Вы находитесь на странице: 1из 63

GENERAL TOPOLOGY

Contents
1. Relations
1.3. Equivalence relations
1.8. Linear Orders
2. The integers and the real numbers
3. Products and coproducts
4. Finite and infinite sets
5. Countable and uncountable sets
6. Infinite sets and the Axiom of Choice (AC)
7. Well-ordered sets
8. The Maximum Principle
9. Topological spaces
10. Basis for a Topology
11. Order topologies
12. The product topology
12.6. The coproduct topology
13. The subspace topology
14. Closed sets and limit points
14.3. Closure and interior
14.10. Limit points and isolated points
14.14. Convergence, the Hausdorff property, and the T1 -axiom
15. Continuous functions
15.6. Homeomorphisms and embeddings
15.10. Maps into product spaces
15.15. Maps out of coproducts
16. Quotient maps
16.1. Open and closed maps
16.2. Quotient maps
17. Metric topologies
17.6. The first countability axiom
17.12. The uniform metric
18. Connected spaces
19. Connected subsets of linearly ordered spaces
19.5. Path connected spaces
19.9. Components and path components
19.12. Locally connected and locally path connected spaces
20. Compact spaces
21. Compact subspaces of linearly ordered spaces
21.5. Compactness in metric spaces
22. Limit point compactness and sequential compactness
23. Locally compact spaces and the Alexandroff compactification
24. Countability axioms
25. Separation Axioms
26. Normal spaces
27. Second countable regular spaces and the Urysohn metrization theorem
27.1. An embedding theorem
27.5. A universal second countable regular space
Date: 24th March 2004.
1

3
3
4
5
6
7
8
11
12
14
16
16
18
18
19
19
22
22
23
24
24
25
26
27
28
28
28
34
34
36
36
39
40
40
41
43
45
47
48
48
52
53
55
57
57
57

GENERAL TOPOLOGY

28. Completely regular spaces and the StoneCech


compactification

28.6. The StoneCech construction


29. Manifolds
References

59
59
61
63

GENERAL TOPOLOGY

1. Relations
Definition 1.1. A relation R on the set A is a subset R A A.
We shall be interested in these properties of relations
Re: (Reflexivity) aRa for all a A
Sy: (Symmetry) aRb bRa for all a, b A
Tr: (Transitivity) aRb and bRc aRc for all a, b, c A
Com: (Comparability) For all a, b A, we have either a = b, aRb, or bRa
NonRe: (Non-reflexivity) aRa for no a A.
where aRb means (a, b) R A B.
Example 1.2. We may define a relation D on Z+ by aDb if a divides b. The relation D Z+ Z+
is reflexive and transitive. It has none of the other properties.
1.3. Equivalence relations.
Definition 1.4. An equivalence relation on a set A is a relation that is reflexive, symmetric, and
transitive.
This means that
Re: a a for all a A
Sy: a b b a for all a, b A
Tr: a b c a c for all a, b, c A
when is an equivalence relation.
The equivalence class containing a A is the subset
[a] = {b A | a b}
of all elements of A that are equivalent to a. The collection of equivalence classes is denoted A/ .
The canonical map [ ] : A A/ takes the element a A to the equivalence class [a] A.
Example 1.5. (1) Equality is an equivalence relation. The equivalence class [a] = {a} contains
just one element.
(2) a mod b mod n is an equivalence relation on Z. The equivalence class [a] = a + nZ consists of
all integers congruent to a mod n and the set of equivalence classes is Z/nZ = {[0], [1], . . . , [n 1]}.
def

(3) x y |x| = |y| is an equivalence relation in the plane R2 . The equivalence class [x] is a
circle centered at the origin and R2 / is the collection of all circles centered at the origin.
def

(4) If f : A B is any function, a1 a2 f (a1 ) = f (a2 ) is an equivalence relation on A. The


equivalence class [a] = f 1 (f (a)) A is the fibre over f (a) B. we write A/f for the set of
equivalence classes.
(5) [Ex 3.2] (Restriction) Let X be a set and A X a subset. Declare any two elements of A to
be equivalent and any element outside A to be equivalent only to itself. This is an equivalence
relation. The equivalence classes are A and {x} for x X A. One writes X/A for the set of
equivalence classes.
(6) [Ex 3.5] (Equivalence relation generated by a relation) The intersection of any family of equivalence relations is an equivalence relation. The intersection of all equivalence relations containing
a given relation R is called the equivalence relation generated by R.
Lemma 1.6. Let be an equivalence relation on a set A. Then
(1) a [a]
(2) [a] = [b] a b
(3) If [a] [b] 6= then [a] = [b]
Proof. (1) is reflexivity, (2) is symmetry, (3) is transitivity: If c [a] [b], then a c b so a b
and [a] = [b] by (2).

This lemma implies that the collection A/ is a partition of A, a collection of nonempty, disjoint
subsets of A whose union is all of A. Conversely, given any partition of A we define an equivalence
relation by declaring a and b to be equivalent if they lie in the same subset of the partition. We
conclude that an equivalence relation is essentially the same thing as a partition.

GENERAL TOPOLOGY

The canonical map [ ] : A A/ has the following universal property: Any map f : A B
that respects the equivalence relation in the sense that a1 a2 f (a1 ) = f (a2 ) factors uniquely
through A/ . This means that there is a unique map f such that the diagram
/B
AA
AA
}>
}
AA
}
AA
}}
}
[ ]
!f
}
A/
f

commutes.
Example 1.7. Any map f : A B is the composite A

[ ]

/ / A/f / f

/ B . The corestriction

f : A/f f (A) of f is a bijection between the set of fibres A/f and the image f (A).
1.8. Linear Orders.
Definition 1.9. A linear order on the set A is a relation that satisfies comparability, nonreflexivity, and transitivity.
This means that
Com: If a 6= b then a > b or b > a for all a, b A
NonRe: a > a for no a A
Tr: a > b > c a > c for all a, b, c A
when > is a linear order.
Example 1.10. The usual x > y is a linear order on R.
The open interval from a to b > a is the set
(a, b) = {x A|a < x < b}
If (a, b) = then a is the immediate predecessor of b, and b the immediate successor of a.
Definition 1.11. Let (A, <) and (B, <) be linearly ordered sets. An order preserving map is a
map f : A B such that a1 < a2 f (a1 ) < f (a2 ) for all a1 , a2 A.
An order preserving map is always injective. If there exists an order preserving bijection
f : A B, then we say that (A, <) and (B, <) have the same order type.
Definition 1.12. Let (A, <) and (B, <) be linearly ordered sets. The dictionary order on A B
is the linear order given by
def

(a1 , b1 ) < (a2 , b2 ) (a1 < a2 ) or (a1 = a2 and b1 < b2 )


What about orders on A B?
Any subset of a linearly ordered set inherits a linear order. The restriction of a dictionary order
to a product subspace is the dictionary order of the restricted linear orders.
Example 1.13. (1) R and (0, 1) have the same order type. [0, 1) amd (0, 1) have distinct order
types. {1} (0, 1) and [0, 1) have the same order type.
(2) The dictionary order on Z+ [0, 1) has the same order type as [1, ). In the dictionary order
on [0, 1) Z+ each element (a, n) has (a, n + 1) as its immediate successor. Thus Z+ [0, 1) and
[0, 1) Z+ do not have the same order type.
The example shows that, in general, (A, <) (B, <) 6= (B, <) (A, <) as ordered sets.
Definition 1.14. An ordered set (A, <) has the least upper bound property if any subset that
is bounded from above has a least upper bound. It is well-ordered if any nonempty subset has a
smallest element.
Example 1.15. R has the least upper bound property. Z+ is well-ordered. The order type of the
well-ordered set (Z+ , <) is denoted . Any well-orderes set has the least upper bound property.

GENERAL TOPOLOGY

2. The integers and the real numbers


Theorem 2.1. (Induction Principle) Let A Z+ . If 1 A and n A n + 1 A for every
n Z+ , then A = Z+ .
Proof. The hypothesis is that A is inductive. By definition, Z+ is the smallest inductive subset of
R, so Z+ A.

Theorem 2.2. (Well-ordering) Z+ is well-ordered.
Proof. The claim is that any non-empty subset of Z+ has a smallest element.
As a first step we show by induction that any nonempty subset of Sn has a smallest element.
We claim that the subset
(2.3)

{n Z+ | Sn is well-ordered}

in inductive and therefore equal to Z+ . Clearly, 1 is in (2.3). Suppose that the integer n belongs
to (2.3). We must show that n + 1 is in (2.3). Let therefore B be any nonempty subset of Sn+1 . If
B Sn+1 6= then the smallest element of B Sn+1 is also the smallest element of B. If B Sn = ,
then B = {n} and n is the smallest element of B. This shows that B has a smallest element, i.e.
that n + 1 is in (2.3).
Let now A Z+ be any nonempty subset. The intersection A Sn is nonempty for some n, so
it has a smallest element. This is also the mallets element of A.

Theorem 2.4. (General Induction Principle) Let A be a subset of Z+ . If the implication
Sn A n A
is true for all n Z+ , then A = Z+ .
Proof. Suppose that A is not all of Z+ . Let n = min(Z+ A) be the smallest element outside
A (well-ordering), i.e. x 6 A x n or, equivalently, Sn A. Then n A by hypothesis.
Contradiction.

Theorem 2.5. (Archimedean Principle) Z+ has no upper bound in R.
Proof. We assume the opposite and derive a contradiction. Suppose that Z+ is bounded from
above. Let b = sup Z+ be the smallest upper bound (well-ordering). Since b 1 is not an upper
bound, there is a positive integer n Z+ such that n > b 1. Then n + 1 is also an integer (Z+
is inductive) and n + 1 > b contradicting that b is an upper bound for Z+ .

The Archimedean Principle says that for any real number x there is a positive integer n greater
than x.

GENERAL TOPOLOGY

3. Products and coproducts


Definition 3.1. An indexed family of sets consists of a collection A of sets, an index set J, and
a surjective function f : J A.
We often denote the set f (j) by Aj and the whole indexed family by {Aj }jJ . Any collection
A can be made into an indexed family by using the identity map A A as the indexing function.
We define the union and the intersection of the indexed family as
[
\
Aj = {a | a Aj for at least one j J},
Aj = {a | a Aj for all j J}
jj

jj

and we define the Cartesian product to be


Y
S
(3.2)
Aj = {x : J Aj | j J : x(j) Aj }
jJ

Observe that the set of maps into a product


Y
Y
map(B,
Aj ) =
map(B, Aj )
jJ

jJ

Q
is a product. In particular there is a map : jJ Aj jJ Aj corresponding to the inclusion
T
Q
maps jJ Aj Aj , j J. The product comes equipped with projection maps j :
Aj Aj ,
j J, inducing the above correspondence.
If the index set J = Sn+1 = {1, . . . , n} then we also write
A1 An , A1 An , A1 An
Q
for jSn+1 Aj , jSn+1 Aj , jSn+1 Aj , respectively. If also and Aj = A for all j Sn+1 we
Q
write An for the product jSn+1 A. The elements of An are all n-tuples (a1 , . . . , an ) of elements
from A.
If the index set J = Z+ then we also write
S

A1 An , A1 An , A1 An
T
Q
for jZ+ Aj , jZ+ Aj , jZ+ Aj , respectively. If also Aj = A for all j we write A for the
Q
product jZ+ A, the set of all functions x : Z+ A, i.e. all sequences (x1 , . . . , xn , . . .) of elements
from A. (The reason for this notation is that the order type of (Z+ , <) is often denoted by .)
S
Example 3.3. (1) S1 S2 Sn = nZ+ Sn = Z+ .
Q
(2) If the collection A = {A} consists of just one set A then the product jJ A = {x : J A}
is the set of all functions from J to A.
(3) There is a bijection (which one?) between {0, 1} = {x : Z+Q
{0, 1}} and P(Z+ ). More
generally, there is a bijection (which one) between the product jJ {0, 1} and the power set
P(J).
S

We define the copoduct to be the set


a
[
(3.4)
Aj = {(j, a) J Aj | a Aj }
jJ

`
S
The coproduct is sometimes also called
the disjoint union because
Aj = A0j is the union of
S
the disjoint sets A0j = {j} Aj J jJ Aj . Observe that the set of maps out of the coproduct
a
Y
map(
Aj , B) =
map(Aj , B)
jJ

jJ

S
is a product. In particular there is a map : jJ Aj jJ Aj corresponding to the inclusion
S
`
maps Aj jJ Aj . The coproduct comes equipped with inclusion maps j : Aj Aj , j J,
inducing the above correspondence.
If the index set J = Sn+1 = {1, . . . , n} then we also write A1 q q An for the coproduct.

GENERAL TOPOLOGY

4. Finite and infinite sets


Definition 4.1. A set A is finite it admits a bijection f : A Sn to some section Sn , n Z+ , of
Z+ . A set is infinite if it is not finite.
Lemma 4.2. Let A be a set, a0 an element of A, and n Z+ a positive integer. Then
There is a bijection A Sn+1 There is a bijection A {a0 } Sn
Proof. : Let f0 : A {a0 } Sn be a bijection. Define f : A Sn+1 by
(
f0 (a) a 6= a0
f (a) =
n + 1 a = a0
: Let f : A Sn+1 be a bijection. If we are so lucky that f (a0 ) = n + 1, we just take the
restriction of f to A {a0 }. If we are not lucky, let g = (f (a0 ), n + 1) : Sn+1 Sn+1 be the
permutation that interchanges f (a0 ) and n + 1 and leaves all other elements fixed. The bijection
g f between A and Sn+1 takes a0 to n + 1, so we are back in the lucky situation.

Could it happen that there A admits bijections Sm+1 o
where m 6= n?

/ Sn+1 to Sm+1 and Sn+1

Corollary 4.3. Let m Z+ and n Z+ be two positive integers. If m 6= n, then there is no


bijection between Sm and Sn .
Proof. We have just seen that
There is a bijection Sm+1 Sn+1 There is a bijection Sm Sn
or, equivalently,
There is no bijection Sm Sn There is no bijection Sm+1 Sn+1
This is precisely (the nontrivial part of) the statement that
{n Z+ | There is no bijection Sm Sn for m 6= n}
is inductive.

This means that a finite set is in bijection with Sn+1 for precisely one positive integer n Z+ ;
this integer is the cardinality of A, denoted card A = n.
Corollary 4.4. Let A be a finite set. Any subset B of A is also finite and
B A card B card A
B ( A card B < card A
Proof. It suffices to consider subsets of A = Sn+1 . Using Lemma 4.2 we see that the sets
{n Z+ | Any B Sn+1 is in bijection with Sm+1 for some m n}
{n Z+ | Any B ( Sn+1 is in bijection with Sm+1 for some m < n}
are inductive. Lets for instance look at the first set. Assume that n belongs to this set. We want
to show that n + 1 belongs to the set. So, let B be a subset of Sn+2 . If B does not contain n + 1,
then B is a subset of Sn+1 and then we know from the induction hypothesis that B is in bijection
with Sm+1 for some m n. If B does contain n + 1, then B {n + 1}, a subset of Sn+1 , is in
bijection with Sm+1 for some m n. Lemma 4.2 says that B is in bijection with Sm+2 where
m + 1 n + 1.

In particular, a finite set A can not be in bijection with any of its proper subsets B ( A (for
their cardinalities are distinct). We can use this property of finite sets to find an example of an
infinite set.
Corollary 4.5. Z+ is infinite.
Proof. The map x x + 1, with inverse function x x 1, is a bijection between Z+ and its
proper subset Z+ {1}.


GENERAL TOPOLOGY

Theorem 4.6 (Characterization of finite sets). Let A be a set. The following statements are
equivalent
(1) A is finite
(2) There exists a surjection Sn+1 A for some n Z+
(3) There exists an injection A Sn+1 for some n Z+
Proof. (1) (2): There even exists a bijection Sn+1 A.
(2) (3): Let f : Sn+1 A be surjective. If n = 0, A = and there is nothing to prove. When
n > 0, we construct a right inverse Sn+1 o

f
g

o / / A to f . The fibres f 1 (a) are nonempty for all

a A so they have smallest elements since Z+ is well-ordered. Define g : A Sn+1 by


g(a) = min f 1 (a)
As a right inverse to f , g is injective.
(3) (1): If there exists an injection A Sn+1 , then there exists a bijection between A and a
subset of Sn+1 . All subsets of the finite set Sn+1 are finite (4.4).

h
x
/ /B .
Remark 4.7. You may get the idea that any surjective map admits a right inverse: A
Can you prove it? Consider for instance the surjective map
proj

{(b, B) A P 0 (A) | b B} P 0 (A) : (b, B) 7 B


where P 0 (A) = P(A) {} is the collection of nonempty subsets of A. The dual statement is that
xx
/ B . Can you prove that?
any injection admits a left inverse: A /
Corollary 4.8.
(1) Subsets of finite sets are finite.
(2) Images of finite sets are finite.
(3) Finite unions of finite sets are finite.
(4) Finite Cartesian products of finite sets are finite.
Proof. (2) Sn+1  A  f (A).
(3) We show first that AB is finite if A and B are finite. Suppose f : Sn+1 A and g : Sm+1 B
are surjective functions. Then h : Sm+n+1 A B given by
(
f (i) i = 1, 2, . . . , n
h(i) =
g(i) i = n + 1, . . . , m + n
is also surjective. Induction now shows that A1 An is finite when A1 , . . . , An are finitely
many finite sets.
(4) Since
[
AB =
{a} B
aA

is finite union of finite sets, it is finite. Induction now shows that A1 An is finite when
A1 , . . . , An are finitely many finite sets.

5. Countable and uncountable sets
Definition 5.1. A set A is countably infinite if it admits a bijection f : A Z+ . It is countable
if it is finite or countably infinite. It is uncountable if it is not countable.
Example 5.2. Z and Z+ Z+ are countably infinite.
Lemma 5.3. Any infinite subset of Z+ is countably infinite.
Let C Z+ be an infinite subset. The claim is that the elements of C can be listed C =
{c1 , c2 , . . .} as an infinite sequence. The obvious thing to do is let c1 be the smallest element of C,
c2 the next element of C, and so on. However, to do this correctly we must use the
Principle of recursive definition Given a set A, an element a1 of A, and a formula that for each
i > 1 defines h(i) in terms of h(1), . . . , h(i 1). Then this formula defines a function h : Z+ A
such that h(1) = a1 .
This follows from the Induction Principle, but we shall not go into details. Recursive functions
can be computed by a Turing machine.

