS AND CONVERGENT SEQUENCES

Im(fm )

Im(fm )

Im(fm )

Im(fm )

54

Im(fm )

Im(fm+1 )

Figure 3. The function fm+1 : [0, 1] [0, 1]2 is constructed from the
function fm : [0, 1] [0, 1]2 as follows: Take 4 copies of the unit square
[0, 1]2 and shrink each by a factor of 2. Arrange these 4 shrunken
squaresinto a new square [0, 1]2 as indicated on the right gure above.
Thus, 2 of the shrunken squares are placed side by side to ll the upper
half of [0, 1]2 while the remaining 2 shrunken squares are rst rotated
by 90 and the placed into the bottom half of [0, 1]2 . The image of
fm+1 in [0, 1]2 is then the union of the 4 shrunken copies of the image of
fm , connected suitably by small straight line segments (indicated in red)
so as to make fm+1 continuous. This gure also explains the indexing
of the subdivision squares Jim+1 , i = 1, ..., 4m+1 . Namely, the rst 4m
of these squares are the 4m squares Jim from the previous stage of the
construction, shrunk by a factor of 2 and placed in the lower left corner
of [0, 1]2 . One repeats this process three more time placing three more
copies of the 4m squares Jim into the upper left, upper right and nally
lower right corner of [0, 1]2 . The indices i in Jim are increased by 4m ,
2 4m and 3 4m respectively, in this process.
Lemma 3.4.3. The function f : [0, 1] [0, 1]2 constructed in (3.2), is continuous.
Proof. Inequality (3.3) shows the sequence {fm }m to converge uniformly to f .
The claim of the lemma is now immediate from Theorem 3.3.2.
Next we check that f is surjective.
Lemma 3.4.4. The function f : [0, 1] [0, 1]2 dened in (3.2), is surjective.

3.4. SPACE FILLING CURVES

Im(f1 )

Im(f2 )

Im(f3 )

55

Im(f4 )

Figure 4. The images of the functions fm : [0, 1] [0, 1]2 for m =


1, 2, 3, 4. These are constructed following the recipe described in Figure
3.
Proof. Let y0 [0, 1]2 be any point, we need to show that there exists a point
m
x0 [0, 1] with f (x0 ) = y0 . Since 4 Jim = [0, 1]2 we can nd for each m N, an index
i=1
im {1, ..., 4m } such that y0 Jim . Let xm [0, 1] be any point with the property
m
that fm (xm ) Jim . Note that such a point xm has to exist by Property (b) from (3.1).
m

2
Then d(fm (xm ), y0 ) 2m .
Since the sequence {xm }m [0, 1] is bounded, it must have a convergent subsequence {xk(m) }m with m k(m) an increasing function. Let x0 be the limit of
{xk(m) }m . By continuity of f , the sequence {f (xk(m) }m has to converge to f (x0 ) (Theorem 1.1.5). Thus, for any > 0 there exists an m0 N such that m m0 implies that

d(f (xk(m) ), f (x0 )) < 3 . For this same , let us chose a value m1 N such that 2m2 < 3 .
1

Since k(m) m, choosing any m m1 will imply that the inequality 2k(m) < 3 still
holds. Putting these observations together, we nd that for any m max{m0 , m1 },
the next inequality holds:
d(f (x0 ), y0 ) d(f (x0 ), f (xk(m) )) + d(f (xk(m) ), fk(m) (xk(m) )) + d(fk(m) (xk(m) ), y0 ),

< + + ,
3 3 3
= .
The bound on the second term on the right hand side of the rst line, comes from
Equation (3.3). Since d(f (x0 ), y0 ) < for every > 0, it must be that d(f (x0 ), y0 ) = 0
and therefore that f (x0 ) = y0 as desired.
Having constructed a continuous surjection f from [0, 1] onto [0, 1]2 , we can easily
obtain surjections from [0, 1] to [0, 1]n . To do this, let us write f (t) = (1 (t), 2 (t))
with 1 , 2 : [0, 1] [0, 1] being the coordinate functions of f . Let Fn : [0, 1] [0, 1]n
be dened inductively as F2 = f and
(3.4)

Fn+1 (t) = (1 (t), Fn (2 (t)))

Since 1 and 2 are continuous, Fn+1 is also continuous given that Fn is continuous. An
induction on n then shows that all Fn are continuous since F2 is. Also by induction on

56

3. CONTINUOUS FUNCTIONS AND CONVERGENT SEQUENCES

n one can show that Fn surjects onto [0, 1]n . For example, when n = 3, let (t1 , t2 , t3 )
[0, 1]3 be any point. By surjectivity of F2 , we can nd an s [0, 1] such that F2 (s) =
(t2 , t3 ). By surjectivity of (1 , 2 ), we can then nd a t [0, 1] such that (1 (t), 2 (t)) =
(t1 , s). But then
F3 (t) = (1 (t), F2 (2 (t))) = (t1 , F2 (s)) = (t1 , t2 , t3 ),
showing that F3 is surjective. The case of n 3 is handled similarly and is left as an
exercise.
Our construction of the functions fm : [0, 1] [0, 1]2 , m N started by choosing
the image of f1 in [0, 1]2 more or less arbitrary, and then constructing the functions
fm , m 2 according to the recipe in Figure 3. In this sense the functions fm , m 2
are pinned down by our choice of f1 , that is they are determined by the choice of
the pattern of the space lling curve. As long as the construction of the functions fm
satises Properties (a) and (b) from (3.1), we have complete freedom for choosing our
pattern for f1 . By varying the pattern, we may construct a multitude of space lling
curves. Figure 5 shows several alternative patterns.
3.5. Exercises
3.5.1. For a function f : X Y between sets X and Y , and for any family of
subsets Vi Y , i I, show that
(a) f 1 (iI Vi ) = iI f 1 (Vi ).
(b) f 1 (iI Vi ) = iI f 1 (Vi ).
3.5.2. Verify whether or not the given function f : R R is continuous. Here the
domain of f is equipped with the topology T1 and the codomain with the topology T2 .
(a) f (x) = sin x, T1 = Tcc , T2 = TEu .
(b) f (x) = cos x, T1 = Tp with p = 0, T2 = Tp with p = 1.
(c) f (x) = x3 , T1 = T p with p = 0, T2 = TF ,p with p = 0.
3.5.3. For R let f : R R be the function f (x) = x + . If both the domain
and codomain of f are equipped with the nite complement topology, for which R
is f continuous?
3.5.4. Consider the open ball Bx (r) Rn equipped with the relative Euclidean
topology inherited from Rn . Show that for every choice of x Rn and radius r > 0,
Bx (r) is homeomorphic to (Rn , TEu ). (Hint: Show rstly that the tangent function
tan : , R is a homeomorphism and use this to solve the exercise for n = 1.
2 2
Apply this in conjunction with polar coordinates to the case of n 2.)
3.5.5. Construct an explicit homeomorphism f : Dn [1, 1]n . Here
Dn = {x Rn+1 | x2 + + x2 1},
1
n
[1, 1]n = {x Rn | 1 xi 1, i = 1, . . . , n},
and each of the spaces comes with the relative Euclidean topology.

3.5. EXERCISES

57

Figure 5. Each of the three rows of gures depicts the images of f1 , f2


and f3 for the choice of pattern given by the rst picture in the row.
3.5.6. Show that a function f : X Y is continuous if and only if f 1 (Int(B))
Int(f 1 (B)) for every subset B Y .
3.5.7. Verify the claim from Example 3.1.8. Specically, let (X, dX ) and (Y, dY ) be
two metric spaces and let X and Y be equipped with the associated metric topologies.
Show that a function f : X Y is continuous at x X if and only if for every > 0
there exists a > 0 such that dx (x , x) < implies dY (f (x ), f (x)) < .

58

3. CONTINUOUS FUNCTIONS AND CONVERGENT SEQUENCES

3.5.8. Verify the claims from Example 3.1.16: Show that f : (R, T ) (R, T ) given
by f (x) = x3 is a homeomorphism for any choice of T {TEu , Tp , T p , Tf c , Tcc , TF ,p }
with p being either 0 or 1.
3.5.9. Find a formula describing the stereographic projection S : S n {S} Rn
from the south pole S = (0, 0, . . . , 0, 1)of the unit n-sphere in Rn+1 (see Example
3.1.19). Verify that it denes a homeomorphism between S n {S} and Rn .
3.5.10. Let (X, d) be a metric space and Td the associated metric topology on X.
Show that for any xed choice of y X, the function fy : X [0, given by
fy (x) = d(x, y) is a continuous function. Here [0, carries the relative Euclidean
topology.
3.5.11. Let (X, dX ) and (Y, dY ) be two metric spaces. Recall from Example 3.1.18
that a function f : X Y is called an isometry if dY (f (a), f (b)) = dX (a, b) for any
pair of points a, b X. To complete the claims from Example 3.1.18, prove that
(a) Every isometry is an injective function.
(b) Show that the inverse of a bijective isometry is again an isometry.
3.5.12. Show that the notions of open mapand closed map(Denition 3.1.21)
are independent. Specically, consider the functions f : R R and g : R2 R given
by
f (x) = x2 , and g(x, y) = x.
Show that f is a closed map that is not open, and show that g is an open map that is
not closed.
3.5.13. Show by example that a local homeomorphism is not necessarily a closed
map (compare to part (c) of Proposition 3.1.22).
3.5.14. Determine whether or not the sequence {xk }k R converges when R is
given the topology T . If the sequence converges, nd all of its limits.
(a) xk = (1)k and T = TP with the partition P given by P = {[n, n + 1 | n Z}.
(b) xk = sin k and T = Tll .
(c) xk = ln k and T is the topology generated by the subbases
S = { n, n | n N}.
3.5.15. Consider R with the countable complement topology Tcc from Example
2.2.11. Show that a sequence {xk }k R converges to x R if and only if there is an
integer k0 such that xk = x for all k k0 .
3.5.16. Show by example that the condition of rst countability cannot be dropped
from part (b) of Theorem 3.2.11. Specically, nd an example of a function f : X Y
that is not continuous at a point x X, but has the property that if {xk }k X is a
sequence converging to x, then the sequence {f (xk )}k converges to f (x).

3.5. EXERCISES

59

3.5.17. Show that the maps Fn : [0, 1] [0, 1]n dened in (3.4), are surjective for
each n 4. The case of n = 3 was addressed in the paragraph following equation
(3.4).

CHAPTER 4

Separation axioms

eparation axioms, introduced in this section, serve as a measure of how ne


or how coarse a topology T is in relation to the set X on which it is dened.
Section 4.1 formally introduces the 7 types of separation axioms and studies
some of the their most basic properties. Section 4.2 is devoted to examples,
and examines the topological spaces from Section 2.2 in light of the Separation Axioms.
Section 4.3 introduces the all important notion of a topological invariant, and give
several examples. Section 4.4 provides an illustration of how topological invariants are
typically applied to tell apart topological spaces up to homemorphism. The chapter
concludes with a proof of the Urysohn Lemma in Section 4.5.
4.1. Degrees of separation
In Chapter 2 we encountered two notions of size of a topology T on a set X, those
of separability (Denition 2.3.5) and second countability (Denition 2.4.8). These two
yardsticks measure the sizeof the topology T , without any regard to X itself. The
notions of sizeof a topological space that we shall encounter in this section, employs
an interplay between X and T . For instance, we shall ask whether T is large enough
so that any pair of distinct points in X has a pair of disjoint neighborhoods. If this is
not the case, we can think of T as being small in relation to X.
Definition 4.1.1. We say that two subsets A and B of the topological space (X, T )

are separated if A B = = A B.
For example, closed and disjoint sets A, B X are separated.
Definition 4.1.2. Let (X, T ) be a topological space. Then
(a) X is called T0 or Kolmogorov if for any two points x, y X there exists an open
set U X such that either x U and y U , or y U and x U .
/
/
(b) X is called T1 or Frchet if for any two points x, y X there exist open sets
e
Ux , Uy X such that x Ux Uy and y Uy Ux .
(c) X is called T2 or Hausdor if for any two points x, y X there exist disjoint
open sets Ux , Uy X such that x Ux and y Uy (Denition 3.2.6).
(d) X is called T3 if for any closed set A X and any point x X A there are
disjoint open sets UA , Ux X such that A UA and x Ux .
(e) X is called T3 1 if for any closed set A X and any point x X A there
2
exists a continuous function f : X [0, 1] with f |A 0 and f (x) = 1.
61

62

4. SEPARATION AXIOMS

(f) X is called T4 if for any two disjoint closed sets A, B X there are disjoint
open sets UA , UB X with A UA and B UB .
(g) X is called T5 if for any two separated sets A, B X (see denition 4.1.1 above)
there are disjoint open sets UA , UB X with A UA and B UB .
From the properties T0 T5 , we derive three additional denitions. Namely, we shall
that that
(h)
(i)
(j)
(k)

X
X
X
X

is
is
is
is

Regular if it is both T0 and T3 .


Completely Regular if it is both T0 and T3 1 .
2
Normal if it is both T1 and T4 .
Completely Normal if it is both T1 and T5 .

The various properties T0 T5 are not entirely unrelated. Their various dependencies
are easily unraveled with the help of the next lemma.
Lemma 4.1.3. Let X be a T1 space, then for any point x X, the set {x} is closed.
Proof. For any point y X {x} let Uy be a neighborhood of y not containing
x. Then
X {x} =
Uy ,
yX{x}

showing that X {x} is a union of open sets and hence open.


This lemma and Denition 4.1.2 yield all but one of the following implications:

(4.1)

T1
T2
T1 and T3
T3 1
2
T1 and T4
T1 and T4
T5

=
=
=
=
=
=
=

T0 ,
T1 ,
T2 ,
T3 ,
T3 1 (Corollary 4.5.4),
2
T3 ,
T4 .

For example, to see that T3 1 implies T3 , let X be a T3 1 space and let A X be a


2
2
closed subset and x X A any point. The T3 1 property guarantees the existence of
2
a map f : X [0, 1] with f |A 0 and f (x) = 1. Then the sets UA = f 1 ([0, 1/2 ) and
Ux = f 1 ( 1/2, 1]) are open and disjoint sets containing A and x respectively, showing
that X is a T3 space.
The next theorem shows that the separation properties Ti are inherited by a subspace, with some subtleties in the case of T4 and T5 .
Theorem 4.1.4. Let X be a topological space and let Y X be a subspace of X.
(a) If X is Ti for some i {0, 1, 2, 3, 3 1 } then so is Y .
2
(b) If Y is closed and X is Ti for some i {4, 5}, then Y is also Ti .

4.2. EXAMPLES

63

Proof. (a) For the cases of i = 0, 1, 2 let x, y Y be two distinct points. If Ux


and Uy are the neighborhoods of x and y in X showing up in the denition of Ti for
X, then Ux Y and Uy Y are neighborhoods of x and y in Y that can be taken to
verify property Ti for Y itself.
For property T3 , let A Y be a closed subset and let x Y A be any point.
Then there is a closed subset A X with A = A Y and clearly x A . Let UA and
/
Ux be disjoint neighborhoods of A and x in X, then UA = Y UA and Ux = Ux Y
are disjoint neighborhoods of A and x in Y .
For property T3 1 , let again A Y be a closed subset of Y and x Y A any point.
2
Let A X be a closed subset of X with A = A Y and note that x A . By the
/
T3 1 property of X, there exists a function f : X [0, 1] with f |A 0 and f (x) = 1.
2
But then f : Y [0, 1] given by f = f |Y is continuous (part (c) of Theorem 3.1.11)
and f |A 0 and f (x) = 1. This shows that Y also has the T3 1 property.
2
(b) For property T4 , let A, B Y be two disjoint closed subsets of Y . Since Y is
assumed to be closed in X, we see that A, B are also closed subsets of X (Exercise
2.5.17). But then they can be separated by neighborhoods U, V in X and are therefore
separated by U Y and V Y in Y . A similar argument works for T5 , the details are
left to the reader (Exercise 4.6.1).
The hypothesis of Y being closed in part (b) of Theorem 4.1.4 cannot be removed,
as the next example shows.
Example 4.1.5. Consider X = R and for a choice of two distinct points p1 , p2 X,
let T be the topology
T = {U X | U = , or U = X, or p1 U and p2 (X U )}.
Then X does not contain disjoint closed non-empty subsets rendering (X, T ) trivially
a T4 space. Let Y be the subspace of X given by Y = X {p2 }, note that Y is not
closed in X. In Y there are many disjoint non-empty closed subsets, but there are no
disjoint non-empty open subsets, showing that Y cannot be T4 .
The following statement is a direct consequence of Theorem 4.1.4.
Corollary 4.1.6. A subspace of a (completely) regular space is (completely) regular. A closed subspace of a normal space is a normal space.
4.2. Examples
Example 4.2.1. A space that is T3 , T4 , T5 but not T0 , T1 , T2 . Let X = R be given
the partition topology TP associated to the partition P = {[2a, 2a + 2 | a Z}. This
space is not a Ti space for i = 0, 1, 2.
To see this, let x = 0 and y = 1. The smallest open set containing x is [0, 2 which
incidentally also contains y. Thus (R, TP ) is not T0 and hence neither T1 or T2 .
Observe that in this example the closed subsets and R agree with the open subsets.
Thus, given any closed subset A R and a point x R A, the open sets UA and Ux

64

4. SEPARATION AXIOMS

from separation axiom T3 can be taken to be UA = A and Ux = X A. Accordingly,


(R, TP ) is T3 .

Let A, B R be two separated sets. Then A B = implies that B R A. But


is open and thus also closed in the partition topology and thus B = R A.

X A
Given this, the two open sets UA and UB from the separation axiom for T5 can be taken

to be UA = A and UB = R A. Recall that the T5 axiom implies the T4 axiom, cf.


(4.1).
Example 4.2.2. A space that is T0 , T1 but not T2 , T3 , T4 , T5 . Let X = R have the
nite complement topology and let x, y R be any two distinct points. Then the two
sets Ux = X {y} and Uy = X {x} are neighborhoods of x and y respectively with
y Ux and x Uy . Thus (R, Tf c ) is T1 and hence T0 . It is not however T2 T5 since
/
/
no disjoint non-empty open subsets of R exist in this topology, while there is a plenty
of non-empty closed and disjoint subsets of R.
Example 4.2.3. A space that is T0 but not T1 T5 . Consider the included point
topology Tp on R with p = 0. There are no disjoint non-empty open sets in this
topology and so (R, Tp ) cannot be T2 T5 . It is also not T1 since every non-empty open
set has to contain p. It is however T0 since given any two points x, y, with x = p, the
set {y, p} (or simply {p} if y = p) is an open set containing y but not x.
Example 4.2.4. A space that is T0 , T4 , T5 but not T1 , T2 , T3 . Consider X = R with
the excluded point topology T p with p = 0. Given any two distinct points x, y R,
one of them, say x, is not equal to 0. But then the set Ux = {x} is a neighborhood of
x not containing y. Thus (R, T p ) is T0 . However, it is not T1 (and thus also not T2 )
since taking y = 0, the only open set containing y is R which of course also contains x.
The closed subsets of R in this topology are those containing p = 0. Thus, given
any closed set A R, the only neighborhood of A is R itself. This immediately shows
that (R, T p ) cannot be T3 . It also shows that it is T4 since no pair of non-empty disjoint
closed subsets of R exist.

Finally, let A, B R be two separated sets. Note that A = A{0} and B = B{0}.

Then A B = implies that 0 B while A B = implies that 0 A But then A


/
/
and B are both already open sets and so we can set UA = A and UB = B showing that
(R, T p ) is T5 .
Example 4.2.5. A space that is T0 , . . . , T5 . Let X = R be given the lower limit
topology Tll . Then X is Ti for every i = 0, 1, ..., 5. To see this is suces to show that it
is T2 and T5 since according to (4.1), all the other separation axioms follow from these.
To see that X is Hausdor, pick two points x, y X with x < y. Dene
Ux = [x, y

and

Uy = [y, y + 1

These are disjoint neighborhoods of x and y respectively showing that (R, Tll ) is T2 .