GENERAL TOPOLOGY

Proof of Lemma 5.3. Let C Z+ be an infinite set of positive integers. We define a function
h : Z+ C recursively by
h(1) = min C
h(i) = min(C {h(1), . . . , h(i 1)}),

i>1

using that Z+ is well-ordered. Note that C {h(1), . . . , h(i 1)} is nonempty since C is infinite
(4.6). We claim that h is bijective.
h is injective: The function h is injective (even order preserving) because the descending chain
C ) C {h(1)} ) ) C {h(1), . . . , h(i 1)} ) C {h(1), . . . , h(i 1), h(i)} )
implies that h(1) < h(2) < h(i) < h(i + 1) < .
h is surjective: Let c C. We must find a positive integer m such that c = h(m). Our only hope
is
m = min{n Z+ | h(n) c}
(Note that this has a meaning since the set {n Z+ | h(n) c} is nonempty as we can not
inject the infinite set Z+ into the finite set {1, . . . , c 1} = Sc (4.4). Note also that again we use
well-ordering of Z+ .) By definition of m,
h(m) c and h(n) c n m
The last of these two properties is equivalent to n < m h(n) < c, so c 6 {h(1), . . . , h(m 1)},
or c C {h(1), . . . , h(m 1)}, and therefore
h(m) = min(C {h(1), . . . , h(m 1)}) c
by definition of h. Thus h(m) = c.

Theorem 5.4 (Characterization of countable sets). Let A be a set. The following statements are
equivalent
(1) A is countable
(2) There exists a surjection Z+ A
(3) There exists an injection A Z+
Proof. If A is finite, the theorem is true, so we assume that A is countably infinite.
(1) (2): Clear.
(2) (3): If f : Z+ A is s surjection, then g : A Z+ given by g(a) = min f 1 (a) is a leftinverse to f , hence injective.
(3) (1): Let g : A Z+ be injective. The infinite subset g(A) of Z+ is countable by Lemma 5.3
and it is in bijection with A since g is injective.

Corollary 5.5.
(1) A subset of a countable set is countable
(2) The image of a countable set is countable.
(3) A countable union of countable sets is countable.
(4) A finite product of countable sets is countable.
Proof. (1) B  A  Z+ .
(2) Z+  A  B.
(3) Let {Aj }jJ be an indexed collection of sets where J is countable and each set Aj is countable.
Choose surjective maps f : Z+ J and gj : Z+ Aj . Then the map
[
[
Z+ Z+ =
{m} Z+ 
Aj : (m, n) 7 gf (m) (n)
mZ+

jJ

is surjective. Since Z+ Z+ is countable (5.6) so is its image (5.5).


(4) The product of two countable sets is countable since the product of two surjective maps is
surjective. Now use induction.

Example 5.6. The map f : Z+ Z+ Z+ given by f (m, n) = 2m 3n is injective by uniqueness
of prime factorizations. The map g : Z+ Z+ Q+ given by g(m, n) = m
n , is surjective. Thus
Z+ Z+ and Q+ are countable. Q = Q {0} Q+ is countable.
You may think that a countable product of countable sets is countable but thats false.

10

GENERAL TOPOLOGY

Lemma 5.7. The countable product {0, 1} =


uncountable.

nZ+ {0, 1}

= {x : Z+ {0, 1}} of finite sets is

Proof. (Cantors diagonal argument) Let g : Z+ {0, 1} be any function. We show that g can
not be surjective. Define an element (yn )nZ+ {0, 1} by
(
0 g(n)n = 1
yn =
1 g(n)n = 0
Then (yn ) 6= g(m) for all m Z+ for ym 6= g(m)m .

The power set P(Z+ ) is in bijection with {0, 1} (3.3.(3)). Thus the countable set Z+ has an
uncountable power set P(Z+ ). It is true in general that card A < card P(A) for any set A.
Theorem 5.8. Let A be any set.
(1) There is no injective map P(A) A
(2) There is no surjective map A P(A)
/ A , then there will also exist a surjection (4.7) in
Proof. If there exists an injection P(A) /
ss
/ A . Thus it suffices to show (2). We assume the converse: That
the other direction: P(A) /
g : A P(A) is surjective. The subset
{a A | a 6 g(a)}
is of the form g(a0 ) for some a0 A. This leads to the contradiction
a0 g(a0 ) a0 6 g(a0 )

Conjecture 5.9 (Cantors Continuum Hypothesis). Suppose that X R is an uncountable set.
Then there exists a bijection : X R.
Look up the (generalized) continuum hypothesis [Ex 11.8] somewhere [11, 12].

GENERAL TOPOLOGY

11

6. Infinite sets and the Axiom of Choice (AC)


Recall that a set A is infinite if it does not admit a bijection A Sn with some section of the
positive integers.
Theorem 6.1 (Characterization of infinite sets). Let A be a set. The following are equivalent:
(1) A is infinite
(2) There exists an injective map Z+ A
(3) A is in bijection with a proper subset of itself
The naive student will prove (1) (2) like this: Suppose A is infinite. Let a1 be any element
of A. Define h : Z+ A recursively as follows:
h(1) = a1
h(i) = any element of A {h(1), . . . , h(i 1)}
Then h is injective.
The problem here is that there is no formula for h(i) in terms of h(j) for j < i. In fact, for the
proof of this innocent looking theorem we need the
Axiom 6.2 (Axiom of Choice (AC)). Given any nonempty collection of nonempty disjoint sets,
there exists a set C containing exactly one element from each of the sets in the collection.
Unlike the other axioms os set theory, the AC does not determine C uniquely. In 1938 K. Godel
showed that if the axioms of set theory are consistent (we dont know if they are), then set theory
with AC is also consistent. The reason could be that AC is a consequence of the other axioms.
However, P.J. Cohen showed in 1963 that this is not the case.
Look at (4.7) again.
Here is a reformulation of AC.
Theorem 6.3. The following statements are equivalent:
(1) The Axiom of Choice
(2) For any collection A of S
nonempty sets (not necessarily disjoint) there exists a map (a
choice function) c : A Q AA A such that c(A) A for all A A.
(3) The Cartesian product AA A of any nonempty collection A of nonempty sets (not necessarily disjoint) is nonempty.
Proof. It is clear that (2) and (3) are equivalent by the very definition of Cartesian products.
(1) = (2): Let (A0 )AA be the indexed family of sets
[
A0 = {(A, a) | a A A} A
A
AA
0

Then (A )AA is a collection of disjoint sets. (Let A1 be two distinct sets in the collection A. The
sets A01 and A02 are disjoint for the elements, (A1 , a), a A1 , of A01 have first coordinate A1 while
the elements, (A2 , a), a A2 , of A02 have first coordinate A2 6= A1 .) By the Axion of Choice there
contains a single element for each A0 . If we write C A0 = {(A, c(A))}
is set C such that C A0 S
then c is a function A S AA A.
(2) = (1): Let c : A AA A be a choice function for A, a nonempty collection of nonempty
disjoint sets. The set we seek is the image C = c(A) of c: Since c(A) A and the sets in A are
disjoint, c(A) does not belong to any other of the sets in the collection, so C A = {c(A)} for all
A A.

Here is a special, but often used, case. Let A be any set and P 0 (A) = P(A) {} the collection
of nonempty subsets of A. Then there exists (6.3.(2)) a choice function c : P 0 (A) A such that
c(B) B for any nonempty B A. (The choice function picks out an element in each nonempty
subset of A.)
Proof of Theorem 6.1. (1) = (2): Let P 0 (A) be the collection of nonempty subsets of A and
c : P 0 (A) A a choice function (c(B) B for each nonempty subset B of A.) Let h(1) be any

12

GENERAL TOPOLOGY

element of A and h : Z+ A the recursively defined function given by


h(i) = c(A {h(1), . . . , h(i 1)})
Then h is injective (if i < j then h(j) A {h(1), . . . , h(i), . . . , h(j 1)} so h(i) 6= h(j)).
(2) = (3): We view Z+ as a subset of A. Then A = (A Z+ ) Z+ is in bijection with the proper
subset A {1} = (A Z+ ) (Z+ {1}).
(3) = (1): We already know that a finite set can not be in bijection with any of its proper subsets
for cardinality reasons.

7. Well-ordered sets
A very convenient property of (Z+ , <) is that any nonempty subset has a smallest element. We
now focus on this property.
Definition 7.1. A set A with a linear order < is well-ordered if any nonempty subset has a
smallest element.
Any well-ordered set has a smallest element. Any element (but the largest) in a well-ordered set
has an immediate successor, the smallest successor [Ex 10.2]. A well-ordered set can not contain
an infinite descending chain x1 > x2 > , in fact, a linearly ordered set is well-ordered if and
only if it does not contain a copy of the negative integers Z [Ex 10.4].
Definition 7.2. Let (A, <) be a well-ordered set and an element of A. The subset
S (A) = S = [min A, ) = {a A | a < }
is called the section of A by .
Theorem 7.3 (Transfinite Induction). [Ex 10.7] Let (A, <) be a well-ordered set and J A an
inductive subset, that is
A : S J = J
Then J = A.
Proof. By contradiction. Assume that the inductive set J does not completely fill up A, i.e. that
the complement A J is nonempty. The smallest element of A J belongs to (A J) J: Let
be the smallest element not in J. Then A J by definition of , and J since S J and
J is inductive. This is absurd.

Let A o

AB

/ B be the projection maps.

Proposition 7.4.
(1) A subset of a well-ordered set is well-ordered.
(2) A finite product of well-ordered sets is well-ordered.
(3) The union of a well-ordered family of disjoint well-ordered sets is well-ordered [Ex 10.8]
Proof. (1) Clear.
(2) Let (A, <) and (B, <) be well-ordered and C a nonempty subset of A B. Let a0 = min 1 (C)
be the smallest first coordinate of any point in C. Let b0 = min 2 ({a0 } B) C) be the smallest
second coordinate above a0 . Then (a0 , b0 ) = min C is the smallest element in C. This shows that
the product of two well-ordered sets is
(3) See [Ex 10.8].

Example 7.5. (1) The positive integers (Z+ , <) is a well-ordered set.
(2) Z and R are not well-ordered sets.
(3) Any section Sn+1 = {1, 2, . . . , n} of Z+ is well-ordered.
(4) Sn+1 = {1, 2, . . . , n} of Z+ is well-ordered. The product Sn+1 Z+ is well-ordered. The finite
products Zn+ = Z+ Z+ Z+ are well-ordered.
(5) The infinite product {0, 1} is not well-ordered for it contains the infinite descending chain
(1, 0, 0, 0, . . .) > (0, 1, 0, 0, . . .) > (0, 0, 1, 0, . . .) > .

GENERAL TOPOLOGY

13

(6) The element = 2 1 Z+ Z+ is the smallest element of {1, 2} Z+ such that the section
by this element is countably infinite. The well-ordered set S = [1 1, ] has as its largest
element, the section S by is countably infinite but every other section of S is finite. Any
finite subset of S has an upper bound because the set of not upper bounds is finite (a finite
union of finite sets) while S itself is countably infinite.
Which of these well-ordereds have the same order type [Ex 10.3]? Draw pictures of examples of
well-ordered sets.
Any finite ordered set is well-ordered. To see this we first prove
Lemma 7.6. Any finite linearly ordered set has a maximal element.
Proof. The proof is by induction over the cardinality of the set. Any ordered set (A, <) of cardinality 1 has a largest element. (What is it?) Suppose that any finite ordered linearly set of
cardinality n has a largest element. Let (A, <) be a finite ordered set of cardinality n + 1. Let a
be any element of A. Then max{a, max(A {a})} is the largest element of A. (This is a kind of
sorting algorithm.)

Theorem 7.7 (Finite order types). [Ex 6.4] Any finite linearly ordered set of cardinality n has
the order type of (Sn+1 , <); in particular, it is well-ordered.
Proof. Again we use induction over the cardinality of the ordered set (A, <). Suppose that we
know that all finite ordered sets of cardinality n satisfy the theorem and let (A, <) be a finite
ordered set of cardinality n + 1. Let a be the largest element of A (7.6). By induction hypothesis,
A {a} has the order type of Sn . Then A has the order type of Sn+1 .

Can you find an explicit order preserving bijection Sm+1 Sn+1 Sm+n+1 ?
So there is just one order type of a given finite cardinality n. There are many countably infinite
well-ordered sets (7.5). Is there an uncountable well-ordered set? Our examples, R and {0, 1} , of
uncountable sets are not well-ordered (7.5.(2), 7.5.(5)).
Theorem 7.8 (Well-ordering theorem). (Zermelo 1904) Any set can be well-ordered.
We focus on the minimal criminal, the minimal uncountable well-ordered set.
Lemma 7.9. There exists a well-ordered set S with the following properties:
(1) S has a largest element
(2) The section S by is uncountable but every other section is countable.
(3) Any countable subset of S has an upper bound (cf 7.5.( 6)).
Proof. (Cf 7.5.(6)) Take any uncountable well-ordered set A. Append a greatest element to A. Call
the result A again. Now A has at least one uncountable section. Let be the smallest element of A
such that the section by this element is uncountable, that is = min{ A | S is uncountable}.
Put S = [a0 1, ] where a0 is the smallest element of A. This well-ordered set satisfies (1) and
(2). Let C be a countable subset of S . We want to show that it has an upper bound. We consider
the set of elements of S that are not upper bounds, i.e
countable

uncountable
z[}| {
z}|{
{x S | x is not an upper bound for C} = {x S | c C : x < c} =
Sc (
S
cC

This set of not upper bounds is countable for it is a countable union of countable sets. But S is
uncountable, so the set of not upper bounds is a proper subset.

See [SupplExI : 8] for an explicit construction of a well-ordered uncountable set. Z+ is the
well-ordered set of all finite (nonzero) order types and S is the well-ordered set of all countable
(nonzero) order types. See [Ex 10.6] for further properties of S .
Recall that the ordered set Z+ [0, 1) is a linear continuum of the same order type as [1, ) R.
What happens if we replace Z+ by S [Ex 24.6, 24.12]?

14

GENERAL TOPOLOGY

8. The Maximum Principle


The Axiom of Choice has several reformulations [3, Thm B.18].
If we do not insist on comparability in our order relation we obtain:
Definition 8.1. A strict partial order on a set A is a relation (precedes) on A which is nonreflexive and transitive:
(1) a a holds for no a A
(2) a b and b c implies a c.
We do not require that any two nonidentical element can be compared. For instance, power sets
are strictly partially ordered by proper inclusion (.
Theorem 8.2 (Hausdorffs Maximum Principle). Let A be a set with a strict partial order. Then
there exists a maximal linearly ordered subset B of A.
Proof. As a special case, suppose that A is infinitely countable. Then there is some bijection
Z+ A meaning that we can write A = {ai | i Z+ }. Define h : Z+ {0, 1} by recursion as
follows:
h(1) = 0
(
0
h(i) =
1

The subset {aj | j < i, h(j) = 0} {ai } is linearly ordered


otherwise

Then the set B = {ai | h(i) = 0} is linearly ordered and maximal.


The general case is formally the same as above. Let I be some set in bijection with A so that
A = {ai | i I}. Let < be a well-ordering of I (7.8). Let i0 denote the smallest element of I.
Define h : I {0, 1} by transfinite recursion as follows
h(i0 ) = 0
(
0 The subset {aj | j < i, h(j) = 0} {ai } is linearly ordered
h(i) =
1 otherwise
Then the set B = {ai | h(i) = 0} is linearly ordered and maximal.

Definition 8.3. Let (A, ) be a set with a strict partial order. An element m of A is a maximal
if no elements of A are greater than m.
Let B be a subset of A. If b  c for all b B, then c is an upper bound on B.
Maximal elements are often of great interest.
Theorem 8.4 (Zorns lemma). A partially strictly ordered set in which every linearly ordered
subset has an upper bound contains maximal elements.
Proof. Any upper bound on a maximal linearly ordered subset is maximal: Let B be a maximal
linearly ordered subset. By hypothesis, B has an upper bound m, i.e. b  m for all m B.
Suppose that m d for some d A. Then b  m d for all elements of B so B {d} is linearly
ordered, contradicting maximality of B.

We shall later use Zorns lemma to prove Tychonoffs theorem (20.17) that a product of compact
spaces is compact. In fact, The Axiom of Choice, Zermelos Well-ordering theorem, Hausdorffs
maximum principle, Zorns lemma, and Tychonoffs theorem are equivalent.
Here are two typical applications.
A basis for a vector space over a field is a maximal independent subset.
Theorem 8.5. [Ex 11.8] Any vector space has a basis, in fact any linearly independent subset of
a vector space is contained in a basis.
Proof. Let M V be a linearly independent subset. Apply Zorns lemma to the set, strictly
partially ordered by proper inclusion (, of independent subsets containing M . A maximal element
is a basis.


GENERAL TOPOLOGY

15

As a corollary we see that R and R2 are isomorphic as vector spaces over Q.


A maximal ideal in a ring is a maximal proper ideal.
Theorem 8.6. Any nontrivial ring has a maximal ideal, in fact any proper ideal is contained in
a maximal ideal.
Proof. Let I be a proper ideal. Consider the set, strictly partially ordered by proper inclusion, of
proper ideals containing the given ideal I.

Other authors prefer to work with partial orders instead of strict partial orders.
Definition 8.7. [Ex 11.2] Let A be a set. A relation  on A is said to be a partial order precisely
when it is symmetric (that is a  a for all a in A), transitive (that is a  b and b  c implies
a  c), and anti-symmetric (that is a  b and b  a implies a = b). anti-symmetric.

16

GENERAL TOPOLOGY

9. Topological spaces
What does it mean that a map f : X Y between two sets is continuous? To answer this
question, and much more, we equip sets with topologies.
Definition 9.1. Let X be a set. A topology on X is a collection T of subsets of X, called open
sets, such that
(1) and X are open
(2) The intersection of finitely many (two) open sets is open
(3) The union of an arbitrary collection of open sets is open
A topological space is a set X together with a topology T on X.
Example 9.2. (1) The indiscrete topology, or trivial topology, T = {, X}, has only two open
sets.
(2) The discrete topology T = P(X). All subsets of X are open.
(3) The Sierpinski space is the set X = {0, 1} with the topology T = {, {0}, X}. The open sets
are and the subsets containing the point 0.
(4) The standard topology on the real line R is T = {unions of open intervals}.
(5) More generally, suppose that (X, d) is a metric space. The open r-ball centered at x X is
the set Bd (x, r) = {y X | d(x, y) < r} of points within distance r > 0 from x. The metric
topology on X is the collection Td = {unions of open balls}. The open sets of the topological
space (X, Td ) and the open sets in the metric space (X, d) are the same. (See 2021 for more
on metric topologies.)
(6) The finite complement topology, or cofinite topology, T = {U X | X U is finite} {}.
The open sets are and the cofinite subsets.
The Sierpinski topology and the finite complement topology on an infinite set are not metric
topologies.
Topologies are partially ordered by inclusion. For instance, the finite complement topology
(9.2.(6)) on R is contained in the standard topology (9.2.(4), and the indiscrete topology (9.2.(1))
on {0, 1} is contained in the Sierpinski topology (9.2.(3)) is contained in the discrete topology
(9.2.(2)).
Definition 9.3 (Comparison of topologies). Let T and T 0 be two topologies on the same set X.
)
T is finer than T 0
def
T T0
T 0 is coarser than T
The finest topology is the topology with the most opens sets, the coarsest topology is the one
with fewest open sets. (Think of sandpaper!) The discrete topology is finer and the indiscrete
topology coarser than any other topology: P(X) T {, X}. Of course, two topologies may
also be incomparable.
10. Basis for a Topology
Any intersection of topologies is a topology [Ex 13.4]. Thus we may (and shall shortly) speak
of the coarsest topology containing some given collection of subsets of X. In praxis, we restrict
ourselves to collections of subsets that cover X.
Definition 10.1.
A subbasis is a collection S of subsets of X, called subbasis sets, that cover X,
S
that is X = S is the union of the subbasis sets. The topology
TS = {unions of finite intersections of subbasis sets}
is called the topology generated by the subbasis S.
Verify that TS is indeed a topology and that it is the coarsest topology in which the subbasis
sets are open [Ex 13.5]. (Use the distributive laws [1] for and .)
This becomes particularly nice when the subbasis is a basis (as in 9.2.(4) and 9.2.(3)).
Definition 10.2. A collection B of subsets of X, called basis sets, is a basis if
(1) X is a union of basis sets.

GENERAL TOPOLOGY

17

(2) The intersection of two basis sets is a union of basis sets.


The topology
TB = {unions of basis sets}
is called the topology generated by the basis B.
TB is the coarsest topology in which the basis sets are open. When we have a basis rather then
a subbasis we do not need first to take finite intersections as these are already contained in the
basis.
A topology is a basis is a subbasis. Conversely, if S is a subbasis, then
BS = {Finite intersections of subbasis sets}

(10.3)

is a basis for the topology TS generated by the subbasis S.


Lemma 10.4 (Comparison). Let B and B 0 be two bases and S a subbasis.
(1)
TB TB0 B TB0 All B-sets are open in TB0
(2)
(
)
All B-sets are open in TB0
TB = TB0
All B 0 -sets are open in TB
(3)

TB = TS

All basis sets are open in TS

Finite intersections of subbasis sets are open in TB

Proof. (1) is obvious since TB is the coarsest topology containing B. Item (2) is immediate from
(1). Item (3) is immediate from (2) with B 0 = BS , the basis generated by the subbasis S.

Two topologies, bases or subbases are said to be equivalent if they generate the same topology.
Given a topology T , letting B 0 = T in 10.4.(2) gives a criterion that B is a basis for T (that B and
T are equivalent bases); letting B = T in 10.4.(3) gives a criterion that S is a subbasis for T (that
S and T are equivalent subbases).
Example 10.5. (1) In a metric space, the set B = {B(x, r) | x X, r > 0} of open balls is (by
definition) a basis for the metric topology Td . The collection of open balls of radius n1 , n Z+ ,
is an equivalent topology basis for Td .
(2) The collection of rectangular regions (a1 , b1 ) (a2 , b2 ) in the plane R2 is a topology basis
equivalent to the standard basis of open balls B(a, r) = {x R2 | |x a| < r}.
(3) Let f : X Y be any map. If T is a topology on Y with basis B or subbasis S , then the
pull-back f 1 (T ) is a topology, the initial topology for f , on X with basis f 1 (B) and subbasis
f 1 (S).
(4) More generally, let fj : X Yj , j J, be a set of maps form a set X into spaces Yj . Let Tj
S
S
S
be a topology on Yj with basis Bj or subbasis Sj , j J. Then fj1 (Tj ), fj1 (Bj ), fj1 (Sj )
are equivalent subbases on X. The topology they generate is called the initial topology for the
maps fj , j J.
Example 10.6. [Topologies on R] We consider three topologies on R:
R: The standard topology with basis the open intervals (a, b).
R` : The lower limit or right half-open interval topology with basis the right half-open intervals [a, b).
RK : The K-topology with basis {(a, b)} {(a, b) K} where K = {1, 21 , 13 , 14 , . . .}.
The
S lower limit topology is strictly finer than the standard topology because (10.4) (a, b) =
a< [, b), and [0, 1) is open in R` but not in R (an open interval containing 0 is not a subset
of [0, 1).) The K-topology is strictly finer than the standard topology because its basis contains
the standard basis and R K is open in RK but not in R (an open interval containing 0 is not a
subset of R K). The topologies R` and RK are not comparable [Ex 13.6].
Example 10.7. The collection of of open intervals is a basis (10.6) and the collection of open rays
S = {(, b)} {(a, +)}
is a subbasis for the standard topology on R by 10.4.(3).

18

GENERAL TOPOLOGY

11. Order topologies


We associate a topological space to any linearly ordered set. You may view the topological space
as a means to study the ordered set, to find invariants, or you may view this construction as a
provider of interesting examples of topological spaces.
Let (X, <) be a linearly ordered set containing at least two points. The open rays in X are the
subsets
(, b) = {x X | x < b}, (a, +) = {x X | a < x}
of X. The collection of all open rays is clearly a subbasis (just as in 10.7).
Definition 11.1. The order topology T< on the linearly ordered set X is the topology with all open
rays as subbasis. A linearly ordered space is a linearly ordered set with the order topology.
The open intervals in X are the subsets of the form
(a, b) = (, b) (a, +) = {x X | a < x < b},

a, b X, a < b.

Lemma 11.2. The collection of all open rays together with all open intervals is a basis for the
order topology T< .
Proof. We get a basis from a subbasis by taking all finite intersections of subbasis sets (10.3). 
If X has a smallest elementSa0 then the open ray (, b) = [a0 , b). If X has no smallest element,
the the open ray (, b) = a<b (a, b) is a union of open intervals and we do not need this open
ray in the basis. Similar remarks apply to the greatest element when it exists. In particular, if X
has neither a smallest nor a greatest element then the collection of open intervals is a basis.
Lemma 11.3. If X has no smallest and no largest element, then the set {(a, b)} of open intervals
is a basis for the order topology.
Example 11.4. (1) The order topology on the ordered set (R, <) is the standard topology (10.6).
(2) The order topology on the ordered set R2 has as basis the collection of all open intervals
(a1 a2 , b1 b2 ). An equivalent basis (10.4) consists of the open intervals (a b1 , a b2 ).
(3) The order topology on Z+ is the discrete topology because (n 1, n + 1) = {n} is open.
(4) The order topology on Z+ Z+ is not discrete. Any open set that contains the element 2 1
also contains elements from {1} Z+ . Thus the set {1 2} is not open.
(5) The order topology on Z Z is discrete.
(6) Is the order topology on S discrete?
(7) I 2 = [0, 1]2 with the order topology is denoted Io2 and called the ordered square. The open sets
containing the point x y Io2 look quite different depending on whether y {0, 1} or 0 < y < 1.
12. The product topology
Q
Let (Xj )jJ be an indexed family of topological spaces. Let k : jJ Xj Xk be the projection map. An open cylinder is a subset of the product space of the form
j1 (Uj ),

Uj Xj open,

j J.

The collection
of all open cylinders is clearly a subbasis. The set k1 (Uk ) consists of the points
Q
(xj ) Xj with kth coordinate in Uk .
Q
Definition 12.1. The product topology on jJ Xj is the topology which has the collection of all
open cylinders as a subbasis.
Q
The product topology is the coarsest topology making all the projection maps j :
Xj Xj ,
j J, continuous.
Q
Lemma 12.2. The collection of all products jJ Vj where Vj Xj is open in Xj and Vj = Xj
for all but finitely many j J is a basis for the product topology.
Proof. We get a basis from a subbasis by taking all finite intersections of subbasis sets (10.3). 
This becomes particularly simple when we consider finite products.

GENERAL TOPOLOGY

19

Lemma 12.3. Let X = X1 X2 Xk be a finite Cartesian product. The collection


B = {U1 U2 Uk | U1 open in X1 , U2 open in X2 , . . . , Uk open in Xk }
of all products of open sets is a basis for the product topology.
Corollary 12.4. Suppose that Bj is a basis for the topology on Xj , j = 1, . . . , k. Then B1 Bk
is a basis for the product topology on X1 Xk .
Proof. Note that B1 Bk is indeed a basis. Compare it to the basis of 12.3 using 10.4.

When (X, <) and (Y, <) are linearly ordered sets, we now have two topologies on the Cartesian
product X Y : The product topology of the order topologies, (X, T< ) (Y, T< ), and the order
topology of the product dictionary order, (X Y, T< ). These two topologies are in general not
identical or even comparable. It is difficult to imagine any general relation between them since X<
Y< is essentially symmetric in X and Y whereas the dictionary order has no such symmetry. But
even when X = Y there does not seem to be a general pattern: The order topology (Z+ Z+ , T< )
is coarser than the product topology (Z+ , T< ) (Z+ , T< ) (which is discrete) (11.4.(4)(5)). On the
other hand, the order topology (R R, T< ) is finer than the product topology (R, T< ) (R, T< )
which is the standard topology, see 11.4.(1)(2) and [Ex 16.9].
Corollary 12.5 (Cf [Ex 16.9]). Let (X, <) and (Y, <) be two linearly ordered sets. Suppose that
Y does not have a largest or a smallest element. Then the order topology (X Y )< = Xd Y<
where Xd is X with the discrete topology. This topology is finer than X< Y< .
Proof. The equations
[
(, a b) =
{x} Y {a} (, b),

(a b, +) = {a} (b, +)

x<a

{x} Y

x>a

show that the order topology is coarser than the product of the discrete and the order topology. On
the other hand, the sets {x}(c, d) is a basis for = Xd Y< (11.3, 12.4) and {x}(c, d) = (xc, xd)
is open in the order topology since it is an open interval.

The corollary shows that the order topology on X Y does not yield anything new in case the
second factor is unbounded in both directions. As an example where this is not the case we have
already seen Z+ Z+ and I 2 o (11.4.(4), 11.4.(7)).
`
12.6. The coproduct topology. The coproduct topology
Xj (3.4) is the
` on the coproduct
finest topology making all the inclusion maps j : Xj Xj , j J, continuous.
This means that
a
U
Xj is open 1
j (U ) is open for all j J
jJ

Alternatively, the open sets of the coproduct

Xj are the coproducts

Uj of open set Uj Xj .

13. The subspace topology


Let (X, T ) be a topological space and Y X a subset. We make Y into a topological space.
Definition 13.1. The subspace topology on Y is the topology TY = {Y U | U T }. A subset
V Y is open relative to Y if it open in the subspace topology.
It is immediate that TY is indeed a topology. The subset V Y is open relative to Y if and
only if V = Y U for some open set U X.
Lemma 13.2. If B is a basis for T , then Y B = {Y U | U B} is a basis for the subspace
topology TY . If S is a subbasis for T , then Y S = {Y U | U S} is a subbasis for the subspace
topology TY .
Proof. This is 10.5.(3) applied to the inclusion map A , X.
If V Y is open, then V is also relatively open. The converse holds if Y is open.
Lemma 13.3. Assume that A Y X. Then
(1) A is open in Y A = Y U for some open set U in X

20

GENERAL TOPOLOGY

(2) If Y is open then: A is open in Y A is open in X


Proof. As the first point is clear by definition of the subspace topology, we only consider the second
point.
: True for any Y .
: We have A = Y U for some open set U X. Since both Y and U are open, A is open. 
The lemma says that an open subset of an open subset is open.
Subspace topology and product topology commute.
Q
Q
Theorem 13.4. Let Yj Xj , j J. The subspace topology that Yj inherits from Xj is the
product topology of the subspace topologies on Yj .
Proof. The collection of sets
Y 
Yj k1 (Uk ),

Uk Xk open,

(k :

Xj Xk )

is a subbasis for the subspace topology (13.2, 12.1). The collection of sets
Q
k1 (Yk Uk ), Uk Xk open, (k :
Yj Yk )
Q
is a subbasis for the product topology on the product Yj of subspaces (12.1, 13.1). These two
Q
[Ex 2.2]
collections are the same because k1 (Yk Uk ) = k1 (Yk ) k1 (Uk ) = ( Yj ) k1 (Uk ). 
When (X, <) is a linearly ordered set and Y X a subset we now have two topologies on Y .
We can view Y as a subspace of the topological space X with the order topology or we can view
Y as a sub-ordered set of X and give Y the order topology. In general, these two topologies are
not the same (13.6.(2)).
Lemma 13.5 (Y< Y X< ). Let (X, <) be a linearly ordered set and Y X a subset. The order
topology (Y, T< ) is coarser than the subspace topology Y T< . These two topologies are identical
when Y is convex, but in general, they are distinct.
Proof. When a and B are points in Y , the equalities
(, b)Y = Y (, b)X ,

(a, +)Y = Y (a, +)X

show that the subbasis for the order topology (11.1) on Y is contained in subbasis for the subspace
topology on Y (13.2). These two topologies are not the same when X = R (with order topology)
and Y = [0, 1) {2} R. (Any basis set (11.2) for the order topology on Y containing the point
2 also contains elements of [0, 1) so {2} is not open.)
If Y is convex and a a point outside Y , then a is either a lower or an upper bound for Y as we
can not have y1 < a < y2 for two points y1 , y2 of Y . Therefore

(a, +)Y a Y
Y (a, +)X = Y
a 6 Y, a a lower bound for Y

a 6 Y, a an upper bound for Y


This and the similar statement for Y (, b)X show that the subbasis for the subset topology
on Y is contained in the subbasis for the order topology on Y .

Example 13.6. (1) S 21 = 1 Z+ {2 1} = [1 1, 2 1] is a convex subset of Z+ Z+ . The
subspace topology (11.4.(4)) is the same as the order topology.
(2) The subset Z+ Z+ is not convex in Z Z so we expect the subspace topology to be strictly
finer than the order topology. Indeed, the subspace topology that Z+ Z+ inherits from the
discrete space Z Z is discrete but the order topology is not discrete (11.4.(4)(5)).
(3) Consider X = R2 and Y = [0, 1]2 with order topologies. Y is not convex so we expect the
subspace topology to be strictly finer than the order topology. Indeed, the subspace topology on
[0, 1]2 , which is [0, 1]d [0, 1], is strictly finer than Io2 (11.4.(7)): The set
1
1
[0, ] [0, 1] = ([0, 1] [0, 1]) (, 2)
2
2
is open in the subspace topology on Y but it is not open in the order topology on Y as any basis
open set (11.2) containing 21 1 also contains points with first coordinate < 12 .

GENERAL TOPOLOGY

21

(4) Consider X = R2 and Y = (0, 1)2 with order topologies. Y is not convex so we expect the
subspace topology to be strictly finer than the order topology. But it isnt! The reason is that
(0, 1) does not have a greatest nor a smallest element (12.5).
(5) The subset Q of the linearly ordered set R is not convex but nevertheless the subspace topology
inherited from R is the order topology. Again, the reason seems to be that Q does not have a
greatest nor a smallest element.

22

GENERAL TOPOLOGY

14. Closed sets and limit points


Let (X, T ) be a topological space and A a subset of X.
Definition 14.1. We say that the set A is closed if its complement X A is open.
Note that X and are closed (and open), a finite union of closed sets is closed, an arbitrary
intersection of closed sets is closed. The finer the topology, the more closed sets.
Example 14.2. (1) [a, b] is closed in R, [a, b) is neither closed nor open, R is both closed and
open.
(2) Let X = [0, 1] (2, 3). The subsets [0, 1] and (2, 3) of X are both closed and open (they are
clopen) as subsets of X. The interval [0, 1] is closed in R while (2, 3) is open in R.
(3) K = {1, 21 , 31 , . . .} is not closed in R but it is closed in RK .
(4) Recall that R` is the set of real numbers equipped with the topology generated by the basis
sets [a, b) of right half-open intervals. All sets that are (open) closed in the standard topology
R
S are also (open) closed in the finer topology R` . Sets of the form (, a), [a, b), and [a, ) =
a<x [a, x) are both open and closed. Sets of the form (, b] or [a, b] are closed and not open
since they are not unions of basis sets. Sets of the form (a, ) are open and not closed. Sets of
the form (a, b] are neither open nor closed.
(5) Let X be a well-ordered set. In the order topology, sets of the form (a, ) = [a+ , ),
(, b] = (, b+ ) and (a, b] = (a, ) (, b] are closed and open. (Here, b+ denotes b if b is
the largest element and the immediate successor of b if b is not the largest element.)
14.3. Closure and interior. We consider the largest open set contained in A and the smallest
closed set containing A.
Definition 14.4. The interior of A is the union of all open sets contained in A:
[
Int A = {U A | U open} = A
The closure of A is the intersection of all closed sets containing A:
\
Cl A = {C A | C closed} = A
Note or prove [6, Ex 17.6, 17.7, 17.8, 17.9] [3, p 9] [5, 7.1.79] that
A A A
A is open, it is the largest open set contained in A, and A is open iff A = A
A is closed, it is the smallest closed set containing A, and A is closed iff A = A
(A B) = A B , A B = A B, X A = X A , (X A) = X A
For any A X and B Y , we have (A B) = A B and A B = A B in the
Cartesian 
product X Y (15.12) [4, p 88].
Int A Int B
AB
Cl A Cl B
Definition 14.5. A neighborhood of the point x X is an open set containing x.
Proposition 14.6. Let A X and let x be a point in X. Then
x A There exists a neighborhood of x that is contained in A
x A All neighborhoods of x intersect A nontrivially
and X A = (X A) , X A = X A.
Proof. The first assertion is clear as A is the union of open sets contained in A. We now compute
\
[
X A = X {C A | C closed} = {X C X A | X C open}
[
= {U X A | U open} = (X A)
and, similarly, X A = X A. The second assertion is just a reformulation,
x 6 A x (X A) x has a neighborhood that does not intersect A,
of the above formula.

GENERAL TOPOLOGY

23

Proposition 14.7 (Closure with respect to subspace). Let A Y X. Then


(1) A is closed in Y A = Y C for some closed set C in X
(2) If Y is closed then: A is closed in Y A is closed in X
(3) ClY (A) = Y A.
Proof. (1) We have A = Y C Y A = Y (X C) for any subset C of X (make a drawing).
The claim therefore follows from (is equivalent to) the corresponding claim for relatively open sets.
(2) is now obvious.
(3) The set ClY (A) is the intersection of all relatively closed sets containing A. The relatively
closed sets are the sets of the form Y C where C is closed in X. This means that
\
\
ClY (A) = {Y C | C A, C closed} = Y {C | C A, C closed} = Y A
by a direct computation.

Definition 14.8. A subset A X is said to be dense if A = X, or, equivalently, if every open


subset of X contains a point of A.
Proposition 14.9. Let A be a dense an U an open subset of X. Then A U = U .
Proof. The inclusion A U U is general. For the other inclusion, consider a point x U . Let
V be any neighborhood of x. Then V (A U ) = (V U ) A is not empty since V U is a
neighborhood of x and A is dense. But this says that x is in the closure of A U .