To verify separation axiom T5 , let A, B X be two separated sets. Then X B

is an open set and so we can, for each a A X B, nd an xa X such that


(since the half-open intervals are a basis for the topology). Dene
[a, xa X B

4.3. TOPOLOGICAL INVARIANTS

65

the open set UA as UA = aA [a, xa and in a similar vein dene also UB . These
are open sets (being a union of open sets) and contain A and B respectively. If we
had UA UB = , then there would have to be some a A and some b B so that
[a, xa [b, xb = . Suppose that a < b (the case b < a is treated analogously), then
b [a, xa (since b is the smallest element of [b, xb ) but this would yield a contradiction

since b B and [a, xa X B. Therefore UA and UB must be disjoint, showing that


(R, Tll ) is T5 .
Example 4.2.6. Metric spaces are always T0 , . . . , T5 . Let (X, d) be a metric space
and consider X equipped with the associated metric topology Td .
(X, Td ) is always T2 (and hence also T0 and T1 ) since, given any two distinct points
x, y X, the sets Bx (r/2) and By (r/2) are disjoint neighborhoods of x and y, where
r = d(x, y).
To see that it is also T3 T5 , it suces to show that it is T5 , see (4.1). Thus let
A, B X be two separated set. Then for every point a A there exists some ra > 0

such that Ba (ra ) B = . For if not, part (c) of Lemma 2.3.2 would imply that a B
= . Let Ua = Ba (ra /2) and dene UA = aA Ua .
contradicting the assumption A B
In a similar vein dene also rb , Ub and UB . It is clear that UA and UB are open sets
containing A and B respectively. It remains to show that they are disjoint. If not, we
could nd a point p UA UB . But then there would be points a A and b B such
that p Ua Ub showing further that
d(a, b) d(a, p) + d(p, b) < ra /2 + rb /2.
For concreteness, suppose that ra rb . Then the above equality implies that d(a, b) <
rb showing that a Ub , a contradiction. We arrive at UA UB = verifying property
T5 for (X, Td ).
The results of this example can be used to show that certain topologies are not
metrizable (Denition 2.4.17). It was already pointed out that the included point
topology Tp and the nite complements topology Tf c on R are not metrizable (see
paragraph following Denition 2.4.17). Example 4.2.4 shows that the excluded point
topology T p is not metrizable either.
Example 4.2.7. (Rn , TEu ) is T0 , . . . , T5 . Recall that the Euclidean topology TEu
on Rn is a metric topology associated to the Euclidean metric d2 . Thus Example 4.2.6
shows that (Rn , TEu ) is T0 , . . . , T5 .
4.3. Topological invariants
We have thus far expanded most of our eorts towards the study of properties
of topological spaces and continuous functions between them. These are necessary
prerequisites for the more involved concepts studied in later chapters and will be used
throughout the remainder of the book. Before delving into these, we pause here to
reect on a question that lies at the heart of most eorts in topology.
Question 4.3.1. Given a pair of topological spaces X and Y , are they homeomorphic?

66

4. SEPARATION AXIOMS

This is a dicult question in general, indeed its full resolution shall forever stay out
of grasp. Indeed, even restricting the choices of X and Y to particularly nice classes
of topological spaces, for example manifolds, this question is undecidable.
Nevertheless, we can try to obtain partial answers to Question 4.3.1. The main tool
in doing so is the use of topological invariants which we now introduce and then apply
to Question 4.3.1.
Definition 4.3.2. A topological invariant is a function
O : { Set of topological spaces } { Set of objects },
such that if X and Y are homeomorphic than O(X) = O(Y ). The Set of objects
can be any set, examples include the set of integers, the set (ring) of polynomials over
a commutative ring, the set of (isomorphism classes of ) groups, etc.
Remark 4.3.3. It is important to note that the implication
XY
=
O(X) = O(Y ),
=
is one directional. It is generally not possible to conclude that X and Y are homeomorphic given O(X) = O(Y ). With regards to Question 4.3.1, topological invariants
are most frequently used to tell that two spaces X and Y are not homeomorphic, as
would follow from O(X) = O(Y ), rather than to conclude that they are.
Many of the properties of topological spaces encountered in the present chapter and
in Chapter 2, are topological invariants. We present one example with full details and
state others with less elaboration.
Example 4.3.4. Consider the set
V={

Is separable. ,

Is not separable. }

Thus V is a set of two elements, each being a phrase. We use V to dene the function
O : { Set of topological spaces } V,
by setting
O(X) =

Is separable.

Is not separable.

If X is a separable topological space.

If X is not a separable topological space.

We claim that O is a topological invariant. To see this, suppose that X is a separable


space with A X being a countable dense subset of X. Let f : X Y be a
homeomorphism and let B = f (A) Y . Since f is a bijection, it preserves cardinality
of sets and so B is also countable. To see that B is also dense, let V Y be any open
set. Then U = f 1 (V ) is an open subset of X and so A U = according to Corollary
2.3.7. But then B V is also non-empty and so, again by Corollary 2.3.7, B must be
dense.
Similarly, if Y is separable then the same argument shows that X is also separable.
In conclusion, given two homeomorphic spaces X and Y , either both or neither is
separable and therefore O is a topological invariant.

4.4. A FIRST APPLICATION OF TOPOLOGICAL INVARIANTS

67

The preceding example of a topological invariant incorporated a greater than needed


number of details, in an eort to strictly adhere to Denition 4.3.2. Going forward,
we shall take a more relaxed approach. For instance, we may say that the property
of being separableis a topological invariant or, even shorter, that separabilityis a
topological invariant.
Example 4.3.5. Consider the two topological spaces: the Euclidean topology
(R, TEu ) and the excluded point topology (R, T p ) (Example 2.2.7). Then
O(R, TEu ) = Is separable.
O(R, T p ) = Is not separable.
The rst claim is courtesy of Example 2.3.8 and the second follows from Example
2.4.10. We conclude that (R, TEu ) is not homeomorphic to (R, T p ).
Theorem 4.3.6. The properties of a topological space of being
First Countable,
Second Countable,

Separable,
Being Ti for i {0, 1, 2, 3, 3 1 , 4, 5},
2

are all topological invariants.


Proof. The ten proofs of the topological invariance of the ten properties listed are
very similar (the proof for separability was already given in Example 4.3.4). We shall
thus only give here the proof of the topological invariance of the Hausdor property
T2 and defer there rest of the to the exercises (Exercise 4.6.5).
Let X be a Hausdor space and f : X Y be a homeomorphism. To show that
Y is also Hausdor, pick a pair of distinct points y1 , y2 Y and let xi = f 1 (yi ) for
i = 1, 2. Then x1 and x2 are distinct points (since f is a bijection) and by the Hausdor
property of X, there exist disjoint neighborhoods U1 and U2 of x1 and x2 respectively.
But then Vi = f (Ui ) for i = 1, 2 are disjoint neighborhoods of y1 and y2 , proving the
claim.
Example 4.3.7. Let TP be the partition topology on R associated to the partition
P = {[2a, 2a+2 | a Z}. Then the spaces (R, TP ) and (R, Tf c ) are not homeomorphic,
for the former is not T1 (Example 4.2.1) while the latter is (Example 4.2.2).
Example 4.3.8. The spaces (R, TEu ) and (R, T p ) are not homeomorphic as the
Euclidean line is a T3 space (Example 4.2.7) while (R, T p ) is not (Example 4.2.4).
4.4. A rst application of topological invariants
Before proceeding with our exposition, we pause and give an application of the
topological invariants introduced in Section 4.3 to address Question 4.3.1 for the various
example of topological spaces presented in Section 2.2. Thus, we shall consider the
topological spaces obtained by endowing R with one of the nine topologies below, and
ask which, if any, resulting spaces are homeomorphic.
1. TEu - The Euclidean topology from Example 2.2.1.

68

4. SEPARATION AXIOMS

2.
3.
4.
5.
6.
7.
8.
9.

Tdis - The discrete topology from Example 2.2.4.


Tp - The included point topology from Example 2.2.6.
T p - The excluded point topology from Example 2.2.7.
Tf c - The nite complement topology from Example 2.2.10.
Tcc - The countable complement topology from Example 2.2.11.
TF,p - The Fort topology from Example 2.2.12.
Tll - The lower limit topology from Example 2.2.14.
Tul - The upper limit topology from Example 2.2.14.

We organize our results in the table below. Given the row containing the topology
T1 and the column containing T2 , the corresponding entry in the table is either a
checkmark
or a crossmark , indicating whether (R, T1 ) and (R, T2 ) are or are not
homeomorphic. The table entries are clearly symmetric about the subdiagonal and the
subdiagonal itself consists of s as each space is homeomorphic to itself.
TEu Tdis
TEu

Tdis
Tp

p
T

Tf c

Tcc

TF,p

Tll

Tul

Table 1. Homeomorphism
troduced in Section 2.2.

T p T p Tf c







relationships

Tcc TF,p






among the

Tll Tul






topologies on R in-

We explain some of the table entries and leave some for the reader to work out.
This is an instructive exercise that we highly recommend (Exercise 4.6.8). The red
checkmarks linking the nite complement with the countable complement topology is
there to emphasize that all of the topological invariants introduced in Theorem 4.3.6
agree for (R, Tf c ) and (R, Tcc ). To tell that these two spaces are not homeomorphic,
another topological invariant is used, see Exercise 4.6.6. We start by considering the
second countability invariant.
Lemma 4.4.1. Of the nine topologies TEu , Tdis , Tp , T p , Tf c , Tcc , TF,p , Tll and Tul
on R, only the Euclidean topology TEu is second countable.
Proof. The Euclidean topology TEu is second countable according to Example
2.4.4, while the included point topology Tp is not second countable according to Example 2.4.5. Example 2.3.9 shows that (R, Tdis ) is not separable and so by Lemma 2.4.9
it is not second countable either. Here are the remaining six cases:

4.4. A FIRST APPLICATION OF TOPOLOGICAL INVARIANTS

69

T p Every set {x} R with x = p is open and cannot be written as the union of
any other non-empty open sets. Thus every basis B for (R, T p ) must at least
contain {x}, x R {p} which is an uncountable set. Consequently, every
basis is uncountable.
Tf c Suppose a countable basis B = {Ui | i N} existed. Then we could pick an
arbitrary point x R and consider the countable non-empty set Bx = {Ui
B | x Ui }. Since for every y = x, the set R {y} is open, it must be a union of
elements from B showing that for every y = x there exists an element Ui Bx
with y Ui . But then Ui Bx Ui = {x} so that
/
R {x} = R Ui Bx Ui = Ui Bx (R Ui ).

Tcc
TF,p
Tll

Tul

We have thus managed to write the uncountable set R {x} as a countable


union of the nite sets R Ui , a contradiction. Accordingly, (R, Tf c ) cannot be
second countable.
Since Tf c Tcc and Tf c is not second countable then neither is Tcc .
The Fort topology is ner than the excluded point topology which we already
saw was not second countable, showing that neither is TF ,p .
Suppose there were a countable basis B = {Ui | i N} for Tll . Since each
open set in the lower limit topology is a countable union of intervals [a, b
(Exercise 2.5.25), we can assume without loss of generality that Ui = [ai , bi
for some ai , bi R, ai < bi . Since R is uncountable we can nd a point
a R {a1 , a2 , a3 , . . . }. Then the set [a, a + 1 cannot be gotten as a union of
the basis elements. Thus (R, Tll ) is not second countable.
Same argument as for the lower limit topology Tll .

The results of the preceding lemma suce to ll out the rst row and column of
Table 1. We next turn to the Hausdor property.
Lemma 4.4.2. Of the nine topologies TEu , Tdis , Tp , T p , Tf c , Tcc , TF,p , Tll , Tul on R,
the Hausdor ones are precisely TEu , Tdis , TF,p , Tll and Tul .
Proof. Examples 4.2.2, 4.2.3 and 4.2.4 show that (R, Tf c ), (R, Tp ) and (R, T p )
are not Hausdor, while Examples 4.2.5 and 4.2.7 show that (R, Tll ) and (R, TEu ) are
Hausdor. A straightforward adaption of Example 4.2.5 shows that (R, Tul ) is also
Hausdor. Here are the remaining three cases.
Tdis This topology is Hausdor since {x} and {y} are disjoint neighborhood for a
pair of distinct points x, y R.
Tcc As in the nite complement topology, no disjoint non-empty open sets exist.
TF,p Let x, y R be two distinct points with x = p. Then {x} and R {x} are
two disjoint neighborhoods of x and y respectively, showing (R, TF,p ) to be
Hausdor.
Lemma 4.4.2 contributes an additional 32 crossmarks to Table 1.

70

4. SEPARATION AXIOMS

Lemma 4.4.3. The function f : R R given by f (t) = t is a homeomorphism


from (R, Tll ) to (R, Tul ).
Proof. This follows immediately from part (e) of Theorem 3.1.9 and the observation that f 1 ( a, b]) = [b, a and f ([c, d ) = d, c].
The preceding lemma accounts for the two o-diagonal checkmarks
The remaining 22 entries are deferred to Exercise 4.6.8.

in Table 1.

4.5. The Urysohn Lemma


The following lemma gives an alternative characterization of T4 spaces. A similar
characterization can also be given for T3 spaces, see Lemma 6.3.5.
Lemma 4.5.1. A topological space (X, TX ) is T4 if and only if for every closed subset
A X and any open set U X containing A, there exists an open set V X such
that

A V V U.
Proof. = Suppose rstly that X is T4 and let A U X be sets of which A
is closed and U open. Consider the two closed and disjoint sets A, X U X. By the
T4 property of X, there exist disjoint open subsets WA and WXU containing A and
X U respectively. But then
WA WXU =

WA X WXU ,

and since X WXU is a closed set, we see that W A X WXU . On the other
hand, since X U WXU then X WXU U . Thus V = WA has the property

that A V V U , as desired.
= Suppose that X has the property that for every pair A U X with A

closed and U open there is an open set V with A V V U and let A, B X


be two disjoint and closed subsets of X. Then U = X B contains A and so by

assumption there is an open set V with A V V X B. Then V and X V are


disjoint neighborhoods of A and B respectively, showing X to be T4 .
Theorem 4.5.2 (The Urysohn Lemma). Let (X, TX ) be a T4 space and let A, B
X be two closed and disjoint subsets of X. Then there exists a continuous function
f : X R such that f |A 0 and f |B 1. The codomain of f carries the Euclidean
topology.
Proof. Let Q[0,1] = Q [0, 1] be the set of rational numbers between 0 and 1. This
is a countable set and so we can list its elements as
Q[0,1] = {q0 , q1 , q2 , ...}
Any such enumeration of
choose q0 = 0 and q1 = 1.
0
1
1
Q[0,1] =
,
,
,
1
1
2

the elements of Q[0,1] will do but for convenience we shall


An example is
1 2
1 3
1 2 3 4
1 5
1 2 3 4 5 6
, ,
, ,
, , , ,
, ,
, , , , , ,... ,
3 3
4 4
5 5 5 5
6 6
7 7 7 7 7 7

4.5. THE URYSOHN LEMMA

71

in which, for instance, q5 = 1/4 and q9 = 3/5. We note that there is no relation
between the size of the indices i, j N {0} and the size of qi and qj .
In the rst step of the proof, our strategy shall be to construct an open set Uq for
every point q Q[0,1] , subject to the conditions

(a) A Uq0 and B X Uq1 .

(b) If x, y Q[0,1] with x < y then Ux Uy .


In the second step of the proof, these sets Uq are then used to construct the function
f : X R whose existence is claimed by the theorem.
Step 1 We shall dene the sets Uqi , qi Q[0,1] by explicitly dening Uq1 and Uq0
and then proceed by induction for i 2. Thus, dene Uq1 to be X B and note
that A Uq1 . According to Lemma 4.5.1, there is an open set Uq0 X such that

A Uq0 Uq0 Uq1 . With these denitions of Uq0 and Uq1 , properties (a) and (b)
above are met.
Proceeding by induction, suppose that the sets Uq0 , , ..., Uqn , subject to conditions
(a) and (b) above, have been dened. Consider the nite set {q0 , q1 , ..., qn , qn+1 } and
suppose that, with respect to the ordering of rational numbers, qn+1 ts in between qi
and qj with i, j {0, 1, ..., n}. Thus, the indices i, j are chosen so that qi < qn+1 < qj
and so that for any other index k {0, ..., n} {i, j} we either get qk < qi or qj < qk .
Such indices i, j must exist since q0 = 0 and q1 = 1 and qn+1 [0, 1]. By our inductive

assumption, the inclusion Uqi Uqj must hold and by Lemma 4.5.1, an open set

q Uqn+1 Uqn+1 Uq must exist. This last double inclusion shows


Uqn+1 X with U i
j
that properties (a) and (b) continue to hold for the sets Uq0 , ..., Uqn , Uqn+1 , completing
the induction.
Step 2 Using the sets Uqi , qi Q[0,1] dened in Step 1, we are now in the position
to dene a function f : X R as
f (x) =

1
inf{q Q[0,1] | x Uq }

;
;

x B,
x B.
/

Note that f is well dened since if x is not in B, then the set {q Q[0,1] | x Uq }
contains 0 and is therefore non-empty. We will prove that f satises the conditions
asserted by the theorem.
If x A then x U0 (and x B) so that inf{q Q[0,1] | x Uq } = 0 and hence
/
f |A 0. By denition of f , the relation f |B 1 is evident. To see that f is continuous
we will show that for every x X and every open set V R containing f (x) V , there
exists a neighborhood U of x in X with f (U ) V . This will show that f is continuous
at x and since x is arbitrary, we will have shown f to be continuous. Observe that
If x Uqi then f (x) qi .
If x Uqj then qj f (x).
/
The rst point above is obvious while the second point follows from the observation
that if f (x) = y < qj then x Uq for some rational number q in y, qj , an impossibility
since if x Uqj while Uq Uqj .
/

72

4. SEPARATION AXIOMS

U0
U2
3

U3
4

U1
Figure 1. A depiction of some of the sets Uq , q Q [0, 1] used in the
3
proof of the Urysohn lemma. The values of q = 2 , 4 have been picked
3
randomly, however, there is less freedom in choosing the sets U0 and U1
and they have been portrayed as chosen in the proof of the Urysohn
lemma.
Pick any x X and let V R be an open set with f (x) V . Find an > 0
so that f (x) , f (x) + V . If f (x) = 1, nd a rational number qj 1 , 1],

then U = X Uqj has the property f (U ) V . If f (x) = 0, nd a rational number


qi [0, and set U = Uqi , then f (U ) V . Finally, if 0 < f (x) < 1, nd rational

numbers qi , qj with f (x) < qj < x < qi < f (x) + and set U = Uqi Uqj . Then
again f (U ) V .
Before considering some of the consequences of Urysohns Lemma, well work through
an example.
Example 4.5.3. Consider the Euclidean line (R, TEu ) and let A, B R be the
closed subsets A = [2, 1] and B = [1, 2]. We shall follow, step by step, the proof of
Urysohns Lemma (Theorem 4.5.2) to construct a continuous function f : R R with

4.5. THE URYSOHN LEMMA

73

f |A 0 and f |B 1. The sets Uq with q [0, 1] Q can be chosen as


U1 = , 1 2,

Uq = , q ,

and

q [0, 1 Q.

Recall that the set U1 was dened as R B in the proof of Theorem 4.5.2, while the

sets Uq for q = 1 had to be chosen so that A U0 and whenever q1 < q2 then Uq1 Uq2 .
q1 = , q1 ] which is clearly contained in , q2 if
With our choices as above, U
q1 < q2 . The function f : R [0, 1] is then dened as
1
inf{q Q[0,1] | x Uq }

f (x) =

;
;

xB
xB
/

showing that f takes on the form

0
x
f (x) =
1

x0
0x1
; 1x

;
;

This is clearly a continuous and A f 1 (0) and B f 1 (1), see Figure 2 for a graph
of f .

Figure 2. The graph of the function f (in green) constructed using the
proof of the Urysohn Lemma for the subsets A = [2, 1] (in blue) and
B = [1, 2] (in red) of the Euclidean line.
Here are some immediate consequences of Urysohns Lemma.
Corollary 4.5.4. Every normal space is completely regular.
Proof. Suppose that X is normal, i.e. that it possesses the separation properties
T1 and T4 . To show that X is completely regular, we only need to demonstrate that X
is T3 1 . Let A X be any closed set and let x X A be any point. By Lemma 4.1.3,
2

74

4. SEPARATION AXIOMS

the set B = {x} is a closed set and so by the Urysohn Lemma (theorem 4.5.2), there
exists a map f : X R with f |A 0 and f (x) = 1. Thus X is T3 1 , as needed.
2

The next statement follows from the Urysohn Lemma and from Example 4.2.6 by
which every metric space (X, d) is T0 ,. . . ,T5 .
Corollary 4.5.5. Let (X, d) be a metric space and equip X with the associated
metric topology Td . Then (X, Td ) is T3 1 and completely regular. Thus, given any two
2
closed and disjoint subsets A, B X, there exists a continuous function f : X [0, 1]
with f |A 0 and f |B 1.
In the case of metric spaces, the function f from the preceding corollary can be
made completely explicit. To see this, we rst digress to make a denition and an
observation (Lemma 4.5.7).
Definition 4.5.6. Let (X, d) be a metric space and let A X be a subset of X.
We dene the function dA : X [0, as
dA (x) = inf{d(x, a) | a A},
and call it the distance function from A (or simply the distance from A).
Lemma 4.5.7. Let (X, d) be a metric space.
(a) If A X is closed subset, then d(x, A) = 0 if and only if x A.
(b) If A, B X are two disjoint and closed subsets of X, then dA (x) + dB (x) > 0
for every x X.
(c) For every subset A X, the function dA : X [0, is continuous.
Proof. (a) If x A then clearly dA (x) = 0. Suppose conversely that x X is
some element with dA (x) = 0. Thus, for every n N there must exist an element
an A with d(x, an ) < 1/n (by the denition of the inmum). But then lim an = x
and since A is closed, x must lie in A (part (a) of Theorem 3.2.9).
(b) This is immediate from part (a) for if dA (x)+dB (x) = 0 then dA (x) = 0 = dB (x).
By part (a) these two equalities would imply x A and x B and hence x A B,
a contradiction since A and B are assumed to be disjoint.
(c) Let x X be any point and let V [0, be any neighborhood of dA (x). If
dA (x) > 0, nd and > 0 such that dA (x) 2, dA (x) + 2 V and if dA (x) = 0
then nd an > 0 such that [0, 2 V . Let U = Bx () X, we claim that then
dA (U ) V showing continuity of dA at x. Since x X was arbitrary, this proves
continuity of dA .
To see that dA (U ) V , pick a point y U . For every n N there must exist a
point an A such that d(x, an ) < dA (x) + 1/n. By the triangle inequality, this implies
that d(y, an ) d(y, x) + d(x, an ) < + dA (x) + 1/n. Therefore, dA (y) dA (x) .
Arguing the same way by swapping the roles of x and y one obtains the inequality
dA (x) dA (y) showing that |dA (x) dA (y)| and so dA (y) V . Since y U
was arbitrary, this shows that dA (U ) V .