14.10. Limit points and isolated points. Let X be a topological space and A a subset of X.
Definition 14.11. A point x X is a limit point of A if every neighborhood of x contains a point
from A {x}. The set of limit points1 of A is denoted A0 . A point a A is an isolated point if a
has a neighborhood that intersects A only in {a}.
Equivalently, x X is a limit point of A iff x A {x}, and a A is an isolated point of A iff
{a} is open in A. These two concepts are almost each others opposite:
x is not a limit point of A x has a neighborhood U such that U A {x}
x has a neighborhood U such that U (A {x}) = {x}
x is an isolated point of A {x}
Proposition 14.12. Let A be a subset of X and A0 the set of limit points of A. Then A A0 = A
and A A0 = A {a A | a is isolated} so that
A A0 A is closed
A A0 A has no isolated points
A A0 = A is discrete
A0 = A is closed and discrete
If B A then B 0 A0 .
Proof. Since A A A0 A we get that A A0 = A. The intersection A A0 is precisely the set
of nonisolated points of A. If A A0 = then all all points of A are isolated so that the subspace
A has the discrete topology. We have A0 = A A0 , A A0 = .


Example 14.13. (1) Q = R = R Q (RQ contains (Q{0}) 2). Cl Int Q = , Int Cl Q = R.


(2) Let Cr R2 be the circle with center (0, r) and radius r > 0. Then
[
[
[
[
C1/n =
C1/n ,
Cn = (R {0})
Cn
nZ+

nZ+

nZ+

nZ+

The first set of decreasing circles (known as the Hawaiian Earring [6, Exmp 1 p 436]) is closed
because it is also the intersection of a collection (which collection?) of closed sets.
(3) The set of limit points of K = { n1 | n Z+ } is {0} in R and R` and in RK .
1The set of limit points of A is sometimes denotes Ad and called the derived set of A

24

GENERAL TOPOLOGY

14.14. Convergence, the Hausdorff property, and the T1 -axiom. Let xn , n Z+ , be a


sequence of points in X, that is a map Z+ X.
Definition 14.15. The sequence xn converges to the point x X if for any neighborhood U of x
there exists some N such that n > N xn U .
If a sequence converges in some topology on X it also converges in any coarser topology but not
necessarily in a finer topology.
Example 14.16. In the Sierpinski space X = {0, 1}, the sequence 0, 0, converges to 0 and to
1. In R with the cofinite topology, the sequence 1, 2, 3, converges to any point. In RK , the
sequence n1 does not converge; in R and R` it converges to 0 (and to no other point). The sequence
n1 converges to 0 in R and RK , but does not converge in R` . In Z+ Z+ , the sequence 1 n
converges to 2 1 (and to no other point).
Definition 14.17. A topological space X is Hausdorff (or T2 ) if there are enough open sets to
separate points: For any two distinct points x1 6= x2 in X there exist disjoint open sets, U1 and
U2 , such that x1 U1 and x2 U2 .
All linearly ordered spaces are Hausdorff [Ex 17.10]. R` and RK are Hausdorff because the
topologies are finer than the standard Hausdorff topology R.
Theorem 14.18. A sequence in a Hausdorff space can not converge to two distinct points.
A property of a topological space is said to be (weakly) hereditary if any (closed) subspace of
a space with the property also has the property. Hausdorffness is hereditary and also passes to
product spaces.
Theorem 14.19. [Ex 17.11, 17.12] Any subset of a Hausdorff space is Hausdorff. Any product of
Hausdorff spaces is Hausdorff.
Definition 14.20. A topological space is T1 -space if points are closed.
Equivalently, X is T1 iff all finite subsets are closed. Cofinite topologies are T1 by construction.
All Hausdorff spaces are T1 . Not all T1 -spaces are Hausdorff as R with the cofinite topology shows.
The Sierpinski space is not T1 .
Theorem 14.21. Suppose that X is T1 . Let A be a subset of and x a point in X. Then
x is a limit point All neighborhoods of x intersect A in infinitely many points
Proof. : If all neighborhoods of x intersect A in infinitely many points, then, clearly, they also
intersect A in a point that is not x.
: Assume that x is a limit point. Let U be a neighborhood of x. Then U contains a point
of A different from x. Remove this point from U . Since points are closed, the result is a new
neighborhood of x. This new neighborhood of x contains a point of A different from x. In this
way we recursively find a whole sequence of distinct points in U A.

15. Continuous functions
Let f : X Y be a map between two topological spaces.
Definition 15.1. The map f : X Y is continuous if all open subsets of Y have open preimages
in X: V open in X f 1 (V ) open in Y
If TX is the topology on X and TY is the topology on Y then
(15.2)

f is continuous f 1 (TY ) TX f 1 (BY ) TX f 1 (SY ) TX

where BY is a basis and SY a subbasis for TY . The finer the topology in Y and the coarser the
topology in X are, the more difficult is it for f : X Y to be continuous.
Theorem 15.3. Let f : X Y be a map between two topological spaces. The following are equivalent:
(1) f is continuous
(2) The preimage of any open set in Y is open in X: V open in Y f 1 (V ) open in X

GENERAL TOPOLOGY

(3)
(4)
(5)
(6)
(7)

25

The preimage of any closed set in Y is closed in X: V closed in Y f 1 (V ) closed in X


f 1 (B ) (f 1 (B)) for any B Y
f 1 (B) f 1 (B) for any B Y
f (A) f (A) for any A X
For any point x X and any neighborhood V Y of f (x) there is a neighborhood U X
of x such that f (U ) V .

Proof. It is easy to see that the first


 three conditions are equivalent.
f 1 (B ) f 1 (B)
(2) (4):
f 1 (B ) (f 1 (B)) .
f 1 (B ) open
(4) (2): Let B be any open set in Y . Then f 1 (B)
f 1 (B) is open since it equals its own interior.
(3) (5): Similar to (2) (4). 
A f 1 (f (A))
A f 1 (f (A)).
(3) (6):
f 1 (f (A)) is closed
f (f 1 (B))B

(6)

(6) (5): f (f 1 (B)) f (f 1 (B))

(2) (7): This is just a reformulation.

B=B

(4)

f 1 (B ) (f 1 (B)) f 1 (B) so

B


Example 15.4. The function f (x) = x is continuous R R, but not continuous RK RK


(K is closed in RK but f 1 (K) is not closed (14.16)) and not continuous R` R` ([0, ) is open
in R` but f 1 ([0, )) = (, 0] is not open (14.2.(4))). Since (a, b] is closed and open in R` the
map 1[a,b) : R` {0, 1}, with value 1 on [a, b) and value 0 outside [a, b), is continuous.
Theorem 15.5.
(1) The identity function is continuous.
(2) The composition of two continuous functions is continuous.
(3) Let B be a subspace of Y , A a subspace of X, and f : X Y a map taking values in B.
Then
f : X Y is continuous = f |A : A Y is continuous

(restriction)

f : X Y is continuous B|f : X B is continuous

(corestriction)

(4) Let fj : Xj Yj , j J, be an indexed set of maps. Then


Y
Y
Y
fj :
Xj
Yj is continuous fj : Xj Yj is continuous for all j J
S
(5) Suppose that X = U is a union of open sets. Then
f : X Y is continuous f |U is continuous for all
for any map f : X Y .
(6) Suppose that X = C1 Cn is a finite union of closed sets. Then
f : X Y is continuous f |Ci is continuous for all i = 1, . . . , n
for any map f : X Y .
Proof. This easily checked. The =-part of (4) uses 15.11 below.

15.6. Homeomorphisms and embeddings. One of the central problems in topology is to decide
if two given spaces are homeomorphic.
Definition 15.7. A homeomorphism is a bijective continuous map f : X Y with a continuous inverse X Y : f 1 . An embedding is an injective continuous map f : X Y such that
f : X f (X) (where f (X) has the topology as a subspace of Y ) is a homeomorphism.
A bijection f : X Y is a homeomorphism iff
U is open in X f (U ) is open in Y
holds for all subsets U of X iff
f 1 (V ) is open in X V is open in Y
holds for all subsets V of Y .

26

GENERAL TOPOLOGY

An injective map f : X Y is an embedding if


U is open in X f (U ) is open in f (X)
holds for all subsets U of X. An embedding is a homeomorphism followed by an inclusion.
Example 15.8. (1) The map f (x) = 3x + 1 is a homeomorphism R R.
(2) The identity map R` R is bijective and continuous but not a homeomorphism.
(3) The map [0, 1) S 1 : t 7 (cos(2t), sin(2t)) is continuous and bijective but not a homemorphism. The image of the open set [0, 21 ) is not open in S 1 .
(4) The obvious bijection [0, 1) {2} [0, 1] is continuous but not a homeomorphism. There does
not exist any continuous surjection in the other direction.
(5) The spaces [1 1, 2 1] Z+ Z+ and { n1 | n Z+ } R are homeomorphic.
(6) The map R R R : t 7 (t, t) is an embedding. For any continuous map f : X Y , the
map X X Y : x 7 (x, f (x)) is an embedding; see (27.4) for a generalization.
(7) Rn embeds into S n via stereographic projection.
3
3
2
(8) R2 embeds in R
S . Does RS imbed in R ?
(9) Are the spaces Cn and C1/n of 14.13.(2) homeomorphic?
(10) A knot is in embedding of S 1 in R3 (or S 3 ). Two knots, K0 : S 1 , R3 and K1 : S 1 , R3 ,
are equivalent if there exists a homeomorphism h of R3 such that h(K0 ) = K1 . The fundamental
problem of knot theory [1] is to classify knots up to equivalence.
Lemma 15.9. If f : X Y is a homeomorphism (embedding) then the corestriction of the restriction f (A)|f |A : A f (A) (B|f |A : A B) is a homeomorphism (embedding) for any subset A of
X (and any subset B of Y such that
Qf (A)Q B). If the maps fj : Xj Yj are homeomorphisms
(embeddings) then the product map fj :
Xj Yj is a homeomorphism (embedding).
Proof. In case of homeomorphisms there is a continuous inverse in both cases. In case of embeddings, use that an embedding is a homeomorphism followed by an inclusion map.

15.10. Maps into product spaces. There is an easy test for when a map into a product space
Q
is continuous. Recall that the product topology has as basis the collection of all product sets Uj
where Uj is open in Xj for all j and Uj = Xj for all except finitely many j. The collection of all
sets of the from j1 (Uj ) (ie all factors but one is Xj ) is a subbasis.
Q
Theorem 15.11. Let f : X jJ Yj be a map into a product space. Then
Y
f: X
Yj is continuous j f : X Yj is continuous for all j J
jJ

where j :

jJ

Yj Yj is the projection map.

S
Proof. Let TX be the topology on X and Tj the topology on Yj . Then jJ j1 (Tj ) is a subbasis
Q
for the topology on jJ Yj . Therefore
Y
[
f: X
Yj is continuous f 1 (
j1 (Tj )) T
jJ

jJ

f 1 (j1 (Tj )) T

jJ

j J : (j f )1 (Tj ) T
j J : j f is continuous
by definition of continuity (15.2).

Similarly there was (15.5.(3)) an easy test for when a map f : X B into a subspace B Y
is continuous: X B is continuous X B Y is continuous. The secret is that the
subspace is equipped with the initial topology for the inclusion map and that the product space is
equipped with the initial topology for the projection maps.
Q
Let (Xj )jJQbe an indexed family of topological spaces with subspaces Aj Xj . Then Aj is
a subspace of Xj .
Q
Q
Aj = Aj .
Theorem 15.12.

GENERAL TOPOLOGY

27

Proof. Let (xj ) be a point of Xj . Then


Y
Y
Y
Y
(xj )
Aj
Uj
Aj 6= for any basis neighborhood
Uj of (xj )
Y
Y

Uj Aj 6= for any basis neighborhood


Uj of (xj )
j : Uj Aj 6= for any neighborhood Uj of xj
j : xj Aj
Y
Aj
(xj )

It follows that a product of closed sets is closed. (Whereas a product of open sets need not be
open in the product topology.)
Q
Definition
Xj is the toplogy with basis consisting of all
Q 15.13. The box toplogy on the set
products Uj of open set Uj Xj .
The box
Q topology is finer than the product topology (which has as basis the collection of all
products Uj of open set Uj Xj where Uj = Xj except for finitely many j J) and equal to
the product topology for finite products. The projection maps j are continuous and open also in
the box topology and the above theorem also holds for the box topology (with the same proof).
However, since there are more open sets in the box topology it is more difficult for a map into the
product to be continuous. A concrete example is this.
Q
Example 15.14. Let : R Z+ R be the diagonal map of R into the product of countably
many copies of R. is continuous if we use
Q the product topology (15.11) but not continuous if we use the box topology; for instance nZ+ (1/n, 1/n) is open in the box topology but
Q
 T
1
nZ+ (1/n, 1/n) =
nZ+ (1/n, 1/n) = {0} is not open.
Thus the easy characterization (15.11) of continuous maps into the product topology does not
work for the box topology.
15.15. Maps out of coproducts.
`
Theorem 15.16. Let f : jJ Xj Y be a map out of a coproduct space. Then
a
f:
Xj Y is continuous f j : Xj Y is continuous for all j J
jJ

where j : Xj

jJ

Xj is the inclusion map.

Let fj : Xj Yj , j J, be an indexed set of maps. Then


a
a
a
fj :
Xj
Yj is continuous fj : Xj Yj is continuous for all j J

28

GENERAL TOPOLOGY

16. Quotient maps


In this section we will look at the quotient space construction. But first we consider open and
closed maps.
16.1. Open and closed maps. [2, I.5] Let X and Y be topological spaces and f : X Y a
map.


open
Definition 16.1. The map f : X Y is
if for all U X we have:
closed




open
open
U
in X f (U )
in Y
closed
closed
The restriction (16.10) of an open (or closed) map to an arbitrary subspace need not be open
(closed). However,
Proposition 16.2. [6, Ex 22.5] The restriction of an open (or closed) map f : X Y to an open
(closed) subspace A X is an open (closed) map f |A : A f (A) or f |A : A Y .
Proof. Let U A be any subset. The implications
U is open in A

(A is open in X)

U open in X

(f is open)

f (U ) open in Y
f (U ) open in f (A)
show that f |A : A f (A) is open.

A bijective continuous and open or closed map is a homeomorphism.


Example 16.3. (1) The projection map 1 : R R R is continuous and open. It is not a
closed map for H = {(x, y) R R | xy = 1} is closed but 1 (H) = R {0} is not closed.
(2) The map f : [1, 2] [0, 1] given by

0 1 x 0
f (x) = x 0 x 1

1 1x2
is continuous and closed (20.9.(1)). It is not open for f ([1, 1/2)) = {0} is not open.
(3) The map R` {0, 1} with value 1 on [0, 1) and value 0 elsewhere is continuous because
(14.2.(4)) the half-open interval [0, 1) is both closed and open. This map is closed and open
because all subsets of the discrete space {0, 1} are closed and open. See 16.13 for another example
of a closed and open map.
16.2. Quotient maps. Quotient maps are continuous surjective maps that generalize both continuous, open surjective maps and and continuous, closed surjective maps.
Let X and Y be topological spaces and p : X Y a continuous surjective map.
Definition 16.4. The continuous surjective map p : X Y is a quotient map if for all V Y
p1 (V ) is open in X V is open in Y
S
Sets of the form p1 (V ) = yV p1 (y), consisting of unions of fibres, are called saturated sets.
The saturation of A X is the union f 1 f (A) of all fibres that meet A.
Proposition 16.5. For a surjective map p : X Y the following are equivalent:
(1) p : X Y is a quotient map




open
open
(2) For all V Y we have: p1 (V ) is
in X V is
in Y .
closed 
closed



open
open
open sets
(3) p : X Y is continuous and maps saturated
sets to
closed
closed

GENERAL TOPOLOGY

29

Proof. Condition (1) and condition (2) with the word open are clearly equivalent. Suppose now
that: p1 (V ) is open V is open. Then we get
p1 (C) is closed X p1 (C) is open p1 (Y C) is open
Y C is open C is closed
for all C Y . This shows that the two conditions of (2) are equivalent. The content of (3) is just
a reformulation of (2).

Thus a quotient map p : X Y induces a bijective correspondence




Saturated open (closed) sets o U p(U ) / Open (closed) sets
in Y
in X
p1 (V )V
Just as for open (closed) continuous bijective maps we evidently have:
Corollary 16.6. A bijective continuous map is a quotient map if and only if it is a homeomorphism.
Quotient maps behave nicely with respect to composition:
f

Corollary 16.7. Let X

/Y

/ Z be continuous maps. Then

f and g are quotient g f is quotient g is quotient


Proof. The first assertion is a tautology: Assume that f and g are quotient. Then
(g f )1 (V ) open in X f 1 g 1 (V ) open in X

f is quotient

g 1 (V ) open in Y
g is quotient

V open in Z

for any set V Y . Next, suppose that X


Y
Z is a quotient map. Then the last map g is
surjective and
g 1 (V ) is open in Y

f continuous

(g f )1 (V ) is open in X

g f quotient

V is open in Z

for any set V Z.

If p : X Y is quotient there is a simple criterion for continuity of maps out of Y .


Lemma 16.8. Let p : X Y be a quotient map and let f : Y Z be some map of Y into the
topological space Z. Then
f

Y
Z is continuous X
Y
Z is continuous
Proof. This is because for all U Z,
f 1 (U ) is open in Y (f p)1 (U ) = p1 f 1 (U ) is open in Y
since p is a quotient map.



open
Corollary 16.9. Any
continuous surjective map is a quotient map. A quotient map
closed






open
open
open
f : X Y is
if and only if all
sets A X have
saturations f 1 f (A).
closed
closed
closed
Proof. This is immediate from the defintion of the quotient topology:
f is open f (A) is open in Y for all open sets A X
f 1 f (A) is open in X for all open sets A X

There are quotient maps that are neither open nor closed [6, Ex 22.3] [2, Ex 10 p 135] and there
are (non-identity) quotient maps that are both open and closed (16.13) [2, Ex 3 p 128].

30

GENERAL TOPOLOGY

Example 16.10. (1) The projection map 1 : R R R is (16.3.(1)) open, continuous, and
surjective so it is a quotient map. The restriction 1 |H {(0, 0)} is continuous and surjective,
even bijective, but it is not a quotient map (16.5) for it is not a homeomorphism: {(0, 0)} is open
and saturated in H {(0, 0)} but 1 ({(0, 0)} = {0} is not open.
Thus the restriction of a quotient map need not be a quotient map in general. On the positive
side we have
Proposition 16.11. The restriction of a quotient map p : X Y to an open (or closed) saturated
subspace A X is a quotient map p|A : A p(A).
Proof. (Similar to the proof of 16.2.) Let p : X Y be a quotient map and B Y an open set.
(The case where B is closed is similar.) The claim is that p|p1 (B) : p1 (B) B is quotient. For
any U B the implications
p1 (U ) open in p1 (B)

(p1 (B) is open)

p1 (U ) open in X
U is open in Y
U is open in B

(p is quotient)

show that p|p1 (B) : p1 (B) B is quotient.

We can always make a surjective map defined on a topological space into a quotient map.
Definition 16.12. Let X be a topological space, Y a set, and p : X Y a surjective map. The
quotient topology on Y is the collection
{U Y | p1 (U ) is open in X}
of subsets of Y .
The quotient topology2 is the finest topology on Y such that p : X Y continuous.
A typical situation is when R is an equivalence relation on the space X and X X/R is the
map that takes a point to its equivalence class. We call X/R with the quotient topology for the
quotient space of the equivalence relation R. A set of equivalence classes is an open subset of
X/R
if and only if the union of equivalence classes is an open subset of X: V X/R is open
S
[x]V [x] X is open. We shall often say that X/R is the space obtained by identifying equivalent
points of X.
Example 16.13. (1) The real projective n-space RP n is the quotient space of S n by the equivalence relation with equivalence classes {x}, x S n . The quotient map p : S n RP n is both
open and closed since (16.9) the saturation U of an open (closed) set U S n is open (closed)
because x x is a homeomorphism. Elements of RP n can be thought of as lines through the
origin of Rn+1 . A set of lines is open if the set of intersection points with the unit sphere is open.
Does RP 2 embed in R3 ?
(2) More generally, let G X X be the action of a discrete group G on a space X. Give the
orbit space G\X the qutient topologySand let pG : X G\X be the quotient map. The saturation
of any set A X is the union GA = ginG gA of orbits intersecting A. If A is open, its saturation
GA is also open. Thus pG is an open map (16.9). The saturated open sets are precisely the sets of
the form GA where A is open in X. The quotient map is closed if G is finite.
A continuous map f : X Y respects the equivalence relation R if equivalent points have
identical images, that is if x1 Rx2 f (x1 ) = f (x2 ). The quotient map X X/R respects
the equivalence relation R and it is the universal example of such a map.
Theorem 16.14. Let f : X Y be a continuous map. Then
(1) The map p : X X/R respects the equivalence relation R.
2The quotient topology is sometimes called the final topology [2, I.2.4] with respect to the map f

GENERAL TOPOLOGY

31

(2) If the continuous map f : X Y respects R then there exists a unique continuous map
f : X Y such that
f
/Y
XA
AA
}>
}
AA
}
A
}}
p AA
}} f
X/R

commutes. (We say that f factors uniquely through X/R.) Conversely, if f factors through
X/R then f respects R.
(3) If f exists then: f is quotient f is quotient
Proof. If f exists then clearly f respects R. Conversely, if f respects R then we can define
f [x] = f (x) and this map is continuous by 16.8 and it is the only possibility. The rest follows from
16.7.

The theorem says that there is a bijective correspondence




f f
Continuous maps X Y o
/ Continuous maps
gpg
X/R Y
that respect R
taking quotient maps to quotient maps.
Example 16.15. (1) Let f : X Y be any surjective
continuous map. Consider the equivalence
S
relation corresponding to the partition X = yY f 1 (y) of X into fibres fibres f 1 (y), y Y .
Let X/f denote the quotient space. Thus X/f is the set of fibres equipped with the quotient
topology. By constuction, the map f respects this equivalence relation so there is a unique
continuous map f such that the diagram
/Y
XC
CC
{=
{
CC
{{
CC
{{f
C!
{
{
X/f
f

commutes. Note that f is bijective. The bijective continuous map f : X/f Y is a homeomorphism if and only if f is quotient (16.6, 16.14). Thus all quotient maps have up to homeomorphism
the form X X/R for some equivalence relation R on X.
(2) Let f : X Y be any surjective continuous map. The induces map f : X/f Y is a continuous bijection but in general not a homeomorphism. Instead, the quotient space X/f is homeomorphic to Y with the quotient topology which is finer than the given topology. For instance,
X/f is Hausdorff if Y is Hausdorff.
(3) Let f : X Y be any continuous map. Then f has a canonical decomposition
p

X
X/f
f (X)
Y
where p is a quotient map, f is a continuous bijection, and is an inclusion map.
(4) Let X be a topological space and A1 , A2 , . . . , Ak a finite collection of closed subsets. Consider
the equivalence relation where the equivalence classes are the sets A1 , A2 , . . . , Ak together with
the sets {x} for x 6 A1 A2 Ak . The quotient space X/(A1 , . . . , Ak ) is obtained from X
by identifying each of the sets Ai to the point p(Ai ). The quotient map
S p : X X/(A1 , . . . , Ak )
is closed because (16.9) closed sets A X have closed saturations A Ai A6= Ai . A continuous
map f : X Y factors through the quotient space X/(A1 , . . . , Ak ) if and only if it sends each of
the sets Ai X to a point in Y (16.14). The restriction p|X (A1 Ak ) : X (A1 Ak )
X/(A1 , . . . , Ak ) {p(A1 ), . . . , p(Ak )} to the complement of A1 Ak is a homeomorphism
(16.2). In case of just one closed subspace A X, the quotient space is denoted X/A.
(5) The standard map f : [0, 1] S 1 that takes t [0, 1] to (cos(2t), sin(2t)) is quotient because it is continuous and closed. (If you cant see this now, we will prove it later (20.8.(1)).)
The induced map [0, 1]/{0, 1} S 1 is a homeomorphism. More generally, the standard map
Dn /S n1 S n is a homeomorphism where Dn Rn , the unit disc, is the set of vectors of
length 1.

32

GENERAL TOPOLOGY

(6) Let R be the equivalence relation zero or not zero on R. The quotient space R/R is
homeomorphic to Sierpinski space {0, 1}.
S
(7) There is an obvious continuous surjective map f : Z+ S 1 Cn (14.13.(2)) that takes
Z+ {1} to the point common to all the circles. This map is continuous because its restriction
to each of the open sets {n} S 1 is continuous (15.5.(5). However, f is not a quotient map
(16.5) for the image of the closed saturated set consisting of the points n (cos( 2 ), sin( 2 ))
is not closed as it does not
S contain all its limit points. The induced bijective continuous map
f : Z+ S 1 /Z+ {1} Cn isStherefore not a homeomorphism. There is an obvious continuous
surjective map g : Z+ S 1 C1/n (14.13.(2)) that takes Z+ {1} to the point common to
all the circles. This map is continuous because its restriction to each of the open sets {n} S 1
is continuous (15.5.(5)). However, g is not a quotient map for the image of the closed saturated
set Z+ {1} is not closed as it does not S
contain all its limit points. The induced bijective
C1/n is therefore not a homeomorphism either.
continuous map g : Z+ S 1 /Z+ {1}
(Actually
(17.10.(5)),
the
quotient
space
Z

S 1 /Z+ {1}, known as the countable wedge of


+
W
1
circles nZ+ S [6, Lemma 71.4], is not homeomorphic to any subspace of the plane.)
(8) [6, p 451] Let P4g be a regular 4g-gon and with edges labeled a1 , b1 , a1 , b1 , . . . , ag , bg , ag , bg in
counter-clockwise direction. The closed orientable surface Mg of genus g 1 is (homeomorphic
to) the quotient space P4g /R where R is the equivalence relation that makes the identifications
1
1 1
a1 b1 a1
1 b1 ag bg ag bg on the perimeter and no identifications in the interior of the polygon.
See [6, p 452] for the case g = 2. Are any of these surfaces homeomorphic to each other? [6, Thm
77.5]
(9) [6, p 452] Let P2g be a regular 2g-gon with edges labeled a1 , a1 , . . . , ag , ag in counter-clockwise
direction. The closed non-orientable surface Ng of genus g 1 is the quotient space P2g /R where
R is the equivalence relation that makes the identifications a21 a2g on the perimeter and no
identifications in the interior of the polygon. For g = 1 we get the projective plane RP 2 and for
g = 2 we get the Klein Bottle [6, Ex 74.3].
Example 16.16. (The adjunction space) [6, Ex 35.8] [4, p 93] [7, Chp 1, Exercise B p 56] Consider
f
i ?_
/ Y consisting of a space X and a continuous map f : A Y defined
the set-up X o
Y
on the closed subspace A X. Let R be the smallest equivalence relation on X q Y such that
aRf (a) for all a A; the equivalence classes of R are {a} q f 1 (a) for a A, {x} for x X A,
anf f 1 (y) q {y} for y Y . The adjunction space is the quotient space
X f Y = X q Y /R
for the equivalence relation R. Let f : X X f Y be the map X X q Y X f Y and let
i : Y X f Y be the map Y X q Y X f Y . These two continuous maps agree on A in the
sense that f i = i f and the adjunction space is the universal space with this property. For any
other space Z receiving maps X Z Y that agree on A there exists a unique continuous map
X f Y such that the diagram
f

A
_
i


X

/Y
i


/ X f Y
!

#, 
Z
commutes. The map i : Y X f Y is closed for closed sets B Y X q Y have closed
saturations f 1 (B) q B. Since i is injective it is an embedding; its image is a closed subspace
of X f Y homeomorphic to Y . The map f |X A : X A X f Y is open for open sets
U X A X q Y have open saturations U q . Since f |X A is injective it is an embedding; its
image is an open subspace of X f Y homeomorphic to X A. The quotient map X q Y X f Y
is closed if the map f is closed for then also closed subsets B X X q Y have closed saturations
B f 1 f (B A) q f (A B). We shall later see [6, Ex 35.8] that X f Y is normal when X and
Y are normal.

GENERAL TOPOLOGY

33

Only few topological properties are preserved by quotient maps. The reason is that surjective
open maps and surjective closed maps are quotient maps so that any property invariant under
quotient maps must also be invariant under both open and closed maps.
As we saw in 16.15.(6) the quotient space of a Hausdorff space need not be Hausdorff, not even
T1 . In general, the quotient space X/R is T1 if and only if all equivalence classes are closed sets.
(For instance, if X and Y are T1 then also the adjunction space X f Y is T1 .) The quotient space
X/R is Hausdorff if and only if any two distinct equivalence classes are contained in disjoint open
saturated sets. We record an easy criterion for Hausdorfness even though you may not yet know
the meaning of all the terms.
Proposition 16.17. If X is regular and A X is closed then the quotient space X/A is Hausdorff.
The product of two quotient maps need not be a quotient map [2, I.5.3]. See [6, Ex 29.11] for
an important example where the product of two quotient maps is quotient, though.

34

GENERAL TOPOLOGY

17. Metric topologies


If X is a set with a metric d : X X [0, ) the collection {Bd (x, ) | x X, > 0} of balls
Bd (x, ) = {y X | d(x, y) < } is a basis for the metric topology Td induced by d.
Definition 17.1. A metric space is the topological space associated to a metric set. A topological
space is metrizable if the topology is induced by some metric on X.
Hausdorff dimension, fractals, or chaos are examples of metric, not topological, concepts.
Which topological spaces are metrizable? To address this question we need to build up an
arsenal of metrizable and non-metrizable spaces and to identify properties that are common to all
metrizable spaces. Here are the first such properties: All metric spaces are Hausdorff and first
countable.
Theorem 17.2 (Continuity in the metric world). Let f : X Y be a map between metric spaces
with metrics dX and dY , respectively. The following conditions are equivalent:
(1) f is continuous
(2) x X > 0 > 0y X : dX (x, y) < dY (f (x), f (y)) <
(3) x X > 0 > 0 : f (BX (x, )) BY (f (x), )
Proof. Essentially [6, Ex 18.1].

Proposition 17.3 (Comparison of metric topologies). Let d and d0 be metrics on X and Td , Td0
the associated metric topologies. Then
Td Td0 x X > 0 > 0 : Bd0 (x, ) Bd (x, )
Proof. Td Td0 if and only if the identity map (X, Td0 ) (X, Td ) is continuous [6, Ex 18.3] .

Lemma 17.4 (Standard bounded metric). Let d be a metric on X. Then d0 (x, y) = min{d(x, y), 1}
is a bounded metric on X (called the standard bounded metric corresponding to d) that induce the
same topology on X as d.
Proof. Either use the above proposition or use that for any metric the collection of balls of radius
< 1 is a basis for the metric topology. These bases are the same for the two metrics.

Theorem 17.5. Any subspace of a metrizable space is a metrizable. Any countable product of
metrizable spaces is metrizable.
Proof. See [6, Ex 21.1] for the first assertion. To prove the second assertion, let Xn , n Z+ , be a
countable collection of metric spaces. We may assume that each Xn has diameter at most 1 (17.4).
Put
1
d((xn ), (yn )) = sup{ dn (xn , yn ) | n Z+ }
n
Q
for points (xn ) and (yn ) of Xn and convince yourself that d isQa metric.
Since d(x, y) < /n dn (x, y) < , the projection maps n :
Xn Xn are continuous (17.2)
for all n and therefore the metric
topology
is
finer
than
the
product
topology.
Q
Conversely, let B((xn ), ) Xn be an open ball in the metric topology. Then
B1 (x1 , ) BN (xN , ) XN +1 XN +2 B((xn ), )
when N is so big that 1/N < . As the left hand side is open in the product topology, this shows
that any set that is open in the metric topology is also open in the product topology.

An uncountable product of metric spaces need not be metrizable (17.10.(4)).
17.6. The first countability axiom. Let X be a topological space.
Definition 17.7. The space X has a countable basis at the point x X if there is a countable
collection {Un }nZ+ of neighborhoods of x such that any neighborhood U of x contains at least one
of the neighborhoods Un . The space X is first countable if all points of X have a countable basis.
Any metrizable space is first countable.
Proposition 17.8. [6, Thm 30.2] Any subspace of a first countable space is first countable. Any
countable product of first countable spaces is first countable.

GENERAL TOPOLOGY

35

Proof. The first assertion


Q is immediate. Let Xn be a countable product of first countable spaces.
Let
(x
)
be
a
point
of
Xn . Let Bn be countable basis at xn Xn . The collection of all products
n
Q
Un where Un Bn for finitely many n and Un = Xn for all other n is then a countable (5.5)
basis at (xn ).

An uncountable product of first countable (even metric) spaces and the quotient of a first
countable space may fail to be first countable (17.10.(4), 17.10.(5)). Some linearly ordered spaces
are first countable, some are not (17.10.(3)).
In a first countable (eg metric) space X, the points of the closure of any A X can be
approached by sequences from A in the sense that they are precisely the limit points of convergent
sequences in A. This is not true in general (17.10.(3)).
Lemma 17.9. Let X be a topological space.
(1) (The sequence lemma) Let A X be a subspace and x a point of X. Then
x is the limit of a sequence of points from A x A
The converse holds if X is first countable.
(2) (Continuous map preserve convergent sequences) Let f : X Y be a map of X into a space
Y . Then
f is continuous f (xn ) f (x) whenever xn x for any sequence xn in X
The converse holds if X is first countable.
Proof. (1) The direction is clear. Conversely, suppose that x A. Let Un be a countable basis
at x. For each n choose a point xn A U1 . . . Un . Then xm Un for all m n. We claim
that the sequence (xn ) converges to x. Let U be any neighborhood of x. Then Un U for some
n so that xm Un U for m n.
(2) The direction is clear. Conversely, suppose that f (xn ) f (x) whenever xn x. The
idea is to approximate points in the closure of A X by sequences, which we can do in the first
countable space X, to prove that f (A) f (A) (15.3.(6)) for any A X. Let x A. Since X is
first countable, there is by (1) a sequence of points xn A converging to x. By hypothesis, the
sequence f (xn ) f (A) converges to f (x). Thus f (x) A by (1) again.

We say that X satisfies the sequence lemma1 if for any A X and for any x A there is a
sequence of points in A converging to x.
To summarize:
X is metrizable X is first countable X satisfies the sequence lemma
Examples show that neither of these arrows reverse.
Example 17.10. (1) R is first countable as is any metric space.
(2) R` is first countable. The collection of half-open intervals [a, b) where b > a is rational is a
countable basis of neighborhoods at the point a. Is R` metrizable? (24.3) [6, Ex 30.6]
(3) S does not satisfy the sequence lemma (so is not first countable and not metrizable). The
largest element lies in the closure of S but it can not be the limit of any sequence of points
from S for any such such sequence has an upper bound in S . The section S is first countable:
{1} = [1, 2) is open so S has a finite basis at the first element 1. For any other element, > 1, we
can use the collection of neighborhoods of the form (, ] for < . Is S metrizable? (22.3)[6,
Exmp 3 p 181, Ex 30.7].
(4) RJ is not first countable when J is uncountable: Let {Un }nZ+ be any countable collection
of neighborhoods of, say, the point (0)jJ of RJ . The idea is to construct a neighborhood that
does not contain any of the Un . I claim that there is an index j0 J such that j0 (Un ) = R for
all n. Indeed, the set of js for which this is not true
[
{j J | n Z+ : j (Un ) 6= R} =
{j J | j (Un ) 6= R}
nZ+
1Aka a Frech
et space

36

GENERAL TOPOLOGY

is a countable union of finite sets, hence countable (5.5.(3)). Then the neighborhood j1
(R{1})
0
of (0) does not contain any of the neighborhoods Un in the countable collection. (RJ does not
even satisfy the sequence`
lemma [6, Exmp
W 2 p 133].)
`
(5) The quotient map p : nZ+ S 1 nZ+ S 1 is closed (16.15.(4)). The domain nZ+ S 1 =
W
Z+ S 1 R R2 is first countable (17.8) but the image nZ+ S 1 (16.15.(7)) is not: Let
{Un }nZ+ be any collection of saturated neighborhoods of Z+ {1}. The saturated neighborhood
U which at level n equals Un with one point deleted
W (cf Cantors diagonal argument (5.7)) does
not contain any of the Un . It follows (17.8) that nZ+ S 1 does not embed in R2 nor in any
other first countable space. For instance, the universal property of quotient spaces (16.14) gives
a factorization
`
Q
f
1
/ nZ S 1
(17.11)
nZ+ S
+
KKK
ss9
KKK
s
s
K
ss
p KKK
sss f
s
%
W
1
nZ+ S
of the continuous map f such that m (f |{n} S 1 ) is the identity function when m = n and the
constant function when m 6= n. The induced
Q map f is an injective continuous map. It can not be
an W
embedding for the countable product S 1 of circles is Q
first countable (17.8). The topology
on S 1 is finer than the subspace topology inherited from S 1 .
(6) The closed (16.15.(4)) quotient map R R/Z takes the first countable space R to a space
that is not first countable [4, 1.4.17] (at the point corresponding to Z). This can be seen by an
argument similar to the one from item (5): Let {Un }nZ be any countable collection of open
neighborhoods of Z R. Let U be the open neighborhood of Z such that U (n, n + 1), n Z,
equals Un (n, n + 1) with one point deleted. Then U does not contain any of the Un .
17.12. The uniform metric. Let Y be a metric space with a bounded metric d and let J be a
set. We shall discuss uniform convergence which is a metric, not a topological, concept.
Q
Definition 17.13. [6, pp 124,266] The uniform metric on Y J = jJ Y is the metric given by
d((xj ), (yj )) = sup{j J | d(xj , yj )}.
It is easy to see that this is indeed a metric.
Theorem 17.14. On RJ we have: product topology uniform topology box topology
Proof. Omitted.

The elements of Y J are functions f : J Y . Note that a sequence of functions fn : J Y


converges in the uniform metric to the function f : J Y if and only if
> 0N > 0j Jn Z+ : n > N d(fn (j), f (j)) <
We say that the sequence of functions fn : J Y converges uniformly to the function f : J Y .
Theorem 17.15 (Uniform limit theorem). (Cf [6, Thm 43.6]) Suppose that J is a topological space
and that Y is a metric space. The uniform limit of any sequence (fn ) of continuous functions
fn : J Y is continuous.
Proof. Well-known.