4.5. THE URYSOHN LEMMA

75

If the set A X is not closed, part (a) of the lemma may fail. For instance, if
(X, d) = (R, d2 ) (the Euclidean metric on R) and A = 0, 1 then dA (0) = 0 = dA (1)
but neither of 0 or 1 belong to A.
Armed with Lemma 4.5.7, we return to the setting of Corollary 4.5.5. Let (X, d)
be a metric space and let A, B X be a pair of disjoint, closed subsets of X. Dene
f : X [0, 1] as
dA (x)
(4.2)
f (x) =
.
dA (x) + dB (x)
By part (c) of Lemma 4.5.7, the denominator of the right hand side above, is never
zero showing that f is well dened. If x A then dA (x) = 0 so that f (x) = 0 while if
x B then dB (x) = 0 giving f (x) = 1 (by part (a) of Lemma 4.5.7). Thus A f 1 (0)
and B f 1 (1) showing that this function f has the properties of the function f from
Corollary 4.5.5. In fact, equality holds in both of these inclusions (Exercise 4.6.9) thus
yielding a slight strengthening of Corollary 4.5.5. With these ndings, we return one
more time to the setting of Example 4.5.3.
Example 4.5.8. Consider the Euclidean line (R, TEu ) with the Euclidean metric
d2 (x, y) = |x y| and let A, B R be the two closed and disjoint subsets A = [2, 1]
and B = [1, 2]. The two distance functions dA and dB take on the forms

; x 1,
; x 2,
1x
2 x
0
; x [1, 2],
0
; x [2, 1],
, dB (x) =
dA (x) =

x2
; x 2.
x+1
; x 1,
Combining these, the function f : R R from equation (4.2) becomes

2+x
; x 2
1+2x

0
; x [2, 1]

x+1
; x [1, 1]
f (x) =
2

; x [1, 2]

x+1

; x2
2x1
Figure 3 shows the graph of this function, the reader may nd it instructive to compare
it to the graph of the function f constructed in Example 4.5.3, see Figure 2.

76

4. SEPARATION AXIOMS

Figure 3. The graph of the function f (in green) from equation (4.2) for
the subsets A = [2, 1] (in blue) and B = [1, 2] (in red) of the Euclidean
line equipped with the standard metric d(x, y) = |x y|. Compare this
to the graph of the function f from gure 2 constructed directly from
the proof of Urysohns lemma.
4.6. Exercises
4.6.1. Finish the proof of of part (b) of Theorem 4.1.4: Show that a closed subspace
Y of a T5 -space X is itself a T5 -space.
4.6.2. Consider the real line R equipped with the Fort topology TF ,p . Show that
1
(R, TF ,p ) is a Ti -space for all i {0, 1, 2, 3, 3 2 , 4, 5}.
4.6.3. Consider R2 equipped with the topology T generated by the basis
B = {{x} U | x R, U TEu }.
Check that B is a basis of a topology, and then verify whether or not (R2 , T ) is
(a) Hausdor.
(b) Normal.
4.6.4. Consider the rationals Q R equipped with the relative Euclidean topology.
Show that Q is a T4 -space. Is it a T5 -space?
4.6.5. Prove the topological invariance of the properties of First and Second Countability, and of being a Ti -space for i {0, 1, 3, 3 1 4, 5}. Recall that the topological
,
invariance of Separability and of being a T2 -space were addressed in Example 4.3.4 and
Theorem 4.3.6.

4.6. EXERCISES

77

4.6.6. Consider the functions Omax , Omin : { Set of topological spaces } N{},
dened as

There exists a conver

max n N gent sequence in X


; There exists M N so

with n limits.
that no convergent seOmax (X, T ) =
quence in X has more

than M limits.

; Otherwise.

Omin (X, T ) =

There exists a conver


min n N

gent sequence in X

with n limits.

There is at least one


convergent sequence
in X with nitely
many limits.

Otherwise.

(a) Show that both Omax and Omin are topological invariants.
(b) Show that
Omax (R, Tf c ) = = Omin (R, Tf c )

and

Omax (R, Tcc ) = 1 = Omin (R, Tcc ).

Note that if (X, T ) is a Hausdor space, then Omax (X, T ) = Omin (X, T ) = 1. The
converse is not true in general.
4.6.7. Show that the function O : { Set of topological spaces } {0, 1}, dened
by
f (X, T ) =

0
1

;
;

There is no surjective map onto ({a, b}, Tdis ),


There is a surjective map onto ({a, b}, Tdis ),

is a topological invariant. Using this invariant, show that (R, TEu ) and (R {0}, TEu )
are not homeomorphic spaces.
4.6.8. Fill in the eleven pairs of missing entries in Table 1 that are not accounted
for by Lemmas 4.4.1, 4.4.2 and 4.4.3. Specically, show that of the nine topologies TEu ,
Tdis , Tp , T p , Tf c , Tcc , TF,p , Tll , Tul on R, the ones that are
(a) T1 are precisely TEu , Tdis , Tf c , Tcc , TF,p , Tll , Tul . Use this to add four pairs of
crossmarks to Table 1.
(b) T4 are precisely TEu , Tdis , T p , TF,p , Tll , Tul . Use this to conclude that (R, Tp )
and (R, T p ) are not homeomorphic, and add an new pair of crossmarks to
Table 1.
(c) separable are precisely TEu , Tp , Tll , Tul . Use this to add four pairs of crossmarks
to Table 1.

78

4. SEPARATION AXIOMS

(d) rst countable are precisely TEu , Tdis , Tp , T p , Tll , Tul . Conclude that (R, Tdis )
and (R, TF ,p ) are not homeomorphic, add a pair of crossmarks to Table 1.
(e) Use the results of Exercise 4.6.6 to show that (R, Tf c ) and (R, Tcc ) are not
homeomorphic, completing Table 1.
4.6.9. Let (X, d) be a metric space and consider X equipped with the associated
metric topology Td . Let A, B X be two closed and disjoint subsets so that the
function f : X [0, in equation (4.2) is well dened. Show that f 1 (0) = A and
f 1 (1) = B.
4.6.10. Consider R2 with the Euclidean metric and let A, B R2 be the sets
A = Cl(B(0,0) (1)) and B = Cl(B(4,0) (1)). Following the proof of the Urysohn Lemma
4.5.2, and the notation therein, show that possible choices of the sets Uq , q Q [0, 1],
are given by

; q=1
B(0,0) (2 + q)
Uq =
R2 B
; q = 1.
2
Determine the associated function f : R [0, 1] and check that is possesses the
properties claimed by the Urysohn Lemma.

CHAPTER 5

Product spaces

roduct spaces are to topological spaces what Cartesian products are to sets.
Indeed, given a family of topological spaces, we discuss how to naturally
introduce a product topology on the Cartisan product of the sets underlying
the given topological spaces.
Section 5.1 addresses the case of nite products, and studies basic concepts such
as, for example, continuity of functions into and from a product. We also take up
the question of how properties of the factors, such a separability or the Ti -properties,
transfer to the product space, and vice versa (Theorem 5.1.10).
Section 5.2 turns to innite products, where besides the already mentioned product
topology, there is a second natural choice of topology, referred to as the box topology.
The main goal of this section is to contrast the product and box topologies through
examples of sequences of functions.
5.1. Finite products
Definition 5.1.1. Let (X1 , T1 ),. . . ,(Xn , Tn ) be a collection of n 2 topological
spaces. Then their Cartesian product X1 ...Xn = {(x1 , . . . , xn ) | xi Xi , i = 1, ..., n}
becomes a topological space with the product topology TX1 ...Xn dened as the topology
generated by the basis
(5.1)

BX1 Xn = {U1 Un X1 Xn | Ui Ti , i = 1, . . . , n}.

The space (X1 ... Xn , TX1 ...Xn ) is called the product of (X1 , T1 ), . . . , (Xn , Tn ) or
simply a product space. We shall refer to the spaces (Xi , Ti ), i = 1, . . . , n as the factors
of (X1 Xn , TX1 Xn ). The functions
i : X 1 X n X i ,

i (x1 , . . . , xn ) = xi ,

shall be referred to as the projection maps or simply as the projections. Specically i


shall be called the i-th projection.
To see that the set BX1 Xn from Denition 5.1.1 forms the basis of a topology, we
invoke Lemma 2.4.2. Since Xi Ti for each i = 1, . . . , n, we see that X1 Xn
BX1 Xn so that the rst condition of Lemma 2.4.2 is met. As for the second, pick
arbitrary elements Ui , Vi Ti for each i = 1, . . . , n and let (x1 , . . . , xn ) (U1
Un ) (V1 Vn ) be any point. Since
(U1 Un ) (V1 Vn ) = (U1 V1 ) (Un Vn ),
79

80

5. PRODUCT SPACES

and because Ui Vi Ti , we see that (U1 Un ) (V1 Vn ) BX1 Xn


showing BX1 Xn to be the basis of a topology. Indeed, we have shown a bit more,
namely that the intersection of every two elements from BX1 Xn again belongs to
BX1 Xn .
For the most part, we shall be interested in Denition 5.1.1 with n = 2. This
causes no loss of generality since the n-fold product space (X1 ... Xn , TX1 ...Xn )
is homeomorphic to the space gotten by taking n 1 successive twofold products (in
arbitrary order) of the spaces (X1 , T1 ), . . . , (Xn , Tn ) (Exercise 5.3.1).
Remark 5.1.2. It may be tempting, but incorrect, to assume that every open set in
(X Y, TXY ) is a product of open sets from X and Y . Rather, according to Denition
5.1.1, every element W TXY is a union of such products:
Ui Vi

W =

with

Ui TX , Vi TY , i I,

iI

where I is some indexing set.


Example 5.1.3. Consider the topological spaces (X, TX ) = (R, Tp ) and (Y, TY ) =
(R, Tq ) where Tp and Tq are the included point topologies associated to the points
p, q R. An open set W X Y is a union of open set U V with U Tp and
V Tq . In particular, if U and V are non-empty, then p U and q V so that
(p, q) W .
The converse of this is false, i.e. a subset W of X Y that contains (p, q) may not
be open. An example is any set W = {(p, q), (x, y)} with x = p and y = q. Thus, the
product topology of Tp and Tq on R2 is not equal to the included point topology T(p,q)
on R2 .
The result of the following lemma is certainly expected though not completely
immediate from Denition 5.1.1.
Lemma 5.1.4. Let X and Y be two topological spaces and let A X and B Y
be two non-empty subsets.
(a) A B is an open subset of X Y with respect to the product topology TXY if
and only if A and B are both open.
(b) A B is a closed subset of X Y with respect to the product topology TXY if
and only if A and B are both closed.
Proof. (a) = If A and B are both open then AB is open simply by denition
of TXY .
= Suppose conversely that AB is open. Then we can write AB = iI Ui Vi
with Ui X and Vi Y each open. But then A = iI Ui and B = iI Vi showing
that they are both unions of open sets and thus open themselves.
(b) = Suppose that A X and B Y are closed subsets and note that
(X Y ) (A B) = X (Y B) (X A) Y.

5.1. FINITE PRODUCTS

81

Both of the sets on the right hand side of the above are open subsets of X Y showing
that (X Y ) (A B) is open and thus that A B is closed.
= Suppose that A B X Y is closed and let x X A be any point. Pick
y B arbitrarily and observe that then (x, y) lies in (X Y ) (A B) which is an
open set. Thus there must be neighborhoods Ux X of x and Vy Y of y such that
Ux Vy (X Y ) (A B). We claim that Ux is disjoint from A. If it were not,
we could pick a point z Ux A giving us the relation (z, y) (Ux Vy ) (A B),
a contradiction since (Ux Vy ) (A B) = . Therefore Ux A = showing that
X A = xXA Ux making X A an open and hence A a closed set. The proof that
B is closed is analogous.
Definition 5.1.5. Let fi : Xi Yi , i = 1, . . . , n be continuous functions and
dene the function
f1 fn : X1 Xn Y1 Yn
by f1 ... fn (x1 , . . . , xn ) = (f1 (x1 ), . . . , fn (xn )). We shall refer to f1 fn as
the product of the functions f1 , . . . , fn and shall call each fi a factor function.
Theorem 5.1.6. Let (X, TX ) and (Y, TY ) be two topological spaces.
(a) The projection maps X : X Y X and Y : X Y Y are continuous,
open and surjective functions.
(b) For any two points x0 X and y0 Y , the two inclusion maps y0 : X X Y
and x0 : Y X Y given by y0 (x) = (x, y0 ) and x0 (y) = (x0 , y), are
continuous functions.
(c) Let (Z, TZ ) be a topological space and let f : Z X Y be a function with
coordinate functions f = (fX , fY ). Then f is continuous if and only if each of
fX : Z X and fY : Z Y is continuous.
(d) A product of nitely many functions is continuous if and only if each of the
factors is continuous.
Proof. (a) Surjectivity of X and Y are obvious. To prove their continuity, let
U X and V Y be open sets. Then
1
X (U ) = U Y

and

1
Y (V ) = X V,

which are both open sets showing that X and Y are continuous.
Let W X Y be an open set and write W = iI Ui Vi with Ui TX and
Vi TY . Then
X (W ) = X (iI Ui Vi ) = iI Ui TX ,
showing that X is an open map. By the same token, so is Y .
(b) To show that y0 : X X Y is continuous we shall rely on part (e) of Theorem
3.1.9, by which it suces to show that 1 (W ) is open for W ranging through a basis
y0
B of TXY . Choosing B = BXY from Equation (5.1), and thus picking W = U V

82

5. PRODUCT SPACES

with U TX and V TY , we nd that


1 (U V ) =
y0

;
;

y0 V,
y0 V.
/

In either case 1 (U V ) TX , as needed. The case of x0 follows analogously.


y0
(c) = Suppose rst that f = (fX , fY ) : Z X Y is continuous. Then clearly
so are fX and fY since fX = X f and fY = Y f (by part (a) of the present theorem,
and part (b) of Proposition 3.1.11).
= Assume that fX : Z X and fX : Z Y are continuous and let U X and
V Y be arbitrary open sets. Relying again on part (e) of Theorem 3.1.9, continuity
1
1
of f follows by showing that f 1 (U V ) = fX (U ) fY (V ) is open. This in turns
follows plainly from the continuity of fX and fY and the denition of TXY .
(d) The coordinate functions of the product function f = f1 f2 : X1 X2 Y1 Y2 ,
are the two functions f1 X1 and f2 X2 . If f1 and f2 are continuous, then parts (a)
and (c) of the theorem imply that f is also continuous. Conversely, if f is continuous,
then so are f1 X1 and f2 X2 (by part (c) of the present theorem, with Z = X1 X2 ).
1
But then, for any open set U Y1 , the preimage f1 (U ) equals X1 ((f1 X1 )1 (U ))
which is open since f1 X1 is continuous and X1 is open (by part (a)). This implies
continuity of f1 , a similar argument gives continuity of f2 .
The next example demonstrates that projection maps X , Y dont need to be
closed maps.
Example 5.1.7. Consider the topological spaces (X, TX ) = (Y, TY ) = (R, TEu )
and let f R R be the graph of the function f : R R dened as f (x) =
2
sin(x)(1 ex ). We leave it as an easy exercise to show that f is a closed subset of
R R. On the other hand, Y (f ) = 1, 1 which is not a closed subset of (R, TEu ).
Thus, Y is not a closed map.
Inclusion maps on product spaces can fail to be open or closed maps.
Example 5.1.8. Let (X, TX ) and (Y, TY ) be two topological spaces and consider
the inclusion map y0 : X X Y (remember that y0 (x) = (x, y0 )). For any subset
A X we obtain y0 (A) = A {y0 }. If A is an open set then, according to Lemma
5.1.4, A {y0 } is an open set if and only if {y0 } is open. On the other hand, if A is
closed then, again using Lemma 5.1.4, the set A {y0 } is closed if and only if {y0 } is
closed.
Examples of spaces (Y, TY ) in which {y0 } fails to be both open and closed, abound.
An easy example is to take (Y, TY ) = (R, Tindis ), the indiscrete topology on R. For this
example, the inclusion map y0 is neither open nor closed for any choice of y0 R.
Corollary 5.1.9. Let X and Y be two topological spaces. Then, for any choice
of y0 Y , the inclusion map y0 : X X Y is a homeomorphism onto its image.
This allows us to view X as a subspace of X Y . If additionally Y is T1 , then X is a
closed subspace of X Y (for any choice of y0 Y ). Identical conclusions hold for Y .

5.1. FINITE PRODUCTS

83

Proof. Clearly y0 is injective and it is continuous by Theorem 5.1.6, part (b). Its
inverse function 1 = X |X{y0 } is also continuous since the restriction of a continuous
y0
function is continuous (Theorem 3.1.11). If Y is T1 then {y0 } is a closed subspace of
Y (Lemma 4.1.3) rendering X {y0 } a closed subspace of X Y by Lemma 5.1.4.
We next explore how properties of topological spaces, such as second countability
and the Hausdor property, transfer from factors to the product space and vice versa.
Theorem 5.1.10. Let X, Y be two topological spaces and let X Y be equipped with
the product topology. Let A and B be chosen from the sets of properties of topological
spaces

Separabiltiy,

Normality ( T1 and T4 ),
First countability,
.
, B
A
Complete Normality ( T1 and T5 ).
Second countability,

T ,T ,T ,T .
0

(a) X Y possesses property A if and only if both X and Y possess property A.


(b) If X Y possesses property B, then so do X and Y (the converse is generally
not true, see Example 5.1.12).
Proof. With the exception of separability, each A-property is inherited by a subspace (Theorems 2.4.18 and 4.1.4). As Corollary 5.1.9 allows us to view X and Y as
subspaces of X Y , it follows that if X Y has any of the A-properties (other than
separability), then so do X and Y .
Similarly, the B-properties are inherited by a closed subspace (Theorem 4.1.4).
Lemma 4.1.3 shows that a singleton in a T1 space is a closed set, from which it follows
(with the help of Corollary 5.1.9) that X and Y are closed subspaces of X Y . We
conclude that if X Y possesses any of the B-properties, then so do X and Y .
The remaining claims of the theorem are each addressed separately.
A is Separability
= Assume that X and Y are separable and let A X and B Y be countable

dense subsets. Then A B is a countable subset of X Y and A B = A B


(Exercise 5.3.2), showing that X Y is separable.
= Assume X Y is separable with C X Y a countable dense subset. Let
A = X (C) X and B = Y (C) Y . For any open set U X we must have
A U = for otherwise we would obtain C (U Y ) = . Since U Y is open in
X Y , this latter equality of sets produces a contradiction to the denseness property of
C according to Corollary 2.3.7. Thus A U = and so this same corollary guarantees
the denseness of A, and thus the second countability of X. The case of showing that

B = Y follows similarly.
A is Second Countability
Assume that X and Y are second countable and let BX = {Ui | i N} and BY =
{Vj | j N} be two countable bases for X and Y respectively. Dene B as B =
{Ui Vj | i, j N}. This is a countable set of open subsets of X Y . To see that it is

84

5. PRODUCT SPACES

a basis, let W X Y be any open set. By Denition 5.1.1 of the product topology,
we can nd open sets Ai X and Bi Y , i I, such that W = iI Ai Bi . For
each i I, let Ji and Ki be indexing sets such that Ai = jJi Uj and Bi = kKi Vk ,
then
W = iI Ai Bi = iI [(jJi Aj ) (kKi Bk )] = iI,jJi ,kKi Aj Bk .
This shows B to be a basis and thus X Y to be second countable.
A is First Countability
This proof mirrors the one just given for second countability and the details are left to
the reader (Exercise 5.3.3).
A is the T0 property, the T1 propertyor the T2 property.
We address the T0 T2 properties together since their proofs are very similar. Let
(x1 , y1 ) and (x2 , y2 ) be two distinct points in X Y .
T0 : If x1 = x2 , nd either a neighborhood U1 X of x1 that doesnt contain x2 or
else a neighborhood U2 X of x2 that doesnt contain x1 . Then either U1 Y
is a neighborhood of (x1 , y1 ) not containing (x2 , y2 ) or U2 Y is a neighborhood
of (x2 , y2 ) not containing (x1 , y1 ). If x1 = x2 then y1 = y2 and one proceeds in
a similar fashion.
T1 : If x1 = x2 , let U1 , U2 X be neighborhoods of x1 and x2 respectively with
x2 U1 and x1 U2 . Then U1 Y and U2 Y are neighborhoods of (x1 , y1 )
/
/
and (x2 , y2 ) respectively with (x2 , y2 ) U1 Y and (x1 , y1 ) U2 Y . The case
/
/
of y1 = y2 goes similarly.
T2 : If x1 = x2 , let U1 , U2 X be disjoint neighborhoods of x1 and x2 . Then U1 Y
and U2 Y are disjoint neighborhoods of (x1 , y1 ) and (x2 , y2 ) respectively. The
case of y1 = y2 goes accordingly.
A is the T3 property
Assume that X and Y are T3 . Let C X Y be a closed subset and pick a point
(x, y) X Y C. Consider X Y C as a neighborhood of (x, y) and nd
neighborhoods U of x and V of y such that U V X Y C. By regularity of X
we can separate x and the closed set X U by neighborhoods Wx and WXU . Note
that
Wx X WXU = W x X WXU U,
since X WXU is a closed set. In the same vein we nd neighborhoods Wy and
WY V of y and Y V with W y V . Then Wx Wy is a neighborhood of (x, y) with
W x W y U V X Y C so that X Y W x W y becomes a neighborhood
of C. Clearly Wx Wy and X Y W x W y are disjoint.
Corollary 5.1.11. The product X Y is a regular space ( T0 and T3 ) if and only
if each of the factors X, Y is regular.
The following example shows even when X and Y are normal spaces, X Y need
not be, demonstrating that part (b) of Theorem 5.1.10 cannot strengthened.