18. Connected spaces

Definition 18.1. The topological space X is connected if it cannot be represented as the union
X = X1 X2 of two disjoint open non-empty subsets X1 and X2 .
Two subsets A and B are separated if A B = = A B. This means that the two sets
are disjoint and neither contains a limit point of the other. Two disjoint open (closed) sets are
separated. If C A and D B and A and B are separated, then C and D are separated. A
separation of X consists of two separated non-empty subsets A and B with union X = A B.
Theorem 18.2. The following are equivalent:

GENERAL TOPOLOGY

(1)
(2)
(3)
(4)

37

X is connected
The only closed and open subsets of X are and X
If X = A B where A and B are separated then A = or B = .
Every continuous map X {0, 1} to the discrete space {0, 1} is constant.

Proof. (1) (2): If C is a closed and open subset of X then also the complement X C is closed
and open. Thus X = C (X C) is the union of two disjoint open subsets and therefore one of
these sets must be empty.
(2) (3): Suppose that X = A B where A and B are separated. Then A A for A does
not meet B. Thus A is closed. Similarly, B is closed. Therefore A is both closed and open. By
hypothesis, A is either the empty set or all of X.
(3) (4): We prove the contrapositive. Suppose that there exists a continuous map f : X {0, 1}
that is not constant. Then X = f 1 (0) f 1 (1) is the union of two non-empty separated subsets.
(4) (1): We prove the contrapositive. Suppose that X is non-connected. Then X = U1 U2
where U1 and U1 are disjoint, open, and non-empty. The map f : X {0, 1} given by f (U1 ) = 0
and f (U1 ) = 1 is a non-constant continuous map.

Theorem 18.3. The image under a continuous map of a connected space is connected.
Proof. We use the equivalence of (1) and (4) in Theorem 18.2. Let f : X Y be a continuous
map. If f (X) is not connected, there is a non-constant continuous map f (X) {0, 1} and hence
a non-constant continuous map X {0, 1}. So X is not connected.

Example 18.4. (1) R {0} = R R+ is not connected for it is the union of two disjoint open
non-empty subsets.
(2) We shall later prove that R is connected (19.3) and that the connected subsets of R are
precisely the intervals, rays, R, and .
(3) R` is not connected for [a, b) is a closed and open subset whenever a < b. In fact, any subset
Y of R` containing at least two points a < b is disconnected as Y [a, b) is closed and open but
not equal to or Y . (R` is totally disconnected.)
(4) Q is totally disconnected (and not discrete): Let Y be any subspace of Q containing at least
two points a < b. Choose an irrational number t between a and b. Then Y (t, ) = Y [t, )
closed and open but not equal to or Y .
(5) RK is connected [6, Ex 27.3].
(6) The Sierpinski space {0, 1} is connected.
A subspace of X is said to be connected if it is connected in the subspace topology. A subspace
of a connected space need obviously not be connected.
Lemma 18.5. A subspace of X is connected if and only if it is not the union of two non-empty
separated subsets of X.
Proof. Let Y X and suppose that Y = Y1 Y2 . Observe that
Y1 and Y2 are separated in Y Y1 and Y2 are separated in X
because ClY (Y1 ) Y2 = Y1 Y Y2 = Y1 Y2 by 14.7.(3). The lemma now follows immediately
from 18.2, the equivalence of (1) and (3).

Corollary 18.6. Suppose that Y X is a connected subspace. For every pair A and B of separated
subsets of X such that Y A B we have either Y A or Y B.
Proof. The subsets Y A and Y B are separated (since the bigger sets A and B are separated)
with union Y . By 18.5, one of them must be empty, Y A = , say, so that Y B.

Theorem 18.7. Let (Yj )jJ be an indexed family of connected subspaces of the space X. Suppose
that there is an index j0 J such that Yj and Yj0 are not separated for any j J. Then the union
S
jJ Yj is connected.
S
Proof. We have to prove that the subspace Y = jJ Yj is connected. We use the criterion of
Lemma 18.5. Suppose that Y = A B is the union of two separated subsets A and B of X. Each
of the sets Yj is by Corollary 18.6 contained in either A or B. Lets say that Yj0 is contained in A.

38

GENERAL TOPOLOGY

Then Yj must be contained in A for all j J for if Yj B for some j then Yj and Yj0 would be
separated. Therefore Y = A.

Corollary 18.8. The union of a collection of connected subspaces with non-empty intersection is
connected.
Proof. No two subspaces in the collection are separated.

Corollary 18.9. Let C be a connected subspace of X. If C Y C then Y is connected.


Proof. Apply Theorem 18.7 to the collection consisting of C and {y} for y Y C. Then C and
{y} are not separated since y C.

Corollary 18.10. Suppose that for any two points in X there is a connected subspace containing
both of them. Then X is connected.
Proof. Let x0 be some fixed point
x X, let Cx be a connected subspace
S of X. For each point T
containing x0 and x. Then X = Cx is connected since Cx 6= (18.8).

Theorem 18.11. Products of connected spaces are connected.
Proof. We prove first that the product X Y of two connected spaces X and Y is connected. In
fact, for any two points (x1 , y1 ) and (x2 , y2 ) the subspace X {y1 } {x2 } Y contains the two
points and this subspace is connected since (18.8) it is the union of two connected subspaces with
a point in common. Thus X Y is connected by Corollary 18.10.
Next, induction shows that the product of finitely many
connected spaces is connected.
Q
Finally [6, Ex 23.10], consider an arbitrary product Xj of connected spaces Xj , j J. Choose
a point xj in each of the spaces Xj (assuming
that all the spaces of the product are non-empty).
Q
For every finite subset F of J let CF Xj be the product of the subspaces Xj if j F and
{xj } if j 6 F . Since CF is connected
for each finite subset F and these subsets have the point
S
(xj )jJ in common, the union F FQCF , where F is the collection of all finite subsets of J, is
connected. This union is not all of
Xj but its closure is, so that Corollary 18.9 tells us that
S
Q

F F CF =
jJ Xj is connected.

GENERAL TOPOLOGY

39

19. Connected subsets of linearly ordered spaces


Recall that a subset C of a linearly ordered set is convex if a, b C [a, b] C.
Lemma 19.2. Let X be a linearly ordered space and C a subspace of X. Then
C is connected C is convex
Proof. Let C X a subspace that is not convex. Then there exist points a < x < b in X such
that a and b are in C and x is outside C. Since C (, x) (x, +) is contained in the union
of two separated subsets and meet both of them, C is not connected (18.6).

The converse to 19.2 is of course not true as for instance Z+ is convex but not connected (there
are gaps).
Recall that a linear continuum is a linearly ordered set with the least upper bound property
where (x, y) 6= for any two points x < y (no gaps). We shall now identify the linearly ordered
spaces that are connected. (You may read the following theorem and its proof as if we discussed
X = R only.)
Theorem 19.3. [2, Ex 7 p 382] Let X be a linearly ordered space. Then
X is connected in the order topology X is a linear continuum
The connected subsets of a linear continuum are intervals, rays, the whole space and the empty set.
Proof. : This is an exercise [6, Ex 24.4] [4, Problem 6.3.2] [2, Ex 7 p 382].
: Assume X is a linear continuum. We will show that for C X we have
C is connected C is convex
Thanks to 19.2 we only need to show that convex subsets are connected. It isSenough (18.8) to
show that all closed intervals [a, b] in X are connected (any convex set C = bC [a0 , b] is the
union of the closed intervals in C emanating from some fixed point a0 C). Suppose that [a, b] is
not connected. Then [a, b] = A B is the union of two separated sets A 6= and B 6= (18.5).
We may assume that b B. The sets A and B are closed and open in [a, b] (since they form a
separation of [a, b]) and closed in X (since (14.7) [a, b] is closed (11.1) in X). Let c be the least
upper bound of the bounded set A. Then c A since A is closed in X. (The least upper bound of
any bounded set always belongs to the closure of the set since otherwise it wouldnt be the least
upper bound.) In particular, c < b so A [a, b). But an element c of an open subset A of [a, b)
(with the subspace topology or, what is the same (13.5), the order topology) can not be an upper
bound for A because when c A also [c, e) A for some e > c and [c, e) is not empty when X is
a linear continuum. (In short, c A since A is closed in X and c 6 A since A is open in [a, b).)
We now know that the connected and the convex subsets of X are the same. In particular, the
linear continuum X itself, which certainly is convex, is connected in the order topology.
We finally show [2, Prop 1 p 336] that the non-empty convex subsets of X are the (bounded or
unbounded, closed, open, or half-open) intervals. Let C be a convex subset of X. If C is neither
bounded from above nor below it must be all of X, for if x is any point of X there exist a, b C
so that a < x < b and then also x C. If C is bounded from above but not from below, let c be
its least upper bound. Then for any x < c, x is neither a lower nor an upper bound for C so there
exist a, b C such that a < x < b c so that x C. Thus C is either (, c) or (, c]. The
argument is similar for the other cases. Note that X also has the greatest lower bound property
[6, Ex 3.13].

The real line R, the ordered square Io2 , the (ordinary) (half) line [1, ) = Z+ [0, 1), and the
long (half) line S [0, 1) [6, Ex 24.6, 24.12] are examples of linear continua.
Theorem 19.4. [Intermediate Value Theorem] Let f : X Y be a continuous map of a connected
topological space X to a linearly ordered space Y . Then f (X) is connected, hence convex. If Y is
a linear continuum, f (X) is an interval (bounded or unbounded, closed, half-open, or open)
Proof. f (X) is connected (18.3) and therefore convex (19.2). For subsets of a linear continuum we
know (19.3) that connected = convex = interval.


40

GENERAL TOPOLOGY

Any linearly ordered space containing two consecutive points, two points a and b with (a, b) =
is not connected as X = (, b) (a, +) is the union of two disjoint open sets.
Any well-ordered set X containing at least two points is totally disconnected in the order
topology. For if C X contains a < b then a 6 C (a, b] 3 b is closed and open in C since (a, b]
is closed and open in X.
19.5. Path connected spaces. Path connectedness is a stronger property than connectedness.
Definition 19.6. The topological space X is path connected if for any two points x0 and x1 in X
there is a continuous map (a path) f : [0, 1] X with f (0) = x0 and f (1) = x1 .
The image under a continuous map of a path connected space is path connected, cf 18.3. Any
product of path connected spaces is path connected [6, Ex 24.8], cf 18.7.
Example 19.7. The punctured euclidian plane Rn {0} is path connected when n 2 since any
two points can be joined by a path of broken lines. Thus also the (n 1)-sphere S n1 , which is
the image of Rn {0} under the continuous map x 7 x/|x|, is path connected for (n 1) 1.
Since the unit interval I is connected (19.3), all paths f (I) are also connected (18.3) so all path
connected spaces are, as unions of paths emanating from one fixed point, connected (18.10). The
converse is not true, not all connected spaces are path connected.
Example 19.8. (1) (Topologists sine curve) Let S be the graph of the function sin(1/x), 0 <
x 1, considered as a subspace of the plane R2 . The (closed) topologists sine curve is the
subspace S = S ({0} [1, 1]) of the plane. It follows from 18.3 and 18.9 that S is connected.
However, S is not path connected. For suppose that there existed a continuous map f : [0, 1] S
1
from f (0) = (0, 0) to some point in S, say f (1) = ( 2
, 0) S. We may arrange it so that
f (t) S for t > 0. For each positive integer n Z+ we have 2 f ([0, 1/n]) = [1, 1]. The
reason is that the image 1 f ([0, n1 ]) on the first axis is an interval (19.4) so that f maps [0, 1/n]
onto S (1 f ([0, 1/n]) R) and then 2 f ([0, 1/n]) = 2 (S (R 1 f ([0, 1/n])) = [1, 1].
This means that for each n Z+ there is a point tn [0, 1/n] such that the second coordinate
2 f (tn ) = (1)n . The sequence (tn ) converges to 0 but the image sequence 2 f (tn ) = (1)n
does not converge, contradicting continuity of f (17.9.(2)).
(2) The ordered square Io 2 (11.4.(7)) is connected since it is a linear continuum [6, Ex 3.15] but
it is not path connected. For suppose that f : [0, 1] Io2 is a path from the smallest element
(the lower left corner) f (0) = 0 0 to the largest element (the upper right corner) f (1) = 1 1.
Then f is surjective for the image contains (19.4) the interval [0 0, 1 1] = Io2 . But this is
impossible since the ordered square Io 2 contains uncountably many open disjoint subsets (eg
x (0, 1) = (x 0, x 1), 0 x 1) but [0, 1] does not contain uncountably many open disjoint
subsets (choose a rational number in each of them).
(3) RK is connected and not path-connected space [6, Ex 27.3].
19.9. Components and path components. The relations on X defined by
x y There is a connected subset C X such that x C and y C
p

x y There is a path in X between x and y


are equivalence relations. Check transitivity!
Definition 19.10. The equivalence classes of the first (second) equivalence relation are called
components (path components) of X.
The component containing the point x X is, by definition, the union of all connected subspaces
containing x. The path component of the point x X consists, by definition, of all points with a
path to x.
Theorem 19.11. (1) The components are connected closed disjoint subsets with union X. Any
connected subset
of X is a subset of precisely one component.
S
(2) If X = C where the C are connected subspaces such that for 6= there is a separation
of A B of X with C A and C B, then the C are the components of X.
(3) The path components of X are path connected disjoint subsets with union X. Any path connected subset is contained in precisely one of the path components.

GENERAL TOPOLOGY

41

(4) The path components are connected. The components are unions of path components.
Proof. (1) The components form a partition of X since they are equivalence classes. The components are closed since the closure of any connected subset is connected. All the points of a connected
subset are equivalent so they are contained in the same equivalence class.
(2) All points in C are equivalent. No point in C can be a equivalent to any point in C for
6= : Assume that there is a connected set C X that intersects both C and C . Choose a
separation X = A B such that C A and C B. Then C C C X = A B and the
connected (18.8) set C C C intersects both A and B, which is impossible (18.6).

If X has only finitely many components the components are closed and open. If X = C1 Cn
where the finitely many subspaces Ci are connected and separated then the Ci are the components
of X.
The path components can be closed, open, closed and open, or neither closed nor open. However,
for locally connected spaces (19.13) components and path components coincide (19.16).
19.12. Locally connected and locally path connected spaces.
Definition 19.13. The space X is locally (path) connected at the point x if every neighborhood of
x contains a (path) connected neighborhood of x. The space X is locally (path) connected if it is
locally (path) connected at each of its points.
Thus a space is locally (path) connected iff it has a basis of (path) connected subsets.
Consider this table
R
[1, 0) (0, 1]
S
Q
Connected
YES
NO
YES NO
Locally connected YES
YES
NO NO
before you draw any conclusions about the relationship between connected and locally connected
spaces. Note in particular that a space can be connected (even path connected 19.17) and not
locally connected.
On the positive side, note that open subsets of locally (path) connected spaces are locally (path)
connected and that quotient spaces of locally connected spaces are locally connected [6, Ex 25.8]
[2, I.11.6].
Proposition 19.14. The space X is locally (path) connected if and only if open subsets have open
(path) components (open in X or, what is the same ( 13.3), open in the open set).
Proof. Assume that X is locally connected and let U be an open subset. Consider a component
C of U . The claim is that C is open. Let x be a point in C. Choose a connected neighborhood
V of x such that V U . Since V is connected in X, V is connected in U so V is contained in a
component of U (19.11) and this component must be C; V C. This shows that C is open.
Conversely, assume that open subsets have open components. Let x be a point of X and U a
neighborhood of x. Let C be the component of U containing x. Then C is an open connected
neighborhood of x contained in U .

It follows for instance that the open subsets of R are unions of at most countably many open
intervals. For any open set is the union of its components which are open, since R is locally
connected, and connected, so they are open intervals. Since each of the open intervals contains a
rational number, there are at most countably many components. Any closed subset of R is the
complement to a union of at most countably many open intervals.
Proposition 19.15. [6, Ex 25.8] [2, I.11.6] Locally connected spaces have locally connected quotient spaces.
Is it true that locally path connected spaces have locally path connected quotient spaces?
Theorem 19.16. The components and the path components are the same in a locally path connected space.
Proof. The path components P of X are connected and open, since X is locally path connected
(19.14). Therefore (19.11.(2)) they are the components.


42

GENERAL TOPOLOGY

For instance, the components and the path components of a locally Euclidean space, such as a
manifold, are the same.
Example 19.17. (1) The comb space X = ([0, 1]{0})(({0}{1/n | n Z+ })[0, 1]) R2 is
clearly path connected but it is not locally connected for (19.16) the open set U = X [0, 1] {0}
has a component (use (19.11.(2)) to identify the components) that is not open. (The same is true
for any (small) neighborhood of (0, 1/2).) See [6, Ex 25.5] for a similar example.
(2) The closed topologists sine curve S has two path path-components, S = S (R+ R), which
is open and not closed for its closure is S, and {0} [0, 1], which is closed and not open for S
is not closed. Since there are two path components and one component, S is not locally path
connected (19.16). S is not even locally connected: Let U be a small neighborhood (well, not too
big) around (0, 1/2), then U has component that is not open. For instance, U = S ([0, 1] {0}
has a component (19.11.(2)), {0} (0, 1] that is not open (no neighborhood of (0, 21 ) is disjoint
from S). (The Warsaw circle is an example of a path connected not locally connected space, so
is the next example.)
S
(3) The space X = Cn (14.13.(2)) is even path connected but still not locally connected for the
component R+ {0} of the open set X {(0, 0)} (use (19.11.(2)) to identify the components)
is not open (any neighborhood of (1, 0) contains points from the circles). Alternatively, X is not
locally connected since any small neighborhood of (1, 0) has a separation.
(4) Any linear continuum is locally connected since the basis for the topology consists of connected
sets (19.3). It need not be locally path connected [6, Ex 25.3].

GENERAL TOPOLOGY

43

20. Compact spaces


Definition 20.2. A topological space X is compact if it satisfies one of the following two equivalent
conditions:
(1) Every collection of open subsets whose union is X contains a finite subcollection whose
union is X.
(2) Every collection of closed subsets whose intersection is contains a finite subcollection
whose intersection is .
A collection of open subsets whose union is the whole space is called an open covering. Thus a
space is compact if any open covering contains a finite open covering.
Any finite space is compact as there are only finitely many open subsets. Any compact discrete
space is finite. The real line R is not compact as the open covering {(n, n) | n Z+ } has no
finite subcovering.
Theorem 20.3. The image under a continuous map of a compact space is compact.
Proof. Let f : X Y be continuous. Let A be an open covering of f (X). Then f 1 (A) =
{f 1 (A) | A A} is an open covering of X. Assuming that X is compact, there is a finite subset
B of A such that f 1 (B) is a covering of X. Then B is a finite covering of f (X).