5.2. INFINITE PRODUCTS

85

Example 5.1.12. Consider the topological space (X, TX ) = (R, Tll ) where Tll is the
lower limit topology from Example 2.2.14 generated by the subbasis Sll = {[a, b | a, b
R, a < b}. It was already shown in Example 4.2.5 that X is T0 T5 and thus both
(completely) regular and (completely) normal. To see that X X is not T4 with
the product topology consider the subdiagonal = {(x, x) | x X}. Since a, b =
1
[a+ n , b is an open subset of X, it is easy to see that is a closed subset of X X.
n=1
The relative topology on is the discrete one because ([a, a+1 [a, a +1 ) =
{(a, a)}. Therefore, the two disjoint sets
AQ = {(a, a) | a Q}

and

AI = {(a, a) | a I = R Q},

are both closed in and thus also in X X.


If AQ and AI were separated by disjoint open sets UQ , UI X X, then for
every point (a, a) AQ there would have to a basis element [, [, with
(a, a) [, [, UQ , and similarly for points (b, b) AI . But the only basis
elements [, [, that contain (a, a) AQ and do not intersect AI , are those
of the form [a, [a, (with a < and a < ). Similarly, basis elements that
contain (b, b) AI and do not intersect AQ are of the form [b, [b, (with b <
and b < ). From this it follows that for every (a, a) AQ there exist a , a and for
every (b, b) AI there exist b , b , such that
UQ =

[a, a [a, a

and

aQ

[b, b [b, b .

UI =
bI

This relation however forces UQ UI = , a contradiction, showing that X X is not


a normal space.
Since X is regular, Corollary 5.1.11 implies that X X is also regular. We have
thus found an example of a regular space that isnt normal.
5.2. Innite products
In this section we briey turn to the innite product spaces. While in the case
of nite products, the product topology is a natural choice, in the case of innite
products there are two such natural choices. Depending on context, one may prove
more advantageous than the other (see for example Tychonos compactness theorem
in Section 9.4).
Let (X, TX ) and (Y, TY ) be two topological space. The product topology TXY on
X Y , generated by the basis {U V | U TX , V TY }, can be characterized in two
ways:
1. TXY is the smallest topology which contains all of the products of open sets
from X and Y .
2. TXY is the smallest topology on X Y for which the two projection maps
X : X Y X

and

Y : X Y Y,

are continuous. Indeed, if we assume their continuity, then for any U TX and
1
any V TY , the sets 1 (U ) = U Y and Y (V ) = X V would have to be

86

5. PRODUCT SPACES

open in X Y . But then their intersection (U Y ) (X V ) = U V is also


forced to be open.
A similar characterization holds for the product topology of nitely many factors. It
is of course easy to see that these two descriptions lead to the same topology. This,
however, is no longer the case when innitely many factors are involved.
Let I be an indexing set, possibly innite, and for each i I let (Xi , Ti ) be a
topological space. Let X = iI Xi denote their Cartesian product, that is let X be
X = {x : I iI Xi | x(i) Xi }.
As customary, we shall denote x(i) simply by xi . For example, if Xi = Y for all i I,
then X is simply the set of functions from I to Y .
Definition 5.2.1. Let (Xi , Ti ), i I be a collection of topological space and let
X = iI Xi .
(a) The box topology TBox on X is the topology generated by the basis
BBox = {iI Ui | Ui Ti }.
(b) The product topology TP rod on X is the topology generated by the basis
BP rod = {iI Ui | Ui Ti and Ui = Xi for all but nitely many i I}.
The verication that the sets BBox and BP rod are indeed bases of a topology, is easy
and is left as an exercise (Exercise 5.3.4). It should be noted that TBox is strictly ner
than TP rod whenever I is innite. When I is nite, the two topologies agree.
Example 5.2.2. Choose I = N and let (Xi , Ti ) = (R, TEu ). To tell copies of R
apart we shall label them as Ri , i I. Denote the product iN R by RN and note
that RN is simply the set of sequences in R. We shall consider the subset Y X of
bounded sequences in R. On Y , consider the metric d : Y Y [0, (Exercise
5.3.5) given by

d(x, y) =
i=1

|xi yi |
.
2i

The associated metric topology is the relative topology on Y induced by Tbox . On the
other hand, the relative topology on Y induced by Tprod is not a metric topology since
it is not Hausdor. For example, there are no disjoint neighborhoods of xi = 1/i and
yi = 1/i in Tprod .
Example 5.2.3. Pick I = R and let (X , T ) equal the Euclidean line for each
I. To distinguish the copy of R corresponding to I, we shall denote it by R .
Well denote the Cartesian product R R by RR . As in the previous example, we
can interpret RR as a set of functions, namely functions from R to R. We will examine
when a sequence of functions fi RR converges to a function f RR with respect to
the two topologies Tprod and Tbox .

5.2. INFINITE PRODUCTS

87

Consider rst RR with the product topology, let {fk }k RR be a sequence with
limit f . Let R be arbitrary and let U R be a neighborhood of f (). Consider
the open set in the product topology
U = (< R ) U (> R ).
Then U is a neighborhood of f and so there must exist an integer k such that fk U
for all k k . This last statement is equivalent to saying that fk () U for all k k
which is simply saying that the sequence fk () converges to f () in the Euclidean
topology on R.
Conversely, suppose that for each R, the equation lim fk () = f () holds with
respect to the Euclidean topology on R. Pick any neighborhood of V of f and suppose
that
V = R{1 ,...,n } R (V1 ... Vn ) ,
where 1 , ..., n R are chosen arbitrarily and where V1 , ..., Vn are arbitrary neighborhoods of f (1 ), ..., f (n ) respectively. Then there exists an integer k0 such that
k k0 implies that fk (j ) Vj for all k k0 and for any choice j {1, ..., n}. This
of course is equivalent to saying that fk V showing that lim fk = f with respect to
the product topology. In conclusion, we have shown that convergence of a sequence
of functions in RR with respect to the product topology, is equivalent to pointwise
convergence with respect to the Euclidean topology
lim fk = f in (RR , Tprod )

lim fk () = f () in (R, TEu ) for every R.

Next consider RR equipped with the box topology and suppose again that {fk }k
RR is a sequence with limit f . For each R, let U be a neighborhood of f () with
respect to the Euclidean topology on R and let U be the neighborhood of f (with
respect to the box topology) dened as
U = R U .
By assumption, there exists an integer k0 such that k k0 implies that fk U.
The latter inclusion is equivalent to fk () U for every R. If we choose U =
f () , f () + for each R and for some xed > 0, then we nd that
|fk () f ()| < for all k k0 , and for all values of R. The latter says that the
sequence of functions fk : R R converges uniformly to the function f : R R (with
respect to the Euclidean topology on R). Thus

(5.2)

lim fk = f in (RR , Tbox )

lim fk = f uniformly as functions (R, TEu ) (R, TEu ).

For instance, dene fk : R R as fk (t) = etk . Then, for every xed but arbitrary
t R, the relation lim fk (t) = 0 holds in (R, TEu ) so that lim fk = 0 in (RR , Tprod ).
However, the convergence lim fk (t) = 0 in (R, TEu ) is not uniform (since for every xed

88

5. PRODUCT SPACES

k, the quantity |fk (t)| can be arbitrarily large as we vary t) so that lim fk = 0 in
(RR , Tbox ).
The implication in (5.2) is in one direction only, as evidenced by the next example,
which exhibits a sequence of functions fk : R R that converges uniformly to a
function f : R R, but such that the sequence {fk }k does not converge to f in
(RR , TBox ).
2

Example 5.2.4. Consider the sequence of functions fk : R R, fk () = e( +k) ,


k N. Then fk uniformly converges to f 0 because for every > 0 and any k0 with
ek0 < , we obtain for each k k0 and for every R, the inequality
|fk () f ()| = e(

2 +k)

ek ek0 < .

On the other hand, {fk }k does not converge to f in (RR , TBox ) for if it did, then for
4
the neighborhood U = R B0 (e ) of f , we should be able to nd an integer k0
with k k0 implying fk U. The latter inclusion is tantamount to the inclusions
4
fk () B0 (e ) for every k k0 and every R. This in turn amounts to having
2
4
the inequality e( +k) < e hold for all k k0 and all R, clearly an impossibility.
5.3. Exercises
5.3.1. Prove the following two claims about nite product spaces.
(a) Given topological spaces (X1 , T1 ), . . . , (Xn , Tn ) show that
(X1 . . . Xn , TX X ) (X1 . . . Xn1 , TX X ) (Xn , Tn ).
=
1

n1

Said dierently, an n-fold product space can be constructed by taking the product of two factors at a time.
(b) For topological spaces (X, TX ) and (Y, TY ), show that
(X Y, TXY ) (Y X, TY X ).
=
Thus, up to homeomorphism, the order of the factors in a product spaced does
not matter.
5.3.2. Let (X, TX ) and (Y, TY ) be two topological spaces, and let A X and B Y
be a pair of given subsets. Viewing A B as a subset of (X Y, TXY ), show that
(a) A B = A B.

(b) (A B) = A B.

(c) (A B) = (A B) (A B).
5.3.3. Show that if X and Y are rst countable spaces, then X Y , equipped with
the product topology, is also a rst countable space.
5.3.4. Prove that BBox and BP rod from Denition 5.2.1, are each a basis of some
topopology.
5.3.5. Fill in the missing details from Example 5.2.2. Namely, let Y RN be the
subset of bounded sequences. Show that

5.3. EXERCISES

89

y
(a) The function d(x, y) = iN |xi2i i | denes a metric on Y .
(b) The associated metric topology Td on Y agrees with the relative box topology
TBox inherited from RN .
(c) Show that the relative product topology on Y is not Hausdor, and thus not
metrizable.

5.3.6. Let (X, TX ) be a topological space and consider X X equipped with the
product topology TXX .
(a) Show that the function : X X X given by (x) = (x, x) is continuous.
(b) Show that X is Hausdor if and only if the image of is a closed subset of
X X.
5.3.7. Consider [0, 1 and [0, 1] each equipped wit the relative Euclidean topology.
Show that with respect to the product topologies, the space [0, 1 [0, 1] is homeomorphic to [0, 1 [0, 1 . Does the same claim hold for the spaces [0, 1] [0, 1] and
[0, 1 [0, 1 ?
5.3.8. For any m, n N, show that (Rn , TEu ) (Rm , TEu ) (Rn+m , TEu ).
=

CHAPTER 6

Compactness

compactness of a space is one of the most fundamental and important properties associated to a topological space, and has far reaching consequences
both for subspaces and maps on the compact space. As the name suggests,
a compact space can be thought of as being small. In that sense, compactness is not unlike the notions of separability and second countability introduced in
Chapter 2, but its impact on properties of the underlying space far surpasses those of
the latter two.
Section 6.1 introduces compactness, provides examples and establishes rst properties of compact spaces. Section 6.2 heeds special attention to compactness on metric
spaces, culminating with the explicit characterization of compactness of Theorem 6.2.10
(see Corollary 6.2.11 for compactness in Euclidean spaces). Section 6.3 studies more
in depth properties of compact spaces, some of which rely on results from Section 6.2.
Section 6.4 studies the One Point Compactication, a procedure for exhibiting any
non-compact space X as a subspace of a compact space obtained by adding just a
single point to X. Finally, Section 6.5 touches on other avors of compactness, namely
Local Compactness, -compactness and Paracompactness.
6.1. Denition and rst examples
Definition 6.1.1. Let (X, TX ) be a topological space.
(a) An open cover F of X is a collection of open subsets of X whose union is all of
X.
(b) A subcover F of F is a subset of F which is still an open cover of X. A proper
subcover of F is any subcover of F not equal to F.
(c) A renement F of F is an open cover of X with the property that for every
U F there exists a V F such that U V .
Definition 6.1.2. A topological space (X, TX ) of compact if every open cover F
of X has a nite subcover.
Example 6.1.3. Any topological space (X, TX ) with TX a nite set, is compact. In
particular, if X is a nite set then (X, TX ) is compact for any choice of a topology TX .
Example 6.1.4. The Euclidean line (R, TEu ) is not compact. This can be seen by
considering the open cover
F = { a 1, a + 1 | a Z}
91

92

6. COMPACTNESS

There are no proper subcovers of F at all, let alone nite subcovers, since each integer
a Z lies in precisely one element of F, namely a 1, a + 1 . A similar argument can
be used to show that the n-dimensional Euclidean space (Rn , TEu ) is also not compact
(Exercise 6.6.1).
Example 6.1.5. Consider the nite complement topology Tf c on R and let F be
an open cover. Pick any U F with U = . Then U = R {x1 , ..., xn } for some points
xi R, i = 1, ..., n. For each xi let Vi F be such that xi Vi (such a Vi has to exist
since F is an open cover). But then {U, V1 , ..., Vn } is a nite subcover of F showing
that (R, Tf c ) is a compact space.
Example 6.1.6. Consider the included point topology Tp on R. The innite open
cover
F = {{p, x} | x R {p} }
has no proper subcovers. Consequently (R, Tp ) is not compact.
Example 6.1.7. Consider the excluded point topology T p on R and let F be any
of its open covers. Since R is the only non-empty open set that contains p, R must be a
member of every open cover. Given this observation, the cover {R} is a nite subcover
of any open cover. Therefore, (R, T p ) is compact.
We turn to some simple properties of compact spaces.
Theorem 6.1.8. Let f : X Y be a map. If X is compact then so is f (X). In
particular, compactness is a topological invariant.
Proof. Let Ff (X) = {Vi | i I} be an open cover of f (X) and consider the open
cover FX = {f 1 (Vi ) | i I} of X. By compactness of X, this latter cover has a
nite subcover FX = {f 1 (Vi1 ), ..., f 1 (Vin )} for some indices i1 , ..., in I. But then
Ff (X) = {Vi1 , ..., Vin } is a nite subcover of Ff (X) .
If f : X Y is a homeomorphism and X is compact, then so if Y , being equal
to f (X). If X is not compact then neither is Y since compactness of Y would imply
that of X, being equal to f 1 (Y ). Thus, two homeomorphic spaces X and Y are either
both compact or both non-compact.
Theorem 6.1.9. Let X and Y be two topological spaces.
(a) If X is compact and A is a closed subspace of X then A is also compact.
(b) If X is Hausdor and A X is a compact subspace then A is closed in X.
Proof. (a) Let FA = {Vi | i I} be an open cover of A and let Ui be open subsets
of X such that Vi = A Ui . Then FX = {X A, Ui | i I} is an open cover of X
and thus has a nite subcover FX that looks like FX = {Ui1 , ..., Uin , X A} or FX =
{Ui1 , ..., Uin } for some choice of indices i1 , ..., in I. In either case, FA = {Vi1 , ..., Vin }
is a nite subcover of FA showing that A is compact.
(b) Fix a point a A. The Hausdor condition guarantees that for every point
x X A there exist disjoint neighborhoods Ua,x and Vx,a of a and x respectively. The

6.1. DEFINITION AND FIRST EXAMPLES

93

collection {Ua,x A | a A} is an open cover of A and so by compactness, it has a nite


subcover {Ua1 ,x , ..., Uan ,x } for some points a1 , ..., an A. Let Vx = n Vx,ai . Clearly
i=1
Vx is a neighborhood of x but moreover, Vx and A are disjoint. For if not, we could
nd a point y Vx A. Then we would have y Uaj ,x for some j {1, ..., n}. But
on the other hand y Vx Vx,aj which is impossible since Uaj ,x Vx,aj = . Finally,
we note that X A = xXA Vx showing that X A is open and therefore that A is
closed.
Corollary 6.1.10. Let X be a compact space and Y a Hausdor space and let
f : X Y be a continuous function. Then f is a closed map. In particular, if f is a
bijection then it is a homeomorphism.
Proof. Let A X be any closed subset of X. According to part (a) of Theorem
6.1.9, A is compact. According to part (b) of Theorem 6.1.9, f (A) is compact, which
renders it a closed subset of Y by part (c) of Theorem 6.1.9. Thus f is a closed map.
If f is a bijection, it suces to show that f 1 : Y X is continuous in order to
demonstrate that f is a homeomorphism. According to part (a) of Theorem 3.1.9, the
latter follows from the closedness property of f which has already been established.
Theorem 6.1.11. Let (X, TX ) and (Y, TY ) be two topological spaces and let X Y
be equipped with the product topology TXY . Then X Y is compact if and only if each
of X and Y are compact.
Proof. = Suppose rstly that X and Y are both compact and let W = {Wi | i
I} be an open cover of X Y . For each Wi W, nd open sets Ui,j X and
Vi,j Y , j Ji such that Wi = jJi Ui,j Vi,j . If we can show that the open cover
W = {Ui,j Vi,j | j Ji , i I} has a nite subcover, then clearly so does W. Thus,
without loss of generality, we shall assume that each Wi has the form Ui Vi , i I to
begin with.
Pick a point y Y and consider the subspace X {y} X Y . As we saw in
Corollary 5.1.9, X {y} is homeomorphic to X and so X {y} must also be compact
(Proposition 6.1.8). Therefore, the induced open cover
Wy = {(Ui Vi ) (X {y}) | i I},
of X {y}, must have a nite subcover Wy = {Uj {y} | j Jy } for some nite subset
of indices Jy I. Let Vy = jJy Vj and note that Vy is a neighborhood of y Y with
the property that Uj Vy Uj Vj for all j Jy .
Perform the above procedure for all y Y to arrive at the open cover {Vy | y Y }
of Y . By compactness of Y , this has a nite subcover {Vy1 , ..., Vyn } for some points
y1 , ..., yn Y . We claim that then the set
W = {Uj Vj | j Jyi , i = 1, ..., n}
is a nite subcover of W . Given any point (x, y) X Y , there exists an index
i {1, ..., n} such that y Vyi . Since the sets Uj , j Jyi cover X, we nd that x Uj
for some j Jyi . But then (x, y) Uj Vyi Uj Vj W . So W is an open cover.

94

6. COMPACTNESS

= Since the projections maps X : X Y X and Y : X Y Y are both


continuous and surjective (part (a) of Theorem 5.1.6), and since the continuous image
of a compact space is compact (part (b) of Theorem 6.1.9), the compactness of X Y
immediately implies that of both X and Y .
By induction, the result of the preceding theorem extends without diculty to
arbitrary nite products to show that X = X1 ... Xn is compact (with respect
to the product topology) if and only if each of the factors Xi is compact. The case
of innite products is more subtle. On an innite product Y = iI Yi there are
two natural choices of topologies available: the product topology TP rod and the box
topology TBox (Denition 5.2.1). Regardless of the choice, the compactness of Y still
implies the compactness of the factors Yi (by the same argument as in the second half
of the proof of Theorem 6.1.11, which only relied on the surjectivity of the projection
maps i : Y Yi ). Conversely however, the compactness of Y is only implied by the
compactness of the Yi when Y is given the product topology, but not with respect to
the box topology. See Section 9.4 for more details on this.
Before proceeding to study more in depth properties of compact spaces in Section
6.3, we digress rst to take a closer look at compactness of metric spaces.
6.2. Compactness for metric spaces
Compactness for metric spaces is much easier to characterize than in the case of
general topological spaces (Theorems 6.2.6 and 6.2.10). An important special case of
these ndings is a completely explicit descriptions of compact subspaces of Euclidean
space (Rn , TEu ) (Corollary 6.2.11).
Definition 6.2.1. Let (X, d) be a metric space.
(a) (X, d) is called totally bounded if for every > 0 there exist nitely many points
xi X, i = 1, . . . , n such that X = n Bxi ().
i=1
(b) (X, d) is said to be bounded if there exists some r > 0 such that d(x, y) < r for
all x, y X, otherwise we call (X, d) unbounded.
(c) If (X, d) is bounded, we dene its diameter diam(X) as
diam(X) = sup{d(x, y) | x, y X}.
If (X, d) is unbounded we dene diam(X) to be innite.
Using the triangle inequality (relation (3) from the axioms of a metric, Example
2.2.3), it is easy to show that a a totally bounded metric space is always bounded. The
converse is not generally true (Exercise 6.6.2).
For the remainder of this chapter we shall always assume, without further mention,
that a metric space (X, d) has been made into a topological space with the metric
topology Td (Example 2.2.3).
Lemma 6.2.2. A totally bounded metric space (X, d) is separable.

6.2. COMPACTNESS FOR METRIC SPACES

95

1
Proof. For every n N and for the choice of n = n , there is a nite set of points
1
An = {xn , . . . , xnn } X such that X = xAn Bx n . Note that A = nN An is a
1
countable subset of X. We claim that A is also dense. To see this, let U X be any
non-empty open set and let x U be any element. Then there exists an > 0 such
1
1
that Bx () U . Find an integer n large enough so that n < and let x Bxn ( n ) for
i
1
n
n
some i {1, . . . , n }. Then d(x, xi ) < n < showing that xi Bx () U and thus

A U = . Since U was arbitrary it follows that A = X (Corollary 2.3.7).

Remark 6.2.3. Recall from Proposition 2.4.15 that a metric space is second countable if and only if it is separable. In particular, a totally bounded metric space must
be second countable.
Lemma 6.2.4. Let (X, d) be a metric space with the property that every sequence
{xk }k X has a convergent subsequence. Then (X, d) is totally bounded.
Proof. Suppose X were not totally bounded. Then there would be some > 0
so that X cannot be covered by a nite number of balls of radius . Using this,
we can construct a sequence {xk }k X as follows: let x1 be arbitrary and choose
consecutive elements of the sequence so that xk+1 k Bxj (). Notice that it follows
/ j=1
that xm j=1 Bxj () whenever < m.
/
Let yn = xkn be some convergent subsequence of xk with lim yn = y X. Then
there must exist an n0 N such that n n0 implies yn By (/2). Let , m be integers
such that n0 < m, then
d(xk , xkm ) d(xk , y) + d(y, xkm ) = d(y , y) + d(y, ym ) < /2 + /2 = .
This implies that xkm Bxk () which is a contradiction since k < km and by denition,
/
xkm kk Bxk (). Thus X must be totally bounded.
Lemma 6.2.5 (Lindelfs Theorem). If X is a second countable topological space
o
then every open cover F of X has a countable subcover.
Proof. Let B = {Un | n N} be countable basis for X and let F = {Vi | i I}
be an open cover of X. Since each Vi is an open set, there must exist a subset Ji N
such that Vi = jJi Uj . Dene F = {Uj B | j Ji for some i I}. Clearly F is
countable and it is also easy to see that F is an open cover of X. For if x X is an
arbitrary point then there is some i I with x Vi and thus x Uj for some j Ji .
On the other hand, F is a renement of F by denition. For each Uj F , let j I
be such that Uj Vj . Then F = {Vj | Uj F } is a countable subcover of F.
Theorem 6.2.6. A metric space (X, d) is compact if and only if every sequence
{xk }k X has a convergent subsequence.
Proof. = Let X be a compact space and let {xk }k X be an arbitrary
sequence. Suppose that there exists a point y X such that every neighborhood of
y contains innitely many distinct points of the sequence {xk }k . For such a point y
we dene a subsequence yn = xkn of {xk }k as follows: Pick y1 to be any point in

96

6. COMPACTNESS

By (1) {xk | k N}. With y1 , . . . , yn1 chosen, pick yn to be an arbitrary point in the
1
set By ( n ) {xk | k > kn1 }. It is easy to see that {yn }n is a convergent subsequence of
{xk }k with limit y, proving the theorem.
If this doesnt occur, then every y Y possesses a neighborhood Uy that contains
only nitely many distinct points of the sequence {xk }k . Then F = {Uy | y X}
is an open cover of X and so by compactness of X, it has a nite subcover F =
{Uy1 , . . . , Uym }| for some points y1 , . . . , ym X. But then the innite sequence {xk }k
is contained in the nite union Uy1 ... Uy1 showing (by the Pigeon Hole Principle)
that some point y must equal xk for innitely many distinct values k1 , k2 , k3 , . . . of
k. Then the sequence {xk1 , xk2 , xk3 , . . . }, being a constant sequence, is a convergent
subsequence of {xk }k .
= Suppose that every sequence {xk }k X has a convergent subsequence and
let F be an arbitrary open cover of X. Lemma 6.2.4 shows that X is totally bounded,
and hence by Lemma 6.2.2 it is also separable. By virtue of Proposition 2.4.15 (see
Remark 6.2.3) and using Lemma 6.2.5, we see that there must be a countable subcover
G = {U1 , U2 , U3 , . . . } of F.
We will prove the theorem by showing that G has a nite subcover. Assume to
the contrary that G has no nite subcovers. Form a sequence {xk }k X by choosing
k1
x1 X arbitrarily and by picking xk from the set X i=1 Ui for k 2. By the
assumption, there must be a subsequence yn = xkn with converges to some point y X.
Since G is an open cover of X, there must be some index m N with y Um . But
then, by construction, yn Um for all n > m, a contradiction given that lim yn = y.
/
Therefore G, and hence also F, must have a nite subcover. Since the cover F was
arbitrary, it follows that X is compact.
Corollary 6.2.7. Every compact metric space X is totally bounded, separable and
second countable.
Proof. This follows from Theorem 6.2.6 and Lemmas 6.2.2 and 6.2.4 and Remark
6.2.3.
While Theorem 6.2.6 provides a necessary and sucient condition for a metric
space (X, d) to be compact, it is often tedious in practice to check that X has the
property that every of its sequences has a convergent subsequence. We provide a
second characterization of compactness on metric spaces, one which involved easier to
verify conditions. To state that theorem, we rst introduce two additional concepts.
Recall that a sequence {xk }k Rn is said to be a Cauchy sequence (with respect
to the Euclidean topology) if for every > 0 there exists a natural number n0 such
that for all m, n n0 the inequality d2 (xn , xm ) < holds. This denition involves only
the Euclidean metric d2 on Rn and none of the other more advances structures on Rn .
Hence, it easily transfers to any metric space.
Definition 6.2.8. Let (X, d) be a metric space. A sequence {xk }k X is said to
be a Cauchy sequence if for every > 0 there exists a natural number n0 such that for

6.2. COMPACTNESS FOR METRIC SPACES

97

all m, n n0 we obtain d(xn , xm ) < . The metric space (X, d) is said to be a complete
metric space (or simply complete) if every Cauchy sequence is convergent.
Example 6.2.9. Consider Rn equipped with the Euclidean topology associated to
the Euclidean metric d2 . For a subset A Rn , the metric space (A, d2 |AA ) is complete
if and only if A is closed. In particular, (Rn , d2 ) itself is complete.
To see this, suppose that A is closed in Rn and let {xk }k A be a Cauchy sequence.
Every Cauchy sequence is bounded and every bounded sequence in Rn has a convergent
subsequence yn = xkn (these are standard results in analysis whose proofs we omit)
with limit lim yn = y. By Exercise 6.6.3 the sequence {xk }k is then also convergent
with lim xk = y. Since A is closed it must contain the limits of all its convergent
sequences, and thus y A and A is complete.
Conversely, suppose that A is complete. If A werent closed then Rn A wouldnt
be open and we could nd a point y Rn A all of whose neighborhoods intersect A.
1
We could then dene a sequence {xk }k A by choosing xk arbitrarily from By ( k ) A.
1
As d2 (xk , x ) < k + 1 , the sequence {xk }k is Cauchy, creating a contradiction to the
assumption that A is complete and the fact that lim xk = y A. Thus A must be
/
closed.
The following theorem is the main result of this section.
Theorem 6.2.10. A metric space (X, d) is compact if and only if is complete and
totally bounded.
Proof. = Suppose that X is compact. To see that X is complete, let {xk }k
X be any Cauchy sequence. By Theorem 6.2.6 there is a convergent subsequence
yn = xkn of {xk }k with limit y. But then xk itself has to converge to y (Exercise 6.6.3)
showing that X is complete.
To see that X is totally bounded, pick an arbitrary > 0 and consider the open
cover F = {Bx () | x X} of X. By compactness of X, F has a nite subcover,
showing that X is totally bounded.
= Assume that X is complete and totally bounded. We will show that X is
compact by relying on Theorem 6.2.6. Specically, we will show that every sequence
{xk }k X has a convergent subsequence.
Since X is totally bounded, there is a nite cover of X with balls of radius 1, say
F1 = {Bx1,1 (1), ..., Bxn1 ,1 (1)}. Since the sequence {xk }k has innitely many terms,
one of the balls from F1 must contain innitely many elements of the said sequence.
Without loss of generality, assume that Bx1,1 (1) is such a ball. Using total boundedness
1
of X once more, we nd that there is a nite cover F2 = {Bx1,2 ( 1 ), ..., Bxn2 ,2 ( 2 )} of X
2
by balls of radius 1 . Since Bx1,1 (1) contains innitely many elements of the sequence
2
{xk }k , there must be a ball in F2 whose intersection with Bx1,1 (1) contains innitely
many elements of the sequence {xk }k . Again, and without loss of generality, we assume
1
that Bx1,2 ( 2 ) is that ball so that Bx1,1 (1) Bx1,2 ( 1 ) contains innitely many elements
2
1
of the sequence xk . Proceeding inductively, we obtain a family of balls Bx1,k ( k ) of

98

6. COMPACTNESS

shrinking radii such that for any n N the set n Bx1,j ( 1 ) contains innitely many
j=1
j
elements of the sequence {xk }k .
To nd a convergent subsequence of {xk }k , pick indices k1 < k2 < k3 < . . . such that
xkn n Bx1,j ( 1 ) and set yn = xkn . Then yn is a Cauchy sequence since d(yn , ym ) <
j=1
j
1 1
max{ n , m }. Since X is assumed to be complete, yn must be a convergent sequence and
so we have generated a convergent subsequence of {xk }k , as needed. This completes
the proof.
Corollary 6.2.11. Let A be a subspace of n-dimensional Euclidean space Rn .
Then A is compact if and only if it is closed and bounded.
Proof. If A is bounded then it is also totally bounded (Exercise 6.6.2) and if in
addition it is also closed, then it is complete by Example 6.2.9. Theorem 6.2.10 then
shows that A is compact.
Conversely, if A Rn is compact, then by Theorem 6.2.10 is must be complete and
hence closed by Example 6.2.9. Theorem 6.2.10 also shows that A has to be totally
bounded and therefore bounded (Exercise 6.6.2).
Example 6.2.12. The n-dimensional sphere S n is the subspace of Rn+1 consisting
of vectors of norm 1:
S n = {(x1 , ..., xn+1 ) Rn+1 | x2 + ... + x2 = 1}.
1
n+1
It is quite obviously bounded and it is also closed. An easy way to verify closedness
of S n is to consider the continuous function f : Rn+1 R given by f (x1 , ..., xn+1 ) =
x2 + ... + x2 . Then S n = f 1 (1) and {1} R is a closed set. According to Corollary
1
n+1
6.2.11, S n is compact.
6.3. Properties of compact spaces
In this section we return to the investigation of properties of compact spaces already
initiated in Section 6.1. Some of the properties established in this section rely on the
results about metric spaces obtained in the preceding section.
Theorem 6.3.1. Let X be a compact topological space and let f : X R be
a continuous function to the Euclidean line. Then f attains both its maximum and
minimum value.
Proof. By part (b) of Theorem 6.1.9, the image f (X) is a compact subspace of
R. By Corollary 6.2.11, f (X) must be closed and bounded and thus f (X) = [a1 , b1 ]
[an , bn ] for some ai , bi R with ai bi . This shows that both the supremum
sup f (X) = max{b1 , . . . , bn } and the inmum inf f (X) = min{a1 , . . . , an } are contained
in f (X) and are therefore attained at some points xmax , xmin X.
Theorem 6.3.2. Let (X, d) be a metric space and let F be an open cover of X.
If X is compact then there exists a real number L > 0, called the Lebesgue number of
F, such that for every open ball Bx (L) of radius L (and with x X arbitrary), there
exists an element U F with Bx (L) U .

6.3. PROPERTIES OF COMPACT SPACES

99

Proof. Let y X be an arbitrary element. Since F is an open cover of X, there


must be some element Uy F such that y Uy . Let ry > 0 be chosen so that
y
By (ry ) Uy . Consider then the open cover G of X given by G = {By ( r2 ) | y X}.
r
Since X is compact, G has a nite subcover G = {By1 ( y1 ), . . . , Byn ( ryn )} for some
2
2
points y1 , . . . , yn X. Let L > 0 be chosen that
rx
rx1
,..., n .
L < min
2
2
We claim that this L has the property specied by the theorem. To see this, let
x X be any point. Since G is a cover, there is some index i {1, . . . , n} such that
ry
x Byi ( 2i ). But then, if x Bx (L), we obtain
ry
ry
ry
d(x , yi ) d(x , x) + d(x, yi ) < L + i < i + i = ryi ,
2
2
2
showing that Bx (L) Byi (ryi ) Uyi F. Since x was arbitrary, the theorem is
proved.
The following has already been established for a compact metric space (Theorem
6.2.6). We show here that it remains valid in general compact spaces provided they
are rst countable (which metric spaces always are by Example 2.4.14).
Proposition 6.3.3. Every sequence {xk }k in a rst countable, compact space
(X, TX ) has a convergent subsequence.
Proof. Case 1. Suppose that we can nd a point x X with the property that
every neighborhood U of x contains innitely many elements of the sequence {xk }k .
Let Bx = {Un | n N} be a countable neighborhood basis of x. Dene the subsequence
yn = xkn of {xk }k by picking y1 from U1 {xk | k N} at random, and by picking yn
from the non-empty set U1 ... Un {xk | k > kn1 } in an arbitrary fashion. Then,
given any neighborhood V of x, we can nd a natural number n0 such that Un0 V
and consequently, yn V for all n n0 . In particular, lim yn = x.
Case 2. If the hypothesis from Case 1 fails, then for every point x X there has
to exist a neighborhood Ux containing only nitely many points of the sequence {xk }k .
By compactness of X, we can reduce the open cover {Ux | x X} of X to a nite
subcover {Ua1 , . . . , Uan } with a1 , . . . , an X. But then, by the pigeonhole principle,
one of the sets Uai must contain innitely many elements of the sequence xk , producing
a contradiction. Therefore Case 2 cannot occur while in Case 1 we were able to nd a
convergent subsequence of {xk }k .
The next set of theorems shows that the compactness of X enhancesthe separation axioms T0 T5 that X possesses. By this we mean, for example, that if a compact
space X is T2 then it is automatically T3 , something that if of course false in general
topological spaces. The chief reason for this is that closed subsets of X behave in
many ways like points, given that every open cover of a closed subset of X has a nite
subcover. How exactly this plays out can be seen in the proof of the next theorem.
Theorem 6.3.4. Let X be a compact topological space.

100

6. COMPACTNESS

(a) If X is Hausdor then X is also T3 and T4 .


(b) If X is T3 then it is also T4 .
The next lemma is used in the proof of Theorem 6.3.4, it is the analogue for T3
spaces of what Lemma 4.5.1 is for T4 spaces.
Lemma 6.3.5. A topological space (X, TX ) has the T3 property if and only if for
every closed set A X and every point x X A there exists an open set V X
such that

x V V X A.
Proof. = Let X be a T3 space and let A X be a closed set and x X A
any point. By the T3 property there are disjoint open sets Ux , UA X with x Ux
and A UA . From Ux UA = it follows that Ux X UA . Since X UA is a closed

set, we nd that Ux X UA and since X UA X A we get the desired relation

x Ux Ux X A. Choosing V = Ux establishes one direction of the theorem.


= Suppose that for every closed subset A X and for every point x X A

there exists an open set V with x V V X A. We dene the open sets Ux and
respectively. Then A UA and Ux UA = and so X is a T3
UA as V and X V
space.
Proof of Theorem 6.3.4. (a) The proof of this part of the theorem follows
closely the proof of part (c) of Theorem 6.1.9. Let A X be a closed set and x X A
any point. The Hausdor condition guarantees for every a A, the existence of disjoint
neighborhoods Ua,x and Vx,a of a and x respectively. The collection {Ua,x A | a A}
is an open cover of A and, since A is compact (part (a) of Theorem 6.1.9), it has a
nite subcover {Ua1 ,x , . . . , Uan ,x } for some points a1 , . . . , an A. Let Vx = n Vx,ai .
i=1
Clearly Vx is a neighborhood of x but moreover, Vx and A are disjoint. For if not, we
could nd a point y Vx A. Then we would have y Uaj ,x for some j {1, . . . , n}
and on the other hand y Vx Vx,aj , a contradiction since Uaj ,x Vx,aj = . Let
UA = Ua1 ,x Uan ,x . Then A UA and clearly UA Vx = , showing that X is T3 .
To verify the T4 property of X, let A, B X be two disjoint closed subsets of
X. For a xed point b B, let Vb,A and UA,b be disjoint open sets with b Vb,A and
A UA,b . The set FB = {Vb,A | b B} is an open cover of B which, by compactness of
B, (again, part (a) of Theorem 6.1.9), has a nite subcover FB = {Vb1 ,A , . . . , Vbn ,A } for
some points b1 , . . . , bn B. Set VB = n Vbi ,A and UA = n UA,bi . These are open
i=1
i=1
sets containing B and A respectively. Moreover, VB UA = for if we had a point
y VB UA then y would lie in some Vbj ,A and at the same time UA,bj , a contradiction
since these two sets are disjoint by construction. Therefore X is T4 .
(b) Let A, B X be two disjoint and closed subsets of X. According to Lemma
6.3.5, for every point a A there exists a neighborhood Ua of a such that

a Ua Ua X B.

6.4. THE ONE POINT COMPACTIFICATION

101

Then {X A, Ua | a A} is an open cover of X and has a nite subcover {X


A, Ua1 , Ua2 , . . . , Uan }. In a similar vein, every point b B has a neighborhood Vb with

b Vb Vb X A,
and the open cover {X B, Vb | b B} has a nite subcover {X B, Vb1 , Vb2 , . . . , Vbn }
(for convenience and without loss of generality, we assume that this cover has as many
elements as the cover {X A, Ua1 , Ua2 , ..., Uan }). We now dene the open sets

U1 = Ua1 Vb1 ,
V1 = Vb1 Ua1 ,

U2 = Ua2 (Vb1 Vb2 ),


V2 = Vb2 (Ua1 Ua2 ),
.
.
.
.
.
.
b1 Vbn ),

U = Uan (V
V = Vbn (Ua1 Uan ).
n

Observe that Ui Vj = for all indices i, j {1, . . . , n}. Since Vb1 Vbn X A
n
we see that UA = i=1 Ui is an open set containing A and likewise, VB = n Vj is an
j=1
open set containing B. Moreover, the set UA and VB are disjoint for if they were not
then we would obtain Ui Vj = for some indices i, j, an impossibility. Thus X is
T4 .
6.4. The one point compactication
Definition 6.4.1. A compactication of a topological space (X, TX ) is a compact
topological space (Y, TY ) that contains X as a dense subspace.
The goal of this section is to show that compactications always exist for every
space X. We will describe one method for obtaining a compactication called the One
Point Compactication.
If (X, TX ) is already compact then X is its own compactication. If X is not
compact, let p be an abstract point not contained in X and dene Y (as a set) by
Y = X {p}. Dene TY to be the following collection of subsets of Y :

Either p U and U TX ,
/

U Y
or
(6.1) TY =

p U and X U is a compact closed subset of X.


Remark 6.4.2. Note that if X U is a closed subset of X for any choice of U TY .
Lemma 6.4.3. TY is a topology on Y .
Proof. This is a direct verication of the three axioms of a topology.
1. TY since p and TX .
/
Y TY since p Y and X Y = , the latter being closed and compact in X.
2. Let Ui TY , i I and let U = iI Ui . If for all i I we hve p Ui and
/
Ui TX then p U and U TX showing that U TY . On the other hand, if
/
p Uj for some j I, then X Uj is a closed and compact subset of X. But
then p U and
X U = X iI Ui = iI (X Ui ) X Uj .