Theorem 20.4. Every closed subset of a compact space is compact.
Proof. The theorem follows immediately from the second definition of compactness (20.2.(2)) when
we observe that closed subsets of a closed subset are closed.

Proposition 20.5. Let Y X. Then
Y is compact
Any collection of open sets (in X) that cover Y contains a subcollection that covers Y
S
Proof. : Let A = {Aj | j J} be a collection of open sets Aj X such that Y Aj . Then
{Y Aj | j J} is an open covering of Y . Since Y is compact, there is a finite subcovering of Y .
The corresponding finitely many sets of the collection A then cover Y .
: Let B = {Bj | j J} be an open covering of Y . Then Bj = Y Aj for some open set Aj X.
The collection {Bj | j J} covers Y and so it contains a finite subcollection that covers Y . The
corresponding subcollection of B then also covers Y .

Lemma 20.6. Let X be a Hausdorff space and let C X, x0 X be a compact subset and a
point such that x0 6 C. Then there exist open sets U , V such that U V = and C U and
x0 V .
Proof. For each point a C, the Hausdorff property implies that there are open sets Ua , Va such
that U V = and a Ua and x0 Va . Since C is compact (20.5), C Ua1 . . . Uak for finitely
many points a1 , . . . , ak C. Set U = Ua1 . . . Uak and V = Va1 . . . Vak .

Theorem 20.7. Every compact subset of a Hausdorff space is closed.
Proof. This is an immediate consequence of 20.6.

Corollary 20.8. Let X be a compact Hausdorff space.


(1) Let C X a subset. Then C is compact C is closed.
(2) If A and B are disoint closed sets in X then there exist disjoint open sets U , V such that
A U and B V .
Proof. (1) is 20.7 and is 20.4.
(2) Let A and B be disjoint closed subsets of X. Then A and B are compact as just shown. Now
apply 20.6 and compactness to find disjoint open sets U A, V B.

Corollary 20.9.
(1) Every continuous map of a compact space into a Hausdorff space is closed
(and a quotient map onto its image).
(2) Every injective continuous map of a compact space into a Hausdorff space is an embedding.

44

GENERAL TOPOLOGY

(3) Every bijective continuous map of a compact space onto a Hausdorff space is a homeomorphism.
Proof. Let f : X Y be a continuous map of a compact space X into a Hausdorff space Y .
(1) We have
20.4

20.3

20.7

C is closed in X C is compact f (C) is compact f (C) is closed in Y


which shows that f is a closed map.
(2 Every injective continuous and closed map is an embedding.
(3) Every bijective continuous and closed map is a homeomorphism.

Example 20.10. The (unreduced) suspension of the space X is the quotient space
SX = [0, 1] X/({0} X, {1} X)
What is the suspension of the n-sphere S n ? Define f : [0, 1] S n S n+1 to be the (continuous)
map that takes (t, x), t [0, 1], to the geodesic path from the north pole (0, 1) S n+1 , through
the equatorial point (x, 0) S n S n+1 , to the south pole (0, 1) S n+1 . (Coordinates in
Rn+2 = Rn+1 R.) By the universal property of quotient spaces there is an induced continuous
and bijective map f : SS n S n+1 . Since SS n is compact (as a quotient of a product of two
compact spaces) and S n+1 is Hausdorff (as a subspace of a Hausdorff space), f is a homeomorphism
(20.9). We often write SS n = S n+1 , n 0, where the equality sign stands for is homeomorphic
to.
Corollary 20.11. The Hausdorff image of a locally connected compact space is locally connected
compact.
Proof. We now know that f is a quotient map. The quotient image of a locally connected space is
locally connected (19.15).

Lemma 20.12 (The tube lemma). Let X and Y be topological spaces where Y is compact. Let x0
be a point of X. For any neighborhood N X Y of the slice {x0 } Y there is a neighborhood
U of x0 such that {x0 } Y U Y N .
Proof. For each point y Y there is a product neighborhood such that x0 y Uy Vy N .
Since Y is compact, there are finitely many points y1 , . . . , yk Y such that Y = V1 . . . Vk where
Vi = Vyi . Put U = U1 . . . Uk .

Theorem 20.13. (Cf 20.17) The product of finitely many compact spaces is compact.
Proof. We show that the product X Y of two compact spaces X and Y is compact. The theorem
follows from this by induction. Suppose that A is an open covering of X Y . We show that there
is a finite subcovering. ForSeach point x X, the slice {x} Y , homeomorphic to Y , is compact
and therefore {x} Y {A | A A(x)}
S for some finite set A(x) A. Thanks to the tube
lemma (20.12) we know that the open set {A | A A(x)} actually contains whole tube U (x) Y
for some neighborhood U (x) of x. By compactness, X = U (x1 ) . . . U (xk ) can be covered by
finitely many of the neighborhoods U (x). Now
X Y =

k
[

U (xi ) Y =

i=1

where

Sk

i=1

k [
[

{A | A A(xi )} =

i=1

{A | A

k
[

A(xi )}

i=1

A(xi ) A is finite (4.8).

Here are two small lemmas that are used quite often.
Lemma 20.14. A compact space can not contain an infinite closed discrete subspace.
Proof. Any closed discrete subspace of a compact space is compact (20.4) and discrete, hence
finite.

Lemma 20.15. [6, Ex 28.5] [4, 3.10.2] Let C1 C2T Cn be a descending sequence of
closed nonempty subsets of a compact space. Then Cn 6= .

GENERAL TOPOLOGY

Proof. If

Cn = then Cn = for some n Z+ by 20.2.(2).

45

Theorem 20.16. [Cf [6, Ex 27.5]] Any nonempty compact Hausdorff space without isolated points
( 14.11) is uncountable.
Proof. Let xn be a sequence of points in X. It suffices (5.4) to show that {xn | n Z+ } =
6 X.
Since X is Hausdorff and has no isolated points there is a descending sequence
V1 V2 Vn1 Vn
of nonempty open sets such that xn 6 V n for all n. These are constructed recursively. Put V0 = X.
Suppose that Vn1 has been constructed for some n Z+ . Since xn is not isolated, {xn } =
6 Vn1 .
Choose a point yn Vn1 distinct from xn and choose disjoint (separated) open sets Un X,
Vn Vn1 such that xn Un and ynT Vn . Then Un V n = so xn 6 V n .
By construction, the
T intersection V n contains none of the points {xn } of the sequence and by
compactness (20.15), V n 6= . Thus X contains a point that is not in the sequence.

Is it true that a connected Hausdorff space is uncountable?
Q
Theorem 20.17 (Tychonoff theorem). The product jJ Xj of any collection (Xj )jJ of compact
spaces is compact.
Q
Proof. Put X = jJ Xj . Let us say that a collection of subsets of X is an FIP-collection (finite
intersectionTproperty) if any finite subcollection has nonempty intersection. We mustT(20.2.(2))
show that A 6= for any FIP-collection A of closed subsets. We show that in fact {A | A
A} =
6 forTany FIP-collection A (of not necessarily closed sets). This means that we must find a
point x in AA A.
Step 1. The FIP-collection A is contained in a maximal FIP-collection.
The set (supercollection?) A of FIP-collections containing A is a strictly partially ordered by strict
inclusion. Any linearly
subset (subsupercollection?) B A has an upper bound, namely
S ordered
S
the FIP-collection B = BB B containing A. Now Zorns lemma (8.4) says that A has maximal
elements.
We can therefore assume that A is a maximal FIP-collection.
Step 2. Maximality of A implies
(20.18)

A1 , . . . , Ak A = A1 Ak A

and for any subset A0 X we have


(20.19)

A A : A0 A 6= = A0 A

To prove (20.18), note that the collection A {A1 Ak } is FIP, hence equals A by maximality.
To prove (20.19), note that the collection A {A0 } is FIP: Let A1 , . . . , Ak A. By (20.18),
A1 Ak A so A1 Ak A 6= by assumption.
Step 3. {j (A) | A A} is FIP for all j J.
j (A1 ) j (Ak ) j (A1 Aj ) 6= for any finite collection of sets A1 , . . . , Ak A.
Since Xj is compact and {j (A) | A A} is an FIP collection of closed subsets of Xj , the
T
T
intersection AA j (A) Xj is nonempty (20.2.(2)). Choose a point xj j (A) for each j J
and put x = (xj ) X.
T
Step 3. x AA A.
Note first that j1 (Uj ) A for any neighborhood Uj of xj . This follows from (20.19) since
j1 (Uj ) A 6= , or equivalently, Uj j (A) 6= , for any A A as xj j (A). Since A has
the FIP it follows that j1
(Uj1 ) j1
(Ujk ) A 6= whenever Uji are neighborhoods of the
1
k
T
finitely many points xji Xji and A A. Thus x AA A.

21. Compact subspaces of linearly ordered spaces
Lemma 21.1. Any compact subset C 6= of a linearly ordered space Y has a smallest and a
largest element.

46

GENERAL TOPOLOGY

Proof. The proof is by contradiction. Assume that C has no largest element. This means that
for any a C there is a c C such that a < c or, in other words, that C cC (, c). By
compactness
C (, c1 ) (, ck ) (, c)
where c = max{c1 , . . . , ck } is the greatest (7.6) of the elements c1 , . . . , ck C. But this is a
contradiction as c C and c 6 (, c).

We shall now identify the compact intervals of the real line R. The result says that all closed
and bounded intervals [a, b] R are compact. The argument works just as well in any linearly
ordered space X with the least upper bound property (1.14) so for the sake of generality we shall
work in this context.
Theorem 21.2. Let X be a linearly ordered space. Then
X has the least upper bound property Every closed and bounded interval in X is compact
Proof. : [6, Ex 27.1].
: Let [a, b] X be a closed interval and A and open covering of [a, b] (with the subspace topology
which is the order topology). We must show that [a, b] is covered by finitely many of the sets from
the collection A. The set
C = {y [a, b] | [a, y] can be covered by finitely many members of A}
is nonempty (a C) and bounded from above (by b). Let c = sup C be the least upper bound of
C. Then a c b. We would like to show that c = b.
Step 1 If x C and x < b then y C for some y (x, b]
Proof of Step 1. Suppose first that x has an immediate successor y > x. We can not have y > b > x
for then y would not be an immediate successor. So x < y b. Clearly [a, y] = [a, x] {y} can be
covered by finitely many members of A. Suppose next that x has no immediate successor. Choose
an open set A A containing x. Since A is open in [a, b] and contains x < b, A contains an
interval of the form [x, d) for some d b. Since d is not an immediate successor of x there is a
point y (x, d). Now [a, y] = [a, x] [x, y] [a, x] [x, d) [a, x] A can be covered by finitely
members of A.
Step 2 c C.
Proof of Step 2. The claim is that [a, c] can be covered by finitely many members of A. From Step
1 we have that C contains elements > a so the upper bound c is also > a. Choose A A such
that c A. Since A is open in [a, b] and a < c, A contains an interval of the form (d, c] for some d
where a d < c. Since c is the least upper bound for C, there must be points from C in (d, c]. Let
z be such a point. Now [a, c] = [a, z] (d, c] where [a, z] can be covered by finitely many members
of A and (d, c] is contained in A. Thus [a, c] can be covered by finitely many members of A.
Step 3 c = b.
Proof of Step 3. Since c C by Step 2, Step 1 says that if c < b, then c can not be an upper
bound.

For instance, the ordered square Io2 = [0 0, 1 1] and S = [1, ] are compact linearly ordered
spaces [6, Ex 10.1]. More importantly, since the linearly ordered space R has the least upper bound
property, we get
Corollary 21.3. The compact and connected subsets of a linear continuum are precisely the closed
and bounded intervals.
Proof. Let X be a linear comtinuum. All closed and bounded intervals in X are compact and
connected (21.2, 19.3). Conversely, if C X is compact then C [a, b] for some a, b in C (21.1)
and if C is also connected then C is convex (19.3) so [a, b] C.

In particular, the compact and connected subsets of R are precisely the closed and bounded
intervals.
Theorem 21.4 (Extreme Value theorem). Let f : X Y be a continuous map of a topological
space X into a linearly ordered space Y .
(1) If X is compact then there exist m, M X such that f (m) f (x) f (M ) for all x X.

GENERAL TOPOLOGY

47

(2) If X is compact and connected then there exist m, M X such that f (X) = [f (m), f (M )].
Proof. (1) Since the image f (X) is compact (20.3) it has (21.1) a smallest and a greatest element,
f (m) f (X) and f (M ) f (X), so that f (X) [f (m), f (M )].
21.1

19.2

(2) f (X) [f (m), f (M )] f (X) since f (X) is compact and connected (20.3, 18.3).

21.5. Compactness in metric spaces.


Lemma 21.6 (Lebesgue lemma). Let X be a compact metric space. For any open covering A of
X there exists an > 0 (the Lebesgue number of the open covering A) such that any subset of X
with diameter < is contained in a member of A.
Proof. For each point x X, choose x > 0 such that the ball B(x, 2x ) is contained in a member
of A. The collection {B(x, x )}xX covers X so by compactness
X = B(x1 , x1 ) B(x2 , x2 ) B(xk , xk )
for finitely many points x1 , x2 , . . . , xk X. Let be the smallest of the numbers 1 , 2 , . . . , k .
The triangle inequality implies that this works. (Suppose that A X with diam A < . Choose
any point a in A. There is an i such that d(xi , a) < . Let b be any point in A. Since d(b, xi )
d(b, a) + d(a, xi ) < + i 2i , the point b is contained in B(xi , 2i ) which is contained in a
member of A.)

Definition 21.7. A map f : X Y between metric spaces is uniformly continuous if for all > 0
there is a > 0 such that dY (f (x1 ), f (x2 )) < whenever dX (x1 , x2 ) < :
> 0 > 0x1 , x2 X : dX (x1 , x2 ) < dY (f (x1 ), f (x2 )) <
Theorem 21.8 (Uniform continuity theorem). Let f : X Y be a continuous map between metric
spaces. If X is compact, then f is uniformly continuous.
Proof. Given > 0. Let be the Lebesgue number of the open covering {f 1 B(y, ) | y Y }.
Then we have for all x1 , x2 X
dX (x1 , x2 ) < diam{x1 , x2 } <
y Y : {x1 , x2 } f 1 B(y, )
y Y : {f (x1 ), f (x2 )} B(y, )
dY (f (x1 ), f (x2 )) < 2

The following is a theorem that applies not to metric spaces in general but rather to the special
metric space Rn .
Theorem 21.9 (HeineBorel). Consider the metric space Rn with its standard metric. A subset
C Rn is compact if and only if it is closed and bounded.
Proof. Let C Rn be compact. Since C is compact and {(R, R)n | R > 0} an open covering of
Rn , there is an R > 0 such that C [R, R]n . Thus C is bounded. As a compact subset of a
Hausdorff space, C is closed (20.7).
Conversely, if C is closed and bounded then C is a closed subset of [R, R]n for some R > 0.
But [R, R]n is compact (21.3, 20.13) and closed subsets of compact spaces are compact (20.4);
in particular, C is compact.

The theorem is not true if we replace the standard metric with the bounded metric associated
to the standard metric.
Can you now answer question 15.8.(9)?

48

GENERAL TOPOLOGY

22. Limit point compactness and sequential compactness


Definition 22.1. A space X is
(1) limit point compact if any infinite subset of X has a limit point
(2) sequentially compact if any sequence in X has a convergent subsequence.
In other words, a space is limit point compact if it contains no infinite closed discrete subspaces
(14.12).
We shall not say much about these other forms of compactness (see [4, 3.10] for a more thorough
discussion).
Theorem 22.2. (Cf [6, Ex 28.4, Ex 35.3]) For any topological space X we have
X is compact = X is limit point compact

X 1st countable

X is sequentially compact

All three forms of compactness are equivalent for a metrizable space.


Proof. X is compact = X is limit point compact: A subset with no limit points is closed and
discrete (14.12), hence finite (20.14).
X is limit point compact and 1st countable = X is sequentially compact: Let (xn ) be a sequence
in X. Consider the set A = {xn | n Z+ }. If A is finite there is a constant subsequence (4.8.(3)).
If A is infinite, A has a limit point x by hypothesis. If X is 1st countable, then x is (17.9) the limit
of a sequence of points from A. By rearranging, if necessary, we get (!) a subsequence converging
to x.
X is sequentially compact and metrizable = X compact: This is more complicated (but should
also be well-known from your experience with metric spaces) so we skip the proof.

Example 22.3. (1) The ordered space S of all countable ordinals is limit point compact and
sequentially compact but it is not compact. S is not compact (21.1) for it is a linearly ordered
space with no greatest element [6, Ex 10.6]. On the other hand, any countably infinite subset
of S is contained in a compact subset (7.9.(3), 21.2). Therefore (22.2) any countably infinite
subset, indeed any infinite subset (14.12), has a limit point and any sequence has a convergent
subsequence. (Alternatively, use that S is first countable (17.10.(3)).) It follows (22.2) that S
is not metrizable.

(2) The StoneCech


compactification Z+ of the positive integers is compact but it is not limit
point compact [6, Ex 38.9] since no point of the remainder Z+ Z+ is the limit of a sequence
in Z+ (so Z+ is not first countable and not metrizable).
23. Locally compact spaces and the Alexandroff compactification
Definition 23.1. A space X is locally compact at the point x X if there exist an open set
U X and a compact set C X such that x U C. X is locally compact if X is locally
compact at each of its points.
Compact spaces are locally compact (U = X = C). The real line R (more generally, Rn ) is
locally compact but not compact. All linearly ordered spaces with the least upper bound property
are locally compact. The space Q R of rational numbers is not locally compact [6, Ex 29.1].
Let X be a locally compact Hausdorff space. Let X = X {} denote the union of X with a
set consisting of a single point . The collection
T = {U | U X open} {X C | C X compact}
is a topology on X: It is easy to see that finite intersections and arbitrary unions of open sets of
the first (second) kind are again open of the first (second) kind. It follows that T is closed under
arbitrary unions. It is also closed under finite intersections since U (X C) = U C is open
in X. And X = X and are open.
Theorem 23.2 (Alexandroff). Let X be a locally compact Hausdorff space. Then X is a compact
Hausdorff space and the inclusion map X X is an embedding.

GENERAL TOPOLOGY

49

If c : X cX is another embedding of X into a compact Hausdorff space cX whose image is


cX but one point then there is a unique homeomorphism cX X such that the diagram

cX

|
||
||
|
|} |

XC
CC
CCc
CC
C!
/ X
'

commutes.
Proof. We must verify the following points:
The subspace topology on X is the topology on X: The subspace topology X T is clearly the
original topology on X.
X is Hausdorff: Let x1 and x2 be two distinct points in X. If both points are in X then there
are disjoint open sets U1 , U2 X X containing x1 and x2 , respectively. If x1 X and x2 = ,
choose an open set U and a compact set C in X such that x U C. Then U 3 x and X C 3
are disjoint open sets in X.
X is compact: Let {Uj }jJ be any open covering of X. At least one of these open sets contains
. If Uk , k J, then Uk = (X C) {} for some compact set C X. There is a finite set
K J such that {Uj }jK covers C. Then {Uj }jK{k} is a finite open covering of X = X {}.
Let now c : X cX be another embedding of X into a compact Hausdorff space cX such that
cX X = { 0 } consists of just one point 0 . Define f : cX X by f (x) = x for all x X
and f ( 0 ) = . For any open set U X, f 1 (U ) is open in X and hence in cX since X is open
in cX (13.3.(2)). For any compact set C X, f 1 (X C) = cX C is open in cX since the
compact set C is closed in the Hausdorff space cX (20.8.(1)). This shows that f is continuous. By
construction, f is bijective and hence (20.9) a homeomorphism.