102

6. COMPACTNESS

Since each X Ui is closed in X (Remark 6.4.2) we see that X U is a closed


subset of the compact space X Uj and hence itself compact (part (a) of
Theorem 6.1.9). It follows that U TY .
3. Let V1 , ..., Vn TY and set V = V1 ... Vn . If there is at least one index
j {1, ..., n} with p Vj and Vj TX then p V and V TX . If p Vi for
/
/
all i = 1, ..., n then X Vi is closed and compact for all i. But then
X V = X n Vi = n (X Vi ),
i=1
i=1
showing X V to be a nite union of closed and compact subspaces of X.
Accordingly, X V itself is a closed and compact subspace of X showing that
V TY .

We remark that the denition of TY makes it immediate that the subspace topology
on X induced by TY is simply TX . Therefore we can view (X, TX ) as a subspace of
(Y, TY ).
We next turn to the verication of the density of X in Y . Since Y = X {p},

it is clear that X must either be X or Y (since these are the only two subsets of Y

containing X). But X = X could happen only if X were closed in Y . By denition of


TY , this would mean that {p} would have to be open in Y which in turn would mean
that Y {p} = X would need to be a closed and compact subspace of X. But our
assumption was that X is not compact, and so {p} is not open in Y and consequently

X is not closed in Y . This excludes the possibility X = X leaving only X = Y .


It remains to be seen that (Y, TY ) is a compact space. Thus, let F be an open cover
of Y . There must be some element U0 F which contains p and so by the denition
of TY , X U0 is a compact subspace of X. Thus, X U0 is covered by nitely many
U1 , ..., Un F showing that Y is covered by the nite subcover {U0 , U1 , ..., Un } F of
F. Ergo, Y is compact. We summarize our ndings in the next theorem.
Theorem 6.4.4. Let (X, TX ) be a non-compact space and let Y = X {p} for
some abstract point p X. Then (Y, TY ), with TY dened as in (6.1), is a compact
/
space that contains X as a dense subspace. The space Y is referred to as the one point
compactication of X.
Example 6.4.5. Let (X, TX ) = (Rn , TEu ) be n-dimensional Euclidean space and let
p = be the symbol for an abstract point not in Rn . Consider also the n-dimensional
sphere S n dened as
S n = {(x1 , . . . , xn+1 ) Rn+1 | x2 + + x2 = 1},
1
n+1
equipped with the relative Euclidean topology on Rn+1 . We will show that the one point
compactication of Rn is homeomorphic to S n . Dene the function f : Rn {} S n

6.4. THE ONE POINT COMPACTIFICATION

103

as
2x1
2xn
|x|2 1
,..., 2
, 2
|x|2 + 1
|x| + 1 |x| + 1

x = ,

f (x1 , ..., xn ) =

(0, 0, ..., 0, 1)

x = ,

x2 + ... + x2 . We note that f is a bijection with inverse function


1
n

y1
yn

,...,
; y = (0, . . . , 0, 1),

1 yn+1
1 yn+1
f 1 (y1 , ..., yn+1 ) =

; y = (0, . . . , 0, 1).

where |x| =

To see that f and f 1 are continuous we will show that they are open maps. If
U TRn {} is an open set with U or if V S n is an open set with (0, ..., 0, 1) V ,
/
/
then f (U ) and f 1 (V ) are open since f |Rn : Rn S n {(0, ..., 0, 1)} is quite obviously
a homeomorphism (since the coordinate functions of f 1 |S n {(0,...,0,1)} and of f |Rn are
continuous functions).
Next, pick a set U TRn {} with U . According to (6.1), Rn U is a compact
subspace of Rn and so, according to corollary 6.2.11, it is also bounded. This allows
us to pick an R > 0 so that Rn B0 (R) U (where 0 Rn is the origin). But then
B(0,...,0,1) () S n f (U )

with

2
R2

+1

showing that (0, ..., 0, 1) f (U ) has a neighborhood contained in f (U ). By continuity


of f |Rn , every point y f (U ) {(0, ..., 0, 1)} also has a neighborhood contained in
f (U ) and so f (U ) is open and therefore f 1 is continuous.
A neighborhood basis for S n around (0, ..., 0, 1) is given by {B(0,...,0,1) ()S n | > 0}.
To show that f 1 is open, it suces to show that f 1 (B(0,...,0,1) () S n ) is an open
subset of Rn . But this is a completely explicit computation and one nds that

f 1 (B(0,...,0,1) () S n ) = Rn B0 (R) with R =


.
1 1 2
Note that Rn B0 (R) TRn {} . For any open subset V S n {(0, ..., 0, 1)}, f 1 (V )
is obviously open and so f 1 is an open map and therefore f is continuous and thus a
homeomorphism.
Example 6.4.6. Consider the included point topology Tp on R. The space (R, Tp )
is not compact (Example 6.1.6) and so we can dene Y = R {} to be the one
point compactication of (R, Tp ). To describe the elements of TY , note that a subset
A R is compact with respect to Tp if and only if it is nite. For if A = {a1 , a2 , . . . }
is innite and p A, then the set {{a1 }, {a2 }, . . . } is an open cover of A with no
/
proper subcovers at all. On the other hand, if A is innite and p A, then the set
{{a1 , p}, {a2 , p}, . . . } is again an open cover of A with no proper subcovers. A subset

104

6. COMPACTNESS

A R is closed if it doesnt contain p. Consequently, closed and compact subsets of


(R, Tp ) are precisely nite subsets of R that dont contain p. From this we obtain
TY = {U Y | (p U and U ) or (p, U and R U is nite)}.
/
We see that the one point compactication of (R, Tp ) contains features of the nite
complement topology. Observe that every sequence in Y converges to .
6.5. Flavors of compactness
Definition 6.5.1. Let (X, TX ) be a topological space and F an open cover of X.
We shall say that F is locally nite if every point x X has a neighborhood Ux that
intersects only nitely many elements from F.
For later intentions, it is not sucient to dene local niteness to be the property
that every point x X intersects only nitely many sets in F. A priori, it is not
clear that this is actually a weaker than Denition 6.5.1, though Example 6.5.3 below
conrms that it is. Note that every nite cover F is locally nite.
Example 6.5.2. Consider the cover F = { a + 1, a 1 | a Z} of (R, TEu ). This
is locally nite since for every point x R, the interval x 1, x + 1 intersects at most
2 elements from F.
On the other hand, G = { a, b | a, b Q, a < b} is not locally nite since, given
any point x R and any neighborhood Ux of x, there are innitely many rational
numbers contained in Ux . Let q1 , q2 , q3 , Ux Q be an innite sequence chosen so
that qi < qi+1 , then q1 , qi intersects Ux for every i = 2, 3, 4, . . . .
Example 6.5.3. Consider the excluded point topology T p on R and let
F = {U R | U = R or U = {x} for some x R {p} }.
Then F is an open cover of R with the property that every point x R is contained
in either exactly two elements of F (if x = p) or in only one such element (if x = p).
But p R has no neighborhood that intersects only nitely many elements from F.
We conclude that F is not a locally nite cover even though every point in R only
intersects nitely many sets in F.
Definition 6.5.4. Let (X, TX ) be a topological space.
(a) We shall say that X is locally compact if for every point x X and every
neighborhood U of x, there exists a neighborhood V of x such that x V

V U and with V compact.


(b) We call X -compact if there exists a sequence A1 , A2 , A3 , . . . of compact subsets
of X such that X = Ai .
i=1
(c) The space X is said to be paracompact if every open cover F of X has a locally
nite renement F .
Remark 6.5.5. We comment briey on each point from the previous denition
before moving on. Regarding part (a) from Denition 6.5.4, some authors dene local

6.6. EXERCISES

105

compactness of X as meaning that every point x X has a neighborhood with compact


closure. While this is implied by our denition, it is not equivalent to it (Example ??).
If X is a compact space then we can take each Ai from part (b) of Denition 6.5.4
to be X. Thus a compact space is -compact.
Similarly, if X is compact, then every of its open covers F has a nite subcover F .
Since every nite cover is necessarily locally nite, we see that compactness implies
paracompactness.
Example 6.5.6. Euclidean space (Rn , TEu ) is -compact since each of Ai = B0 (i),
i N is compact and clearly Rn = Ai . As we already saw, Rn is not compact.
i=1
Example 6.5.7. The exluded point topology T p on R is not locally compact.
Namely, the point p itself has only one neighborhood, R itself. But we saw in example 6.1.7 that (R, T p ) is not compact
Theorem 6.5.8. Let (X, TX ) be a locally compact, non-compact, Hausdor space
and let (Y, TY ) be its one point compactication (with Y = X {}, see (6.1)). Then
Y is a compact Hausdor space.
Proof. Let x, y Y be two distinct points. If x, y X then, since X is Hausdor
and since TX TY , x and y possess disjoint neighborhoods Ux and Uy .
On the other hand, if y = and thus x = , then x lies in X and local compactness

of X guarantees the existence of a neighborhood Ux of x in X with Ux compact. But


x belongs to TY since Ux is compact and closed (the latter because

then Uy = Y U
it is a compact subspace of a Hausdor space, see part (c) of Theorem 6.1.9). Clearly
Ux Uy = demonstrating that Y is Hausdor.
Corollary 6.5.9. Let X be a locally compact, Hausdor space. Then X is completely regular (part (b) of Denition 4.1.2).
Proof. If X is compact, let (Y, TY ) = (X, TX ) and if X is not compact, let (Y, TY )
be the one point compactication of (X, TX ). Then Y is a compact, Hausdor space
(either by assumption on X or by Theorem 6.5.8) and so Y is a normal space according
to part (a) of Theorem 6.3.4. The Hausdor condition of Y implies that Y has the
T0 property while Corollary 4.5.4 shows that Y is a T3 1 space. The conclusion of
2
the corollary now follows from part (a) of Theorem 4.1.4 by which the subspace of a
completely regular space is again completely regular.
Example 6.5.10. Let X be a locally compact, Hausdor space. Then for any two
distinct points x, y X there is a continuous function f : X R such that f (x) = 0
and f (y) = 1.
6.6. Exercises
6.6.1. Show that (Rn , TEu ) is not compact by nding an explicit open cover of Rn
that has no nite subcovers.
6.6.2. Referring to Denition 6.2.1, prove that

106

6. COMPACTNESS

(a) A totally bounded metric space is bounded.


(b) Show by example that there exist bounded metric spaces that are not totally
bounded.
(c) Consider Rn with the Euclidean metric d2 . Show that a subset A Rn is
bounded if and only if it is totally bounded.
6.6.3. Let (X, d) be a metric space and {xk }k X a Cauchy sequence. Show that
if {xk }k possesses a convergent subsequence yn = xkn with lim yn = y, then {xk }k is
also convergent and lim xk = y.
6.6.4. Show that the conclusion of part (b) of Theorem 6.3.4 remains true under
a weaker hypothesis. Namely, show that every second-countable space X that is T3 is
also T4 . Hint: Emulate the proof of part (b) of Theorem 6.3.4 and use Lemma 6.2.5
instead of compactness of X, to reduced a cover to a countable subcover.
6.6.5. Prove the following converse of Theorem 6.5.8: If the one point compactication (Y, TY ) of (X, TX ) (with Y = X {p} and X non-compact) is a Hausdor space,
then X is locally compact.

6.6.6. Consider the extended real line R = R {} with being two

abstract points, not belonging to R. Consider the collection TR of subsets of R dened

as
U and U TEu , or
/

TR = U R
.

{} U = and R U is closed and compact in R.

(a) Show that TR is a topology on R.

TR ) is a compact space.
(b) Show that (R,

(c) Show that (R, TEu ) is a dense subspace of (R, TR ).
TR ) is homeomorphic to [0, 1] (with the relative Euclidean topol(d) Show that (R,
ogy on [0, 1]).
Conclude that the two-point compactication of the real line is a segment (of positive
length).
6.6.7. Let (X, d) be a metric space and F1 F2 F3 . . . a sequence of nested,
non-empty compact subspaces of X. Show that if lim diam(Fi ) = 0, then iN Fi
i
consists of a single point.
6.6.8. Which of the following subspaces of (Rn , TEu ) is compact?
(a) The sphere S n1 Rn with nitely many points removed.
(b) The graph f of a map f : [0, 1]n1 R.
(c) The Cantor set C R (part (d) of Example 1.5.1).
6.6.9. Find an example of an open cover of (R, TEu ) that has no (positive) Lebesgue
number (compare to Theorem 6.3.2).
6.6.10. Let (X, d) be a compact metric space.
(a) Show that X has nite diameter.

6.6. EXERCISES

(b) Show that there exist points a, b X with d(a, b) = diam(X).

107

CHAPTER 7

Connectedness and Path-Connectedness

tuitively speaking, a disjoint union of two abstract topological spaces consists


of (at least) two components- the constituent topological spaces. A prime
examples of this phenomenon is the two point space obtained by taking the
disjoint union of two singletons, forcing the two point space to carry the
discrete topology. Indeed, this archetypical example of a disconnectedspace can be
used to characterized disconnectedspaces as being precisely those that continuously
surject on it.
The notions of connectedness and disconnectedness of a topological space are explored in Section 7.1. Section 7.2 explores the notion of path-connectedness of a space,
and compares and contrasts it to that of connectedness. While connectedness is implied
by path-connectedness, the converse if false in general. The nal Section 7.3 introduces
local version of connectivity and path-connectivity. The main goal of introducing these
concepts is to understand when connected and path-connected components of a space
are open and closed (Theorem 7.3.6 and Corollary 7.3.7).
7.1. Connected topological spaces
Definition 7.1.1. A topological space (X, TX ) is said to be connected if there is
no continuous surjection f : X {0, 1} where the two point set {0, 1} is given the
discrete topology. Otherwise X is said to be disconnected.
Theorem 7.1.2. Let (X, TX ) be a topological space. Then the following are equivalent:
(a) X is connected.
(b) The only open and closed subsets of X are the empty set and X itself.
(c) If X = A B with A, B disjoint and open sets, then A = or B = .
Proof. (a)=(b) If A X is both open and closed, then so is X A. If both
of these were nonempty, as would be the case if A = , X, then we could dene the
surjective function f : X {0, 1} as
f (x) =

0
1

;
;

if x A,
if x X A.

An easy check reveals that f is continuous, contradicting the assumption from part (a)
that X is connected.
109

110

7. CONNECTEDNESS AND PATH-CONNECTEDNESS

(b)=(c) If X = A B with A and B open, then A and B are also closed (each
being the complement of the other). According to part (b), one of A or B is then the
empty set.
(c)=(a) If f : X {0, 1} is a surjective map, then f 1 ({0}) and f 1 ({1}) are
disjoint subsets of X that are open and closed (since f is continuous and since {0}
and {1} are open and closed subsets of {0, 1}). By surjectivity of f , they are also both
non-empty contradicting part (c). Therefore, no continuous surjection f : X {0, 1}
can exist.
Theorem 7.1.3. Let f : X Y be a continuous function between topological
spaces. If X is connected then so is f (X). In particular, connectedness is a topological
invariant.
Proof. Suppose that f (X) = A B with A, B f (X) disjoint and open sets.
Then f 1 (A) and f 1 (B) are open subsets of X with X = f 1 (A) f 1 (B). By
connectedness of X, one of f 1 (A) or f 1 (B) has to be the empty set, say f 1 (B).
But then B = showing that f (X) is connected.
If X and Y are homeomorphic, each of them is the image of the other under a
continuous map and so they are both connected or both disconnected.
An important corollary of the previous result is the Intermediate Value Theorem.
Corollary 7.1.4 (Intermediate Value Theorem). Let X be a connected space and
f : X R a continuous function and let a, c Im(f ) be two distinct points with a < c.
If b a, c , then there exists and x X with f (x) = b (that is b Im(f )).
Proof. If we could nd an element b a, c Im(f ), we could write
f (X) = (f (X) , b ) (f (X) b, ) .
This would imply disconnectedness of f (X), contradicting the previous theorem and
the assumption that X is connected.
Example 7.1.5. If X = {x} is a set with only a single point, then it is connected
for there is no surjection from a set with one element onto a set with two elements.
Example 7.1.6. The excluded point topology makes R into a connected space.
This follows from the observation that if R = A B with A, B open, then if p A we
get A = R and if p B then B = R. In either case, A and B can only be disjoint if
one of them is the empty set.
Example 7.1.7. The set of rational number Q R with its
relative Euclidean
topology, is disconnected. This can be seen by setting A = , 2 Q and B =

2, Q. Then A, B Q are two open and disjoint sets and Q = A B.


Example 7.1.8. If X is any set with at least two elements, then (X, Tdis ) is disconnected. On the other hand, for any set Y , the space (Y, Tindis ) is connected (here
Tdis is the discrete and Tindis is the indiscrete or trivial topology).

7.1. CONNECTED TOPOLOGICAL SPACES

111

Example 7.1.9. With the nite complement topology, R is a connected space


since there are no non-empty disjoint open subsets of R. The same holds true with the
countable complement topology.
In the statement of the next theorem we shall use the word interval to mean any
of a, b , a, b], [a, b or [a, b], including intervals of innite length. Said dierently,
intervals are precisely the convex subsets of R. The next theorem characterizes the
connected subsets of the Euclidean line.
Theorem 7.1.10. Consider the Euclidean line (R, TEu ). The nonempty connected
subspaces of R are precisely the intervals in R.
Proof. = Let X be a nonempty connected subspace of R. If X contains only
one point then it is an interval and the claim of the theorem is proved. Suppose that
X contains at least two points, say x and y, and suppose that x < y. If z R is any
point with x < z < y then z has to belong to X. For if not, then X could be written
as the union
X = ( , z X) ( z, X),
with both sets on the right hand side being open and nonempty (and clearly disjoint)
contradicting the connectedness assumption of X. Therefore z X. Let = inf X and
L = sup X, then X equals an interval from to L, the boundaries may be included or
not and may be innite or not.
= Let X = a, b be an open interval for some a < b and assume that X can be
written as X = A B with A, B disjoint, open subsets of X and with A non-empty.
We shall show that then B is forced to equal the empty set.
Let x A be any element. Since X is open in R and A is open in X, then A is also
open in R. Thus there exists an open interval around x that is contained in A, making
the following denitions meaningful:
= inf{r , x | r, x] A}

and

L = sup{s x, | [x, s A}.

For each n N, there exist rn , sn R such that


1
1
0 < rn <
and 0 < L sn < ,
n
n
and rn , sn A. Thus , L = nN rn , sn is also contained in A. However, since A
is an open interval, neither or L can be continued in A.
If we had a < , it would follow that B and since B is open in R, an interval
around would be contained in B. This in turn would force B to intersect A nontrivially, a contradiction, showing that = a. In a similar vein it follows that L = b and
thus that A = X, implying B = .
The cases of X being one of the other types of intervals follow similarly, and we
leave them as an exercise (Exercise 7.4.1). We add that these additional cases are also
implied by the next theorem using the current theorem with X = a, b .
Theorem 7.1.11. Let (X, TX ) be a topological space and let Y be a subspace of X.

If Y is connected and Z X is any set with Y Z Y , then Z is also connected.

112

7. CONNECTEDNESS AND PATH-CONNECTEDNESS

Proof. Let Z Y be any set with Y Z and write Z = A B with A, B open


and disjoint subsets of Z. Then AY = A Y and BY = B Y are open and disjoint
subsets of Y with Y = AY BY . By connectedness of Y , one of AY or BY is the empty

set, say BY = . Let B Y be an open set with B = Z B , then Y Y B .

Since B is open, Y B is closed and contains Y and so Y Y B . This shows that


B = and therefore also B = Z B = . It follows that Z is connected.
Corollary 7.1.12. The closure of a connected subspace is connected. If a space
X possesses a connected dense subset, then X itself is connected.
Example 7.1.13. Consider the included point topology Tp on R. The closure of
{p} is all of R (since R is the only closed set containing p) and since a set with only
one element is always connected (Example 7.1.5), it follows from Theorem 7.1.11 that
(R, Tp ) is connected.
Lemma 7.1.14. Let (X, TX ) be a topological space and Yi X, i I be a family of
connected subspaces. If iI Yi = then iI Yi is connected.
Proof. Let p iI Yi be any point and let A, B iI Yi be two open and
disjoint subsets of iI Yi with iI Yi = A B. Then Ai = A Yi and Bi = B Yi are
open and disjoint subsets of Yi with Yi = Ai Bi . By the connectedness assumption
of Yi , one of Ai or Bi needs to be the empty set. Suppose that for a given index i0 I
we had Bi0 = and notice that then p Ai0 A. For any other index i I, we
cannot have Ai = since that would imply that p Bi B, a contradiction since we
already found that p A. Therefore we obtain that Bi = for every i I. But then
B = iI Bi = showing that iI Yi is connected.
Definition 7.1.15. Let (X, TX ) be a topological space.
(a) A connected component U of X is any maximal connected subset of X. Said
dierently, U is a connected component of X if, whenever V is a connected
subspace of X and U V , then U = V .
(b) We say that X is totally disconnected if every connected component of X consists of just a single point.
Theorem 7.1.16. Let (X, TX ) be a topological space.
(a) X is the disjoint union of its connected components.
(b) Every connected component of X is a closed subset of X. In particular, if X
has nitely many components, then each component is both open and closed.
(c) X is connected if an only if it has precisely one connected component.
Proof. (a) Let x X be any point and let Ux be the subset of X obtained by
Ux = U Ux U

with

Ux = {W X | x W and W is connected}.

By Lemma 7.1.14 Ux is connected and it is clearly a maximal connected subset of X.


Thus every point x X belongs to a connected component of X showing that X is
the union of its connected components.

7.1. CONNECTED TOPOLOGICAL SPACES

113

Let U, V X be two dierent connected components of X. If we had U V =


then U V would be a connected space (by Lemma 7.1.14), a contradiction to the
maximal property of U and V . Thus U V = .

(b) If U is a component of X then U is also connected (Theorem 7.1.11) and so by

maximality, U = U . Accordingly, U is closed.


(c) This follows directly from the denition. If the only component of X is U then,
since every component is connected, so is X. If X is connected, let Ui , i I be
its components. Then X is a connected set containing Ui and so by the maximallity
property of Ui , we must have X = Ui for all i I.
We note that part (c) of the above theorem cannot in general be improved by
claiming that connected components of X are always open subsets (however, compare
to Corollary 7.3.7). Here is an example demonstrating this.
Example 7.1.17. Consider the set of rational numbers Q R equipped with the
relative Euclidean topology. If Y Q is any subset with at least two elements, say
y1 , y2 Y with y1 < y2 , then
A = , Y

and

B = , Y,

with y1 , y2 (R Q) arbitrary, are two disjoint, open and non-empty subsets of


Y with Y = A B. Thus, any such Y is disconnected showing that the only connected
subsets of Q are sets with only one element. Accordingly, Q is totally disconnected
with closed (but not open!) components {q}, q Q.
Example 7.1.18. Consider the lower limit topology Tll on R. The connected components (R, Tll ) are the single point sets {x}, x R since whenever A R has at least
two elements a, b A with a < b, then A can be written as the union of two disjoint
and non-empty open subsets of A:
A = ( , b A) ([b, A)
This shows that any such A is disconnected. Consequently, (R, Tll ) is completely disconnected.
Example 7.1.19. Consider the Euclidean plane (R2 , TEu ) and let Y be its subspace
given by
Y = {(x, y) R2 | xy = 0 and (x, y) = (0, 0)}
Geometrically, Y is the union of the x-axis with the y-axis and with the origin removed.
Then Y has 4 connected components and they are:
U+ = {(x, 0) R2 | x > 0} = Positive half of the x-axis.
U = {(x, 0) R2 | x < 0} = Negative half of the x-axis.
V+ = {(0, y) R2 | y > 0} = Positive half of the y-axis.
V = {(0, y) R2 | y < 0} = Negative half of the y-axis.

114

7. CONNECTEDNESS AND PATH-CONNECTEDNESS

To see that these are indeed the components of Y , note rstly that each of U and V
is homeomorphic to R and therefore connected (Theorem 7.1.10). To see that U+ is
a maximal connected subspace of Y , let A, B R2 be the open and disjoint subsets
given by
A = {(x, y) R2 | x < |y|}

and

B = {(x, y) R2 | x > |y|}

Then U+ A and U V+ V B. If U+ were not a maximal connected subspace

of Y and were contained in a larger connected subset U Y , then A U and B U

would be disjoint, open and non-empty subsets of U with U = (A U ) (B U ), a


. Consequently, U+ is a component of Y , the
contradiction to the connectedness of U
cases of U , V+ , V are handled analogously.
7.2. Path-connectedness
Definition 7.2.1. Let (X, TX ) be a topological space.
(a) A path in X is a continuous function : [0, 1] X where [0, 1] is equipped with
the relative Euclidean topology. The points (0), (1) are called the initial (or
starting) and terminal (or ending) points of . We will say that joins x to y
or that x is joined to y by .
(b) If : [0, 1] X is a path in X then we shall call the path : [0, 1] X given

by (t) = (1 t), the inverse path of .

(c) Given two paths , : [0, 1] X with (1) = (0), we dene the product path
: [0, 1] X of and as

1
; t [0, 2 ],
(2t)
(t) =

1
(2t 1)
; t [ 2 , 1].
A couple of remarks are in order. If x is joined to y by means of a path , then
y is joined to x by the path . The function from part (c) of Denition 7.2.1 is

continuous by Lemma 3.1.12. If joins x to y and joins y to z then joins x to


z. With this in mind, we are ready to introduce the main concept of this section.
Definition 7.2.2. A topological space X is called path-connected if for every two
points x, y X there exists a path in X with (0) = x and (1) = y.
Theorem 7.2.3. Let (X, TX ) and (Y, TY ) be two topological space and let f : X Y
be a continuous function. If X is path-connected then so is f (X). In particular, pathconnectedness is a topological invariant.
Proof. Given two points x, y f (X), let a, b X be points with f (a) = x and
f (b) = y. Let : [0, 1] X be a path connecting a to b, then f is a path in f (X)
connecting x to y.
If X and Y are homeomorphic spaces then they are each the image of the other.
Consequently they are either are both path-connected or neither of them is.

7.2. PATH-CONNECTEDNESS

115

Theorem 7.2.4. If X is path connected then it is connected. The converse is


generally not true (Example 7.2.7).
Proof. Suppose X were not connected. Then we could write X = A B with
A, B two open, disjoint and non-empty sets. Let a A and b B be any two points
and let : [0, 1] X be a path joining a to b. Then the sets
A = 1 (A)

and

B = 1 (B),

are both open (since is continuous), non-empty (since 0 A and 1 B ) and disjoint
(since A and B are disjoint) subsets of [0, 1]. This implies that [0, 1] is disconnected, a
contradiction to Theorem 7.1.10. Therefore X must be connected.
Example 7.2.5. Euclidean n-dimensional space (Rn , TEu ) is path-connected and
hence also connected. Given any two points x, y Rn , the path : [0, 1] Rn given
by (t) = x + t(y x) starts at x and ends at y.
Example 7.2.6. For any point x Rn and for any r > 0, the ball Bx (r) is
path-connected (and thus also connected). This is seen by observing that every point
y Bx (r) can be connected to x via the path y : [0, 1] Bx (r) dened by y (t) =
y + t(x y). Given any pair of points y1 , y2 Bx (r), the path y1 y2 connects y1 to

y2 .
Example 7.2.7. (Of a connected but not path connected space) Let X be the topologists sine curve from Example 2.2.15, that is let X be the subspace of the Euclidean
plane (R2 , TEu ) given by
1
X = {(x, sin x ) | x 0, 1]} ({0} [0, 1]) R2 .
1
Note that the set Y1 = {(x, sin( x )) | x 0, 1]} X is homeomorphic to 0, 1] and is

thus connected by Theorem 7.1.10. According to Theorem 7.1.11, the closure Y1 of Y1

is then also closed as is any set Z R2 with Y1 Z Y1 . We claim that Z = X

satises this double inclusion showing that it is connected. To show that X Y1 , it


suces to show that every element in Y2 = {0} [0, 1] is the limit of a convergent
sequence {xk }k Y1 . This is easily seen for if (0, y) Y2 , let xk = (tk , y) Y1 be any
sequence with tk a convergent sequence with limit 0. For example, choosing tk as
1
tk =
with k N and t1 = arcsin |[ , 3 ] (y),
2 2
t1 + 2k
will do.
We will now show that X is not path-connected. If we suppose to the contrary
that X is path-connected, then we can nd a path : [0, 1] X connecting (0, 0) to
(1, sin 1). Since the y-axis is a closed subset of R2 , the set

A = 1 (y-axis X),
is a closed subset of [0, 1] and it is nonempty since it contains 0. Let t0 A be any
point and suppose that (t0 ) = (0, b0 ) for some b0 [0, 1]. Consider the neighborhood

116

7. CONNECTEDNESS AND PATH-CONNECTEDNESS

V of (0, b0 ) in X given by
V =

1
1 , 1 b0 2 , b0 +
2 2

1
2

X.

By continuity of , there exists a > 0 so that ( t0 , t0 + [0, 1]) V . Since


t0 , t0 + [0, 1] is path-connected, so is its image (Theorem 7.2.3), and it must
therefore lie in the path-connected component of V that contains (t0 ). This however
is simply V y-axis (see Figure 1) and so t0 , t0 + [0, 1] is a subset of A. This
shows that A is both open and closed and so is therefore B = [0, 1] A. Likewise, both
A and B are nonempty seeing as they contain 0 and 1 respectively. This contradicts
the connectedness of [0, 1] and so our supposed path cannot exist. Thus X is not
path-connected.
0.5

.....

0.5

0.5
.....

0.5
Figure 1. The path-connected components of V for the case of (t0 ) =
(0, 0). The picture looks similar for lather values of (t0 ).
With our investigations of connectedness and path-connectedness thus far, we can
now partially answer Question 3.4.1 about when Euclidean n-space (Rn , TEu ) and Euclidean m-space (Rm , TEu ) can be homeomorphic.
Corollary 7.2.8. If the Euclidean line (R, TEu ) is homeomorphic to Euclidean
n-dimensional space (Rn , TEu ), then n = 1.
Proof. Suppose that f : R Rn is a homeomorphism. Then f |R{0} : R {0}
R {f (0)} is also a homeomorphism. However, R {0} is not an interval and
is therefore not connected by Theorem 7.1.10. On the other hand, we claim that
Rn {f (0)} is even path-connected, and therefore connected, when n 2. To see this,
let x, y Rn be any two points and consider the straight line path (t) = x + t(y x).
n

7.2. PATH-CONNECTEDNESS

117

If f (0) does not lie on , then is a path in Rn {f (0)} from x to y. If f (0) does
lie on , let z Rn be any point not collinear with x and y and consider the straight
line paths (t) = x + t(z x) and (t) = z + t(y z) connecting x to z and z to y
respectively. Then is a path from x to y in Rn {f (0)} showing that Rn {f (0)} is
path-connected. Since connectedness is a topological invariant, it follows that n must
be 1.
Theorem 7.2.9. An open subset U of Euclidean space Rn is connected if and only
if it is path connected.
Proof. In view of Theorem 7.2.4, we only need to show that if U is connected
then it is also path-connected. Let x U be any point and dene
A = {y U | x and y can be joined by a path in U },
B = {y U | x and y can not be joined by a path in U }.
Clearly X = A B and A = since x A. We will show that both A and B are open
subsets of U . Since U is connected, this will imply that B = and A = U , as desired.
To see that A is open, let y A be any point and let : [0, 1] U be a path joining
x to y. Let > 0 be such that By () U and for any z By () let z : [0, 1] U
be the radial path from y to z, that is z (t) = (1 t) y + t z. Then z is a path
from x to z showing that By () A. Since y A was arbitrary, we conclude that A
is open.
To see that B is open, we proceed similarly. Let y B be any point and let > 0
be such that By () U . If there were a point z By () A, there would have to be
a path : [0, 1] U from z to x. Letting z be as in the previous paragraph, the
product path z would be a path from y to x contradicting our choice of y B.
Therefore By () is contained in B and hence B is open.
Theorem 7.2.10. Let (X, TX ) and (Y, TY ) be topological spaces and let X Y be
given the product topology. Then X Y is (path-)connected if and only if each of X
and Y are (path-)connected.
Proof. Since X and Y get surjected on by the projection maps X : X Y X
and Y : X Y Y , (path-)connectedness of X Y implies that of X and Y (Theorem
7.1.3 and Theorem 7.2.3).
In the other direction, assume that X and Y are connected. Note that each of
X {y0 },

{x0 } Y

and Z = X {y0 } {x0 } Y

is a connected subspace of X Y . The connectedness of the rst two these follows


from Corollaries 5.1.9 and 7.1.3, for the third it follows from Lemma 7.1.14 (since
X {y0 } {x0 } Y = {(x0 , y0 )} = ). To show that X Y is connected, write X Y
as the union A B with A, B open and disjoint subsets of X Y . By connectedness,
each (X {y0 }) ({x0 } Y ) with x0 X and y0 Y arbitrary, has to be contained
entirely in either A or B. The same is true of (X {y1 }) ({x0 } Y ) for every other

118

7. CONNECTEDNESS AND PATH-CONNECTEDNESS

y1 . But
[(X {y0 }) ({x0 } Y )] [(X {y1 }) ({x0 } Y )] = {x0 } Y,
so that both of (X {y0 }) ({x0 } Y ) and (X {y1 }) ({x0 } Y ) lie in the same
set, either A or B. Since y0 , y1 Y were completely arbitrary, we see that all of X Y
lies in either A or B. Thus, one of A or B has to be the empty set, as claimed.
Finally, if X and Y are path-connected then so is X Y . Namely, given two points
(x1 , y1 ), (x2 , y2 ) (X Y ), let : [0, 1] X and : [0, 1] Y be paths joining x1 to
x2 and y1 to y2 respectively. Then : [0, 1] X Y is a path joining (x1 , y1 ) to
(x2 , y2 ).
The following lemma is the path-connectedversion of Lemma 7.1.14.
Lemma 7.2.11. Let (X, TX ) be a topological space and Yi X, i I be a family
of path-connected subspaces. If iI Yi = then iI Yi is a path-connected subspace of
X.
Proof. Let p iI Yi be any point and for x Yi , let x be a path in Yi
connecting x to p. Given any x, y iI Yi , the path x y connects x to y.

Definition 7.2.12. Let (X, TX ) be a topological space. A path-connected component of X is any maximal path-connected subspace U of X. Maximallity here refers
to the property that if V is a path-connected subspace of X with U V then U = V .
Example 7.2.13. The topologists sine curve dened in Example 2.2.15 and considered again in Example 7.2.7, has two path connected components, namely Y1 =
1
{(x, sin x ) R2 | x 0, 1]} and Y2 = {0} [0, 1], but only has one connected component.
The following is the analogue of Theorem 7.1.16 for the path-connected case.
Theorem 7.2.14. Let (X, TX ) be a topological space.
(a) X is the disjoint union of its path-connected components.
(b) X is path-connected if and only if it has a single path-connected component.
Proof. (a) For a given x X, the path-connected component Vx containing it
can be obtained as
Vx = V V V

with

V = {W X | x W and W is path connected}.

The set Vx is path-connected by Lemma 7.2.11, and it is clearly maximal with respect
to this property. Two path-connected components U, V X are disjoint for if they
were not, Lemma 7.2.11 would imply that U V is also path-connected, contradicting
the maximallity of both U and V .
(b) This proof is the complete analogue of the proof of part (c) of Theorem 7.1.16
and is left as an exercise.
Unlike in the case of connectedness (part (b) of Theorem 7.1.16), a path-connected
component of X need not be closed nor open.

7.3. LOCAL CONNECTIVITY

119

Example 7.2.15. Consider the topologists sine curve X from Examples 2.2.15 and
7.2.7). The two subsets Y1 , Y2 X dened as
Y1 = {(x, sin 1/x) | x 0, 1]},

Y2 = {0} [0, 1],

are path-connected seeing as they are homeomorphic to 0, 1] and [0, 1] respectively.


Since X = Y1 Y2 and X is not path connected (Example 7.2.7), Y1 and Y2 are the
only two path-connected components of X. It is easy to see that Y2 is closed in X for
Y2 = X ({0} R). However, Y2 is not open in X for if it were we could nd an open
set U R2 with Y2 = U X. But then, given any point (0, y) Y2 , we would need to
be able to nd an > 0 with B(0,x) () U , an impossibility since each such B(0,y) ()
will intersect Y1 . We conclude that
The path connected component Y1 is open but not closed in X.
The path connected component Y2 is closed but not open in X.
The root for this phenomenon lies in the lack of a path-connected analogue of part (b)
of Theorem 7.1.11.
7.3. Local connectivity
Definition 7.3.1. Let (X, TX ) be a topological space.
(a) We say that X is locally connected if every point x X has a neighborhood
basis Bx consisting of connected open sets. Said dierently, we require that for
every point x X and every neighborhood U of x there exists a connected
open set V X with x V U .
(b) We say that X is locally path connected if every point x X has a neighborhood
basis Bx consisting of path connected sets. Said dierently, we require that
for every point x X and every neighborhood U of x there exists an pathconnected open set V X with x V U .
The next lemma is the local analogue of Theorem 7.2.4. The proof is omitted as it
is largely identical to that of Theorem 7.2.4.
Lemma 7.3.2. A locally-path connected space is locally connected.
Example 7.3.3. Every open subset U of Euclidean space (Rn , TEu ) is locally-path
connected and hence locally connected. This is obvious since, given any point x Rn
and given any neighborhood of U of x, there exists an > 0 with Bx () U . Any
such Bx () is path-connected by Example 7.2.6.
The notions of path-connectedness and of local path-connectedness are independent
in that neither implies the other. Here are examples to substantiate this claim.
Example 7.3.4 (A path-connected but not locally path-connected space). Let X
R be the innite broom from Example 2.2.16, that is let X be the subspace of the
2

120

7. CONNECTEDNESS AND PATH-CONNECTEDNESS

Euclidean plane given by


X=

In
n=0

with

In =

1
{(x, n x) R2 | x [0, 1]} ;

{(x, 0) R2 | x [0, 1]}

n N,
n = 0.

It is easy to see that X is path-connected (and hence also connected), for is x X is


any point then x : [0, 1] X, given by x (t) = (1t)x, is a path connecting x to the
origin. Given any two points x1 , x2 X, the path x1 x2 connects the former to the

latter. However, X is not locally connected and hence also not locally path connected.
1
To see this, consider the point x0 = (1, 0) X and its neighborhood U = Bx0 ( 2 ) X.
Then U contains no connected neighborhood of x0 . For if V U were a connected
neighborhood of x0 , we could nd some > 0 so that Bx0 () X V . But every
such set has innitely many path connected (and hence connected) components as is
evident from Figure 2.

Bx0 ()

.
.
.
.
.
.
.
.
.

.
.
.
Bx0 ()

Figure 2. The two neighborhoods U and Bx0 () from Example 7.3.4.


The intersection of the innite broom X with Bx0 () is an innite disjoint
union of intervals and is therefore not connected.
Example 7.3.5 (A locally path-connected but not path-connected space). The
subspace X = [0, 1] [2, 3] R of the Euclidean line is both locally connected and
locally path-connected by not connected or path-connected.
Every discrete space (X, Tdis ) with at least two points is locally connected and
locally path-connected by not connected or path-connected.
Theorem 7.3.6. Let (X, TX ) be a topological space.
(a) X is locally connected if and only if every connected component of every open
set U X is an open subset of X.
(b) X is locally path-connected if and only if every path-connected component of X
is an open subset of X.

7.4. EXERCISES

121

Proof. (a) = Suppose that X is locally connected, let U X be a an open


subset and let U0 U be a connected component of U . Then every point x U0
has a connected neighborhood Ux,0 . It is necessary that Ux,0 be contained in U0 for
if not, then U0 Ux,0 would be a connected subspace of U (Lemma 7.1.14), violating
the maximallity condition from the denition of a connected component (Denition
7.1.15). Given this, we can write
U0 = xU Ux,0
showing that U0 is an open set.
= Suppose the connected components of open subsets of X are open sets. Let
x X be any point and U any neighborhood of x. Then Ux , the connnected component
of U that contains x, is a connected, open set with x Ux U , showing that X is
locally connected.
(b) This part of the proof is a verbatim repeat of the proof of part (a), but substituting connected with path-connected.
The next corollary follows from Theorems 7.1.16 and 7.3.6.
Corollary 7.3.7. In a locally connected space, the connected components are both
open and closed. Similarly, in a locally path-connected space, the path connected components are both open and closed.
Proof. The connected case is immediate from Theorems 7.1.16 and 7.3.6.
For the path-connected case, let X be a path-connected space and let Ui , i I be
its path-connected components. If x Ui and Ux is a path-connected neighborhood of
x, then Ux Ui by Lemma 7.2.11. Thus Ui = xUi Ux showing that Ui is open. But
then jI{i} Uj is also open and equal to X Ui showing that Ui is closed.
7.4. Exercises
7.4.1. Using an approach as in the second half of the proof of Theorem 7.1.10, show
that intervals of the type [a, b , a, b] and [a, b] are connected (without appealing to
Theorem 7.1.11).
7.4.2. For n 2, let A Rn be the subspace consisting of all points with at least
one rational coordinate. Is A (path-)connected?
7.4.3. Show that each of the included and excluded point topologies Tp and T p ,
render R a path-connected space.
7.4.4. Examine whether the irrational numbers in R are (path-)connected with
respect to the topology T {Tp , T p , Tcc , TF ,p }. Determine its (path-)connected components in each case.
7.4.5. Let X be a topological space and An X, n N, a family of (path)connected subspaces. If An An+1 = , show that A = An is a (path-)connected
n=1
subspace of X.

122

7. CONNECTEDNESS AND PATH-CONNECTEDNESS

7.4.6. Find an example of a topological space X with a disconnected subspace Y

such that the closure Y is connected.


7.4.7. Show that no two among the spaces 0, 1 , [0, 1 and [0, 1] are homeomorphic.
7.4.8. Let A R2 be a countable subset, and consider R2 equipped with the
Euclidean topology. Show that R2 A is a path-connected space.
7.4.9. Find an example of a topological space X and a connected subspace Y such
that

(a) Y is disconnected.
(b) Y is disconnected.
7.4.10. Show that a connected and locally path-connected space is path-connected.

CHAPTER 8

Quotient spaces

his chapter introduces a powerful method for constructing new topological


spaces from given ones by means of taking quotients. This construction is
commonplace in set theory and we extend it here to the topological setting
by endowing the quotient set with a natural topology - the quotient topology.
Section 8.1 introduces quotient spaces, presents rst examples and explores some of
the properties of quotient spaces, and maps from and into quotient spaces. Of the
remaining sections, each focuses on a particular set of examples.
8.1. Quotient spaces: Denitions, properties and rst examples
We start this section by reviewing quotient sets rst, postponing the relevant topology for a couple of paragraphs. We introduce three mutually equivalent but slightly
dierent viewpoints of quotient sets. We shall use each of these approaches below for
dealing with quotient spaces.
Definition 8.1.1. Let X be a non-empty set.
(a) The quotient set X/ associated to a surjective function : X Y onto a
non-empty set Y is dened to be X/ = Y .
(b) The quotient set X/ associated to an equivalence relation on X is the set
of equivalence classes: (X/ ) = {[x] | x X} with [x] = {x X | x x}.
(c) The quotient set X/P associated to a partition P = {Pi | i I} of X is dened
as X/P = I.
The three notions of quotient sets from Denition 8.1.1 coincide with one another,
provided the surjection , the equivalence relation and the partition P are properly
related. Here are the details:
induces Let : X Y be a surjective function and dene the relation on
X as
x 1 x2

if and only if

(x1 ) = (x2 ).

Then the function (X/) (X/ ) given by x [x] is a bijection of quotient sets.
induces P Given an equivalence relation on X, we dene the partition P on
X by
P = {[x] | x X},
123

124

8. QUOTIENT SPACES

where [x] is as in part (b) of Denition 8.1.1. The function (X/ ) (X/P) given
by [x] [x] (where [x] is thought of an equivalence class and as an element of the
partition P respectively) is again a bijection between quotient sets.
P induces Let P = {Pi | i I} be a partition of X with Pi = for all i I.
We dene : X Y by setting Y = I and
(x) = i

if

x Pi .

The quotient sets (X/P) and (X/) are equal by Denition 8.1.1.
Notice that our considerations establish the commutativity of the diagram below
in which each map is a bijection.
/ {Equivalence relations on X.}
Surjections : X Y .
O

gggg
gggg
gggg
gggg
sgggg

Partitions P = {Pi | i I}
of X with Pi = , i I.
Example 8.1.2. Let X = {1, 2, 3, 4, 5, 6}, let Y = {e, o} and dene : X Y by
(1) = (3) = (5) = o

and

(2) = (4) = (6) = e.

Dene the relation on X by


ab

if and only if

b a is even.

Finally, let P be the partition of X given by P = {{1, 3, 5}, {2, 4, 6}}. Then , and
P are related as above and each of X/, X/ and X/P, is a set of two elements.
With these preliminaries out of the way, we now turn to the topology of quotient
sets. We state our denition in terms of a surjective function : X Y though we
could equally well use equivalence relations or partitions.
Definition 8.1.3. Let (X, TX ) be a topological space, let Y be a any non-empty
set and let : X Y be a surjective function.
(a) The quotient topology TX/ on Y , induced by X and (or simply the quotient
topology), is dened by
TX/ = {U Y | 1 (U ) TX }.
(b) For topologies TX and TY on X and Y respectively, the surjective function
: (X, TX ) (Y, TY ) is called a quotient map if it has the property that
U TY

if and only if

1 (U ) TX .

It is quite straightforward to verify that TX/ is a topology on Y . Clearly the empty


set and Y belong to TX/ since 1 () = and 1 (Y ) = X. The second and third
axiom of a topology (Denition 2.1.1) follow from the equalities
1 (iI Ui ) = iI 1 (Ui )

and

1 (iI Ui ) = iI 1 (Ui ),

where Ui Tx/ , i I (these two equalities were the subject of Exercise 3.5.1).

8.1. QUOTIENT SPACES: DEFINITIONS, PROPERTIES AND FIRST EXAMPLES

125

Remark 8.1.4. While Denition 8.1.3 seemingly denes two new terms, those of
a quotient space and of a quotient map, each of these determines the other. It is easy
to see (Exercise 8.7.1) that : (X, TX ) (Y, TY ) is a quotient map if and only if
TY = TX/ .
We postpone the exploration of concrete examples of quotient spaces for the moment
and turn to some of their general properties instead.
Proposition 8.1.5. Let (X, TX ) be a topological space, let : X Y be a surjective function onto the set Y and let TX/ be the associated quotient topology.
(a) The quotient map : (X, TX ) (Y, TX, ) is continuous.
(b) The quotient topology TX/ on Y is the nest topology (Denition 2.1.3) for
which is continuous. Said dierently, if TY is any topology on Y for which
: (X, TX ) (Y, TY ) continuous, then TY TX, .
(c) If f : (X, TX ) (Y, TY ) is a continuous surjection that is additionally also
either open or closed, then f is a quotient map.
Proof. (a) Immediate from Denition 8.1.3.
(b) If TY is a topology on Y for which the map : X Y is continuous, then for
every U TY the set 1 (U ) belongs to TX . Thus any such U automatically lies in
TX/ verifying the claim.
(c) Let f : (X, TX ) (Y, TY ) be a continuous surjection and let U TY be any
set. By continuity of f , f 1 (U ) lies in TX . Suppose that f is also an open map and
that U TY is a set such that f 1 (U ) TX . The openness property of f implies then
that the set f (f 1 (U )) = U is open in Y . Thus TY = {U Y | f 1 (U ) TX } is the
quotient topology associated to X and f .
If f is closed rather than open and U Y is again a set with f 1 (U ) TX , then
X f 1 (U ) is a closed set and hence so is f (X f 1 (U )) = Y U . Thus U is again
an open set and we arrive at the same conclusion.
We next investigate when functions on a quotient space are continuous. For this
purpose, let (X, TX ) be topological space, let : X Y be a surjection and assume
that Y is given the quotient topology TX/ . Let (Z, TZ ) be another topological space

and let f : Y Z be a function. Associated to f is the function f : X Z dened

by f = f , see commutative diagram below.


/
Y
X

~~
~~
~~ f
 ~~~

Z
(x1 ) = f (x2 ), a fact that we will express by saying

Notice that if x1 , x2 (y) then f


1

that f is constant on (y) for each y Y . Conversely, given a function f : X Z


1
that is constant on (y) for each y Y , we can dene f : Y Z as
1

f (y) = f (xy )

for any choice

xy 1 (y).

126

8. QUOTIENT SPACES

The constancy of f on 1 (y) guarantees that f is well dened. Thus, there is a

bijective correspondence between functions f : Y Z and functions f : X Z, the


1
latter of which need to be constant on (y) for all y Y . The next theorem shows
that this correspondence remains bijective after restricting to the subset of continuous
functions in each of these sets.
Theorem 8.1.6. Let : (X, TX ) (Y, TX/ ) be a quotient map and let (Z, TZ ) be
a topological space. A function f : Y Z is continuous if and only if the function

f : X Z given by f = f , is continuous.

Proof. If f : Y Z is continuous then clearly so is f : X Z, being the

composition of two continuous functions. If on the other hand f is continuous, then,


1
for any open set V Z, consider f (V ). This is open in Y if and only if 1 (f 1 (V ))
is open in X. But

1 (f 1 (V )) = (f )1 (V ) = f 1 (V ),

and f 1 (V ) is open in X by continuity of f . Thus f too is continuous.


The situation of functions to a quotient space is somewhat more precarious. Certainly given a continuous function g : Z X, the induced function g : Z Y given

by g = g is again continuous, see commutative diagram below.

/
X
O
>Y
g

~
~~
~~
~ g
~~

Z
But given a continuous function g : Z Y , there need not be a continuous function
g : Z X with g = g . Instances of this type abound, for illustration see Examples

?? and ??.
In the concrete examples of quotient spaces considered in Section 8.2, it will sometimes prove advantageous to construct a quotient space in several stages. What is
meant by this is that we rst form an intermediatequotient space from a given space
X, say by moding out by a partition P1 of X, and then further moding out by a partition P2 of X/P1 to arrive at (X/P1 )/P2 . The next theorem show that the latter can
be arrived at by just a single quotient space construction.
Theorem 8.1.7. Let (X, TX ) be a topological space, let P1 = {Ui X | i I1 } be
a partition of X and let P2 = {Vi X/P1 | i I2 } be a partition of X/P1 . Then
(X/P1 )/P2 X/P3 ,
=
1
where P3 is the partition of X given by P3 = {Wi X | Wi = 1 (Vi ), i I2 }. Here
1 : X X/P1 is the quotient map.

Proof. Let 2 : (X/P1 ) (X/P1 )/P2 and 3 : X X/P3 be the additional two
quotients maps. Note that as sets, each of (X/P1 )/P2 and X/P3 are dened to equal
the indexing set I2 . Thus we dene f : (X/P1 )/P2 X/P3 to simply be the identity.
The four maps 1 , 2 , 3 and f t into the commutative diagram:

8.1. QUOTIENT SPACES: DEFINITIONS, PROPERTIES AND FIRST EXAMPLES


1

X
3

f =id

X/P3

127

X/P1


(X/P1 )/P2

Showing that f is a homeomorphism is equivalent to showing that a subset W X/P3


is open if and only if that same set W is also open when regarded as a subset of
1
(X/P1 )/P2 . But, W (X/P1 )/P2 is open if and only if 2 (W ) is an open subset of
1
1
X/P1 which in turn is true if and only if 1 (2 (W )) = (2 1 )1 (W ) is an open
subset of X. Since f 3 = 2 1 , the conclusion follows.

Theorem 8.1.8. Let f : X1 X2 be a homeomorphism and let i : Xi Yi ,


i = 1, 2 be two quotient maps such that there exists a injective function f : Y1 Y2 for
which the diagram
X1
1

Y1

X2


Y2

commutes. Then f is a homeomorphism between Y1 and Y2 .

Proof. The function f : Y1 Y2 is surjective since both 2 and f are surjective,


showing that f is a bijection. According to Theorem 8.1.6, f : Y1 Y2 is continuous if

and only if 2 f : X1 Y2 is continuous. The latter is a composition of two continuous


functions and therefore continuous. A symmetric argument shows that f 1 : Y2 Y1
is likewise continuous.
The next theorem points to which topological properties are inherited by quotient
spaces.
Theorem 8.1.9. Let : X Y be a quotient map. If X is either compact,
connected or path-connected, then so is Y .
Proof. This follows from the fact that Y = f (X) and that each of the properties of compactness, connectedness and path-connectedness is preserved under taking
continuous images (Theorem 6.1.8, Theorem 7.1.3 and Theorem 7.2.3 respectively).
With some of the general properties of quotient spaces out of the way, we turn
to examples. Specic groups of examples are considered in sections ?? ??. In the
present section we only look at some very basic examples to illustrate Denition 8.1.3.
We do so after rst singling out two special scenarios of surjective functions : X Y
that play a prominent role in the remainder of this section.
Construction 8.1.10 (Quotienting out by a subset). Let (X, TX ) be a topological
space and let A X be a subset of X. Let Y be the set Y = (X A) {a} where a

128

8. QUOTIENT SPACES

is some abstract element not in X. Dene the function : X Y by


(x) =

x
a

;
;

x X A,
x A,

and note that is surjective. The space (Y, TX/ ) is typically denoted by (X/A, TX/A )
and referred to as the quotient of X by A. Note that it is the quotient space X/PA
associated to the partition PA = {A, {x} | x X A} of X.
Construction 8.1.11 (Gluing of topological spaces). Let (X1 , T1 ) and (X2 , T2 ) be
two disjoint topological spaces, let A X1 be a given subspace and let f : A X2 be
a continuous function. Let X = X1 X2 equipped with the topology TX = TX1 TX2
(that is U TX if and only if U Xi TXi , i = 1, 2) and consider the partition P of
X dened by
P = {{x}, {y} f 1 (y) | x X1 A, y X2 }.
The space Y = X/P is denoted by Y = X1 f X2 and is said to have been obtained by
gluing X1 to X2 along A by using the function f : A X2 . (which we will typically
abbreviate by simply saying X1 f X2 is obtained by gluing X1 to X2 ). Note that Y
is constructed by identifying points x A to their image f (x) Y . Points x X1 A
and points y X2 f (A) do not get identied with other points.
Example 8.1.12. Let Dn = {x Rn | ||x|| 1} and S n1 = {x Rn | ||x|| = 1}
be the n-dimensional closed ball and the (n 1)-dimensional sphere, each equipped
with the relative Euclidean topology. Since S n1 Dn , it makes sense to form the
quotient space Dn /S n1 (Construction 8.1.10).
The main goal of this example is to establish the existence of a homeomorphism
f : (Dn /S n1 , TDn /S n1 ) (S n , TEu ).
To dene the function f , we rst introduce a couple of auxiliary functions. Let g1 :
Int(Dn ) Rn be the homeomorphism given by (Exercise 8.7.2)

ln 1+||x|| x
; x = 0,

1||x||
||x||
g1 (x) =

0
; x = 0,
and let g2 : Rn S n {(1, 0, ..., 0} be the homeomorphism
g2 (y1 , ..., yn ) =

2y1
2y2
2yn
||y||2 1
,
,...,
,
||y||2 + 1 ||y||2 + 1
||y||2 + 1 ||y||2 + 1

Write Dn /S n1 = (Dn S n1 ){s} where s Dn /S n1 represents the image of S n1 in


Dn /S n1 under the quotient map : Dn Dn /S n1 . Then we dene f : Dn /S n1
S n as

; x Dn S n1 ,
g2 (g1 (x))
f (x) =

(0, ..., 0, 1)
; x = s.

8.1. QUOTIENT SPACES: DEFINITIONS, PROPERTIES AND FIRST EXAMPLES

129

Clearly f is a bijection so we only need to verify that it is continuous and open. To

show that f is continuous, it suces to show that f = f : Dn Rn is continuous

(Theorem 8.1.6). From its denition, it is evident that f is continuous at all points
n
x Int(D ) so we are left with verifying that this remains true for x0 S n1 . Noting

that f (x0 ) = (0, 0, ..., 0, 1) for each x0 S n1 , let V S n be any neighborhood of

(0, .., 0, 1). Continuity of f at x0 will be demonstrated if we can nd a neighborhood U

of x0 in Dn such that f (U ) V . Firstly, nd an r > 0 such that B(0,...,0,1) (r) S n V .


Without loss of generality we can assume that r 2 for otherwise V = S n and then
any neighborhood U of x0 maps into V . But if r 2, then U = {x Dn | ||x|| > R}
with
exp

2 2
r

exp

2 2
r

1 1

R=

where

exp(t) = et ,

1 +1

is a neighborhood of x0 with f (U ) = B(0,...,0,1) (r) showing that f is continuous at x0


and consequently that f is continuous. The verication of the continuity of f 1 is left
as an exercise.
Example 8.1.13. Let X1 and X2 be two disjoint copies of the closed unit ball
D = {(x1 , ..., xn ) Rn | x2 + ... + x2 1} equipped with the relative Euclidean
n
1
topology and let A X1 be the subset A = S n1 = {(x1 , ..., xn )Rn | x2 + .... + x2 = 1}.
n
1
Dene f : A X2 be f (x) = x. We will show that X1 f X2 (Construction 8.1.11) is
homeomorphic to S n .
As usual, let : X1 X2 X1 f X2 denote the quotient map and let the function
g : X1 X2 S n be dened as

; (x1 , ..., xn ) X1 ,
(x1 , ..., xn , 1 x2 ... x2 )
1
n
g (x1 , ..., xn ) =

(x1 , ..., xn , 1 x2 ... x2 )


; (x1 , ..., xn ) X2 .
1
n
n

Note that g induces a function g : X1 f X2 S 1 since g (x) = g (f (x)) for all x X1 .

According to Theorem 8.1.6, to show that g is continuous it suces to show that g

is continuous. The continuity of g , however, is immediate and follows from Lemma

3.1.12 after observing that the subsets X1 and X2 are closed in X1 X2 . Finally, g is
a bijection since it has in inverse function, namely

; yn+1 0,
(y1 , ..., yn ) X1
1
g (y1 , ..., yn+1 ) =
(y , ..., y ) X
; yn+1 0.
1
n
2
Since S n is compact and since X1 f X2 is Hausdor (which we leave as an easy exercise), it follows from Corollary 6.1.10 that g is a homeomorphism. Figure 1 illustrates
the formation of X1 f X2 from X1 and X2 .

130

X1
X2

8. QUOTIENT SPACES

X1

X2
X1 f X2
Figure 1. An illustration of the creation of the identication space
X1 f X2 from Example 8.1.13. Each of Xi is a copy of the closed unit
ball Dn (n being 2 in this illustration) and the gluing map f : X1 X2
is the identity.

Example 8.1.14. Consider the Euclidean line (R, TEu ) and let A R be the
subset of all integers A = Z. Using the notation from Construction 8.1.10, we form
the quotient space R/Z. Visually, we think of R/Z as having been gotten by collapsing
Z R to a single point, as in Figure 2. This representation of R/Z is not unlike that
of the Hawaiian earrings H from Example 2.2.17 (also shown in Figure 2). Recall that
H is the union Cn of the circles
n=1
Cn = {(x, y) R2 | (x

1 2
1
) + y 2 = 2 },
n
n

and as a subset of R2 , H is equipped with the relative Euclidean topology.


The goal of this example is to show that in fact R/Z and H are not homeomorphic,
despite the similarities in their appearances. We rst observe that R/Z is not compact.
Namely, let
1
1
1
3
FR =
a ,a +
aZ ,
, a + ,a +
2
2
4
4
be an open cover of R and let FR/Z = {(U ) | U FR } be the associated open cover
of R/Z where : R R/Z is the quotient map. This open cover of R/Z has no nite
1
subcovers at all since every point (a + 2 ), a Z lives in a unique open set ( a +
1
, a + 3 ) while every point (a + 1 ), a Z lives in a unique open set ( a 1 , a + 1 ).
4
4
5
2
2
On the other hand, H is a compact space. Being a subspace of R2 , it suces to
show that H is bounded and closed. The boundedness of H is obvious and to verify
the closedness of H, let p R2 H be any point. Then there exists an integer n0 N
such that p lies inside of Cn for all n < n0 and p lies outside of Cn for all n n0
(if p lies outside of C1 we take n0 = 1). If n0 = 1 let r = dC (p) and if n0 > 1 let
r = min{dCn0 (p), dCn0 +1 (p)}. In either case, the ball Bp (r) lies entirely in R2 H
showing that R2 H is open and thus that H is closed.

8.1. QUOTIENT SPACES: DEFINITIONS, PROPERTIES AND FIRST EXAMPLES

2, 1

131

(x 1)2 + y 2 = 1
1, 0
1
(x 2 )2 + y 2 =

1
4

0, 1

4, 5
1, 2
3, 4

2, 3

(a)

(b)

Figure 2. (a) A visual representation of R/Z. Each of the innitely


many loops represents an open interval a, a + 1 for some integer a Z.
Several such intervals are explicitly labeled. The distinguished point z is
marked by a black dot in the middle. (b) The Hawaiian earrings. Each
1
1
of the circles corresponds to {(x, y) R2 | (x n )2 + y 2 = n2 } for some
integer n N. The distinguished point o = (0, 0) is labeled by a black
dot.
Finally, since compactness is a topological invariant (Theorem 6.1.8), R/Z and H
cannot be homeomorphic.
Example 8.1.15. Consider again the Euclidean line (R, TEu ) and let A R be the
set of rational number Q. By considering the quotient space R/Q, we will demonstrate
that the Hausdor property of a topological space may disappear when passing to a
quotient space.
To see that R/Q is not Hausdor, let x, y R be any two distinct irrational elements
in R and let (x), (y) be their images in R/Q under the quotient map : R R/Q.
Note that (x) = (y). Suppose that U(x) and U(y) are neighborhoods of (x) and
(y) in R/Q. Then Vx = 1 (U(x) ) and Vy = 1 (U(y) ) are neighborhoods of x and
y in R and as such, both contain a rational number, say qx Vx Q and qy Vy Q.
But then we can conclude that (qx ) U(x) and (qy ) U(y) and since clearly
(qx ) = (qy ), we nd that U(x) and U(y) cannot be disjoint. Accordingly, R/Q is
not Hausdor even though R clearly is.

Potrebbero piacerti anche