The space X is called the Alexandroff or the one-point compactification of X.
Example 23.3. (1) If X is compact Hausdorff then X is X together with an isolated point .
({} = X X is a neighborhood of .)
(2) The n-sphere S n = Rn is the Alexandroff compactification of the locally compact Hausdorff
space Rn for there is (15.8.(7)) a homeomorphism of Rn onto the complement of a point in S n .
(3) Let X be a linearly ordered space with the least upper bound property and let [a, b) be
a half-open interval in the locally compact space X. Then [a, b) = [a, b]. For instance the
Alexandroff compactifications of the half-open intervals [0, 1) R, Z+ = [1, ) Z+ Z+ , and
S = [1, ) S Sare [0, 1], [1, ], [1, ] = S , respectively.
(4) (Z+ R) = nZ+ C1/n , the Hawaiian Earring (14.13.(2)).
Corollary 23.4. Let X be a topological space. The following conditions are equivalent:
(1) X is locally compact Hausdorff
(2) X is homeomorphic to an open subset of a compact Hausdorff space
(3) X is Hausdorff and for any point x X and any neighborhood U 3 x there is a neighborhood V 3 x such that V U and V is compact.
Proof. (1) = (2): Use 23.2.
(2) = (3): Suppose that Y is a compact Hausdorff space and that X Y is an open subset. X
is Hausdorff since subspaces of Hausdorff spaces are Hausdorff (14.19). Let x be a point of X and
U X a neighborhood of x. Since Y U is compact (20.8) and does not contain x, there are
(20.6) disjoint open set V 3 x and W Y U . Then x V Y W U . Here V Y W U
since Y W is closed and V is compact since it is closed (20.8).
(3) = (1): Take x U = X and find V X such that C = V is compact.

Corollary 23.5. Any closed subset of a locally compact space is locally compact. Any open or
closed subset of a locally compact Hausdorff space is locally compact Hausdorff.
Proof. Let X be a locally compact space and A X a closed subspace. We use the definition
directly to show that A is locally compact. Let a A. There are subsets a U C X such
that U is open and C is compact. Then A U A C where A U is a neighborhood in A and
C A is compact as a closed subset of the compact set C (20.4).

50

GENERAL TOPOLOGY

If X is locally compact Hausdorff and A X open, it is immediate from 23.4 that A is locally
compact Hausdorff.

An arbitrary subspace of a locally compact Hausdorff space need not be locally compact
(23.6.(1)). The product of finitely many locally compact spaces is locally compact but an arbitrary product of locally compact spaces need not be locally compact [6, Ex 29.2]; for instance
Z
+ is not locally compact. The image of a locally compact space under an open continuous [6,
29.3] or a perfect map [6, Ex 31.7] [4, 3.7.21] is locally compact but the image under a general
continuous map of a locally compact space need not be locally compact; indeed, the quotient of a
locally compact space need not be locally compact (23.6.(2)) [4, 3.3.16].
S
Example 23.6. (1) nZ+ Cn R2 (14.13.(2)) is not locally compact at the origin: Any neighborhood of 0 0 contains a countably infinite closed discrete subspace so it can not be contained
in any compact subspace (20.14).
(2) The quotient space R/Z+ (17.10.(6)) is not locally compact at the point corresponding to
Z+ : Any neighborhood of this point contains an infinite closed discrete subspace so it can not be
contained in any compact subspace`(20.14).
Q 1
W
(3) In diagram (17.11), the space
S 1 is locally compact,
S is compact, and S 1 is not
locally compact (at the one point common to all the circles).
(4) (Wedge sums and smash products) A pointed space is a topological space together with one
of its points, called the base point. The wedge sum of two pointed disjoint spaces (X, x0 ) and
(Y, y0 ) is the quotient space
X Y = (X q Y )/({x0 } {y0 })
obtained from the disjoint union of X and Y (12.6) by identifying the two base points. Let
f : X Y X Y be the continuous map given by f (x) = (x, y0 ), x X, and f (y) = (x0 , y),
y Y . The image f (X Y ) = (X {y0 }) ({x0 } Y ) = 21 ({y0 }) 11 ({x0 }) is closed when
X and Y are T1 spaces. In fact, f is a closed map for if C X is closed then f (C) = {(x, y0 ) |
x C} = 11 (C) 21 ({y0 }) is also closed. Thus f is a quotient map onto its image and its
factorization f as in the commutative diagram (16.14.(2))
/ X Y
X YI
:
II
uu
II
u
u
II
u
II
uu f
$
uu
X Y
f

is an embedding. We can therefore identify X Y and X {y0 } {x0 } Y . The smash product
is defined to be
X Y = (X Y )/(X Y )
the quotient of the product space X Y by the closed subspace X Y ( = ).
If A X and B Y are closed subspaces, the universal property of quotient maps (16.14.(2))
produces a continuous bijection g such that the diagram
g

gX gY
/ X/A Y /B X/AY /B / X/A Y /B
X Y R
RRR
k5
RRR
kkkk
k
k
RRR
k
RRR
kkk g
R)
kkkk
(X Y )/(X B A Y )

commutes. However, g may not be a homeomorphism since the product gX gY of the two closed
quotient maps gX : X X/A and gY : Y Y /B may not be a quotient map. If X/A and Y are
locally compact Hausdorff spaces ([Ex 31.7] may be useful here) then gX gY is quotient [Ex
29.11] so that g is a homeomorphism (16.7, 16.14.(3), 16.6) and
(23.7)

(X Y )/(X B A Y ) = X/A Y /B

in this case. Here is an application. The nsphere S n is (homeomorphic to) the quotient space
I n /I n . Using the formula [Exam June 2003] for the boundary of the product of two sets we find

GENERAL TOPOLOGY

51

that the smash product of an msphere with an nsphere


(23.7)

S m S n = I m /I m I n /I n = I m I n /(I m I n I m I n ) = I m+n /I m+n = S m+n


is an (m + n)sphere.
(5) (The Alexandroff compactification of a product space) If X and Y are locally compact Hausdorff spaces the map X Y X Y X Y is an embedding. This follows from (16.11,
16.6) when we note that the first map embeds X Y into an open saturated subset of X Y
(15.9). The universal property of the Alexandroff compactification (23.2) implies that
(23.8)

(X Y ) = X Y

for any two locally compact Hausdorff spaces X and Y . Once again we get
(23.8)

S m S n = Rm Rn = (Rm Rn ) = Rm+n = S m+n


when we view the spheres as Alexandroff compactifications of euclidian spaces.
The locally compact Hausdorff space X is open in X. This is actually true whenever X is
embedded as a dense subspace of a Hausdorff space.
Proposition 23.9. Let X be a locally compact and dense subspace of the Hausdorff space Y . Then
X is open.
Proof. Let x X. Since X is locally compact Hausdorff the point x has a neighborhood X U ,
where U is open in X, such that its relative closure
14.7

14.9

ClX (X U ) = X X U = X U
is compact and hence closed in the Hausdorff space Y (20.7). But no part of U can stick outside
X U for since U (X U ) is open
X dense

U (X U ) 6= = (X U ) (X U ) 6=
which is absurd. Thus we must have U X U and of course X U X so U X. Therefore
X is open.

Some mathematicians prefer to include Hausdorffness in the definition of (local) compactness.
What we call a (locally) compact Hausdorff space they simply call a (locally) compact space; what
we call a (locally) compact space they call a (locally) quasicompact space.

52

GENERAL TOPOLOGY

24. Countability axioms


Let us recall the basic question: Which topological spaces are metrizable? In order to further
analyze this question we shall look at a few more properties of topological spaces.
We have already encountered the first countability axiom. There are three more axioms using
the term countable. Here are the four countability axioms.
Definition 24.1. A topological space
that has a countable neighborhood basis at each of its points is called a first countable space
that has a countable basis is called a second countable space
that contains a countable dense subset is said to contain a countable dense subset1
in which every open covering has a countable subcovering is called a Lindelof space
A subset A X is dense if A = X, that is if every nonempty open subset contains a point from
A.
Any second countable space is first countable. Any subspace of a (first) second countable space
is (first) second countable. A countable product of (first) second countable spaces is (first) second
countable, see 17.8.
The real line R is second countable for the open intervals (a, b) with rational end-points form
a countable basis for the topology. The real line R` with the right half-open interval topology is
first but not second countable (25.4).
All metric spaces are first countable but not all metric spaces are second countable [6, Example
2 p 190].
Any compact space is Lindel
of. Any closed subset of a Lindelof space is Lindelof [6, Ex 30.9]
(with a proof that is similar to that of 20.4). A product of two Lindelof spaces need not be Lindelof
(25.5).
The next result implies that R, in fact any subset of the 2nd countable space R, is Lindelof.
Theorem 24.2. Let X be a topological space. Then
X has a countable dense subset ks
X is 2nd countable

+3 X is Lindel
of


X is 1st countable
If X is metrizable, the three conditions of the top line equivalent.
Proof. Suppose first that X is second countable and let B be a countable basis for the topology.
Pick a point bB B in each basis set. The image of the map B X : B 7 bB is dense. Let now
U be an open covering of X. For each basis set B that is contained in some member of U pick a
UB U such that B UB . Then the image of the map (from a subset of) B U : B UB is an
open covering: Let x be any point in X. Since x is contained in a member of U and every open
set is a union of basis sets we have x B U for some basis set B and some U U. But then
also x B UB .
To prove the arrows in the other direction, let X be a metric space with a countable dense
subset A X. Then the collection {B(a, r) | a A, r Q+ } of balls centered at points in A and
with a rational radius is a countable (5.5.(4)) basis for the topology: It suffices to show that for
any y B(x, ) there are a A and r Q+ such that y B(a, r) B(x, ). Let r be a positive
rational number such that 2r < d(x, y) and let a A B(y, r). Then y B(a, r), of course,
and B(a, r) B(x, ) for if d(a, z) < r then d(x, z) d(x, y) + d(y, z) d(x, y) + d(y, a) + d(a, z) <
d(x, y) + 2r < .
Finally, let X be a metric Lindel
of S
space. For each positive rational
number r, let Ar be a
S
countable subset of X such that X = aAr B(a, r). Then A = rQ+ Ar is a dense countable
(5.5.(3)) subset: Any open ball B(x, ) contains a point of Ar when 0 < r < , r Q.

Example 24.3. R` has a dense countable subset and is not second countable (25.4). The ordered
square Io2 is Lindel
of and is not second countable since it contains uncountably many disjoint open
sets (x 0, x 1), x I. Thus R` and Io2 are not metrizable [6, Ex 30.6].
1Or to be separable

GENERAL TOPOLOGY

53

25. Separation Axioms


Definition 25.1. A space X is called a
T1 -space if points {x} are closed in X
T2 -space or a Hausdorff space if for any pair of distinct points x, y X there exist disjoint
open sets U and V such that x U and y V
T3 -space or a regular space if points are closed and for every closed set B X such that
x 6 B there exist disjoint open sets U and V such that x U and B V
T4 -space or a normal space if points are closed and for every par of disjoint closed sets
A, B X there exist disjoint open sets U and V such that A U and B V .
We have the following sequence of implications
X is normal = X is regular = X is Hausdorff = X is T1
where none of the arrows reverse (25.5, 25.6, [6, Ex 22.6]).
Lemma 25.2. Let X be a T1 -space. Then
(1) X is regular For every point x X and every neighborhood U of x there exists an
open set V such that x V V U .
(2) X is normal For every closed set A and every neighborhood U of A there is an open
set V such that A V V U .
Proof. Let B = X U .

Theorem 25.3. (Cf 14.19) Any subspace of a regular space is regular. Any product of regular
spaces is regular.
Proof. Let X be a regular space and Y X a subset. Then Y is Hausdorff (14.19)). Consider
a point y Y and a relatively closed set B Y such that y 6 B. Then B = ClY B = B Y
so y 6 B. By definition of regularity, there exist disjoint open sets U and V such that y U
and B V . The relatively open sets U Y and V Y are disjoint and they contain y and B,
respectively. Q
Let X =
Xj be the Cartesian product of regular space Xj , j J. Then X is Hausdorff
Q
(14.19)). We use 25.2.(1) to show that X is regular. Let x = (xj ) be a point in X and U = Uj
a basis neighborhood of x. Put Vj = Xj whenever Uj = Xj . Otherwise, choose Vj such that
Q
Vj is a neighborhood of x in the product topology and
xj Vj V j Uj . Then V =
Q
Q
V = V j Uj = U . Thus X is regular.

A subspace of a normal space need not be normal (28.5) and the product of two normal spaces
need not be normal ((25.5),[6, Example 2 p 203], [4, 2.3.36]). Any closed subspace of a normal
space is normal [6, Ex 32.1].
Example 25.4 (Sorgenfreys half-open interval topology). The half-open intervals [a, b) form a
basis for the space R` (10.6, 14.2.(4), 17.10.(2), 18.4.(3)).
R` is 1st countable: At the point x the collection of open sets of the form [x, a) where
a > x is rational, is a countable local basis at x.
R` is not 2nd countable: Let B be any basis for the topology. For each point x choose a
member Bx of B such that x Bx [x, x + 1). The map R B : x 7 Bx is injective for
x = inf Bx .
R` has a countable dense subset: Q is dense since any (basis) open set in R` contains
rational points.
R` is Lindel
of: It suffices (!) to show that Sany open covering by basis open sets contains
a countable subcovering, ie that if R = jJ [aj , bj ) is covered by a collection of right
half-open intervals [aj , bj ) then R is actually already covered by countably many of these
intervals. Write


[
[
R=
(aj , bj ) R
(aj , bj )
jJ

jJ

as the (disjoint) union of the corresponding open intervals and the complement of this
union. The first set can be covered by countably many of the intervals (aj , bj ) for any

54

GENERAL TOPOLOGY

subset of R (with the standard topology) is Lindelof (24.2). Also the second set can be
covered by countably many of the intervals (aj , bj ) simply because it is countable: The
second set
[
[
[
[
R
(aj , bj ) =
[aj , bj )
(aj , bj ) = {x
{ak } | j J : ak 6 (aj , bj )}
jJ

jJ

jJ

kJ

consists of some of the left end-points of the intervals. For any element x of this set, choose
an interval I(x) = [ak , bk ) from the collection such that x = ak is the left end-point of
I(x). The corresponding open intervals Int I(x) = (ak , bk ) are disjoint for the left endpoint of Int I(x) is not in Int I(y) when x 6= y so there can be no overlap. Therefore the
map I is injective. But there is only room for countably many open disjoint intervals in
R (choose a rational point in each of them) so the second set contains at most countably
many elements.
R` is normal: Let A, B R` be disjoint closed sets.SFor each point a A RS B there
is an xa R such that [a, xa ) R B. Let UA = aA [a, xa ). Define UB = bB [b, xb )
similarly. Then UA and UB are open sets containing A and B, respectively. If UA UB 6=
then [a, xa ) [b, xb ) 6= for some a A, b B. Say a < b; then b [a, xa ) R B which
is a contradiction So UA and UB are disjoint. (R` is even completely normal [6, Ex 32.6]
by this argument.)
Example 25.5 (Sorgenfreys half-open square topology). The half-open rectangles [a, b) [c, d)
form a basis (12.4) for the product topology R` R` . The anti-diagonal L = {(x, x) | x R} is
a closed discrete (L [x, ) [x, ) = {(x, x)}) subspace of the same cardinality as R. Q Q
is a countable dense subspace.
R` R` is not Lindel
of: A Lindelof space can not contain an uncountable closed discrete
subspace (cf 20.14, [6, Ex 30.9]).
R` R` is not normal: A normal space with a countable dense subset can not contain
a closed discrete subspace of the same cardinality as R [4, 2.1.10]. (Let X be a space
with a countable dense subset. Since any continuous map of X into a Hausdorff space is
determined by its values on a dense subspace [6, Ex 18.13], the set of continuous maps
X R has at most the cardinality of RZ . Let X be any normal space and L a closed
discrete subspace. The Tietze extension theorem (26.4) says that any map L R extends
to continuous map X R. Thus the set of continuous maps X R has at least the
cardinality of RR which is greater (5.8) than the cardinality of RZ .) See [6, Ex 31.9] for a
concrete example of two disjoint closed subspaces that can not be separated by open sets.
R` R` is completely regular: since it is the product (28.3) of two compeletely regular
(even normal (25.4)) spaces.
Example 25.4 shows that the arrows of Theorem 24.2 do not reverse. Example 25.5 shows that
the product of two normal spaces need not be normal, that the product of two Lindelof spaces
need not be Lindel
of, and provides an example of a (completely) regular space that is not normal.
Example 25.6 (A Hausdorff space that is not regular). The open intervals plus the sets (a, b) K
where K = { n1 | n Z+ } form a basis for the topology RK (10.6, 19.8.(3)). This is a Hausdorff
topology on R since it is finer than the standard topology. But RK is not regular. The set K
is closed and 0 6 K. Suppose that U 3 0 and V K are disjoint open sets. We may choose
U to be a basis open set. Then U must be of the form U = (a, b) K for the other basis sets
containing 0 intersect K. Let k be a point in (a, b) K (which is nonempty). Since k is in the
open set V , there is a basis open set containing k and contained in V ; it must be of the form (c, d).
But U V ((a, b) K) (c, d) and ((a, b) K) (c, d) 6= for cardinality reasons. This is a
contradiction

GENERAL TOPOLOGY

55

26. Normal spaces


The two most important theorems about normal spaces are the Urysohn lemma and the Tietze
extension theorem.
Theorem 26.1 (Urysohn lemma). Let X be a normal space and let A and B be disjoint closed
subsets of X. Then there exists a continuous function (a Urysohn function) f : X [0, 1] such
that f (a) = 0 for a A and f (b) = 1 for b B.
Proof. We shall recursively define open sets Ur X for all r Q [0, 1] such that (make a
drawing!)
(26.2)

r < s = A U0 Ur U r Us U s U1 X B

To begin, let U1 be any open set such that A U1 X B, for instance U1 = X B. By


normality (25.2.(2)) there is an open set U0 such that A U0 U 0 U1 . To proceed, arrange
the elements of Q [0, 1] into a sequence
Q [0, 1] = {r0 = 0, r1 = 1, r2 , r3 , r4 , . . .}
such that the first two elements are r0 = 0 and r1 = 1. Assume that the open sets Uri satisfying
condition (26.2) have been defined for i n, where n 1. We shall now define Urn+1 . Suppose
that if we put the numbers r0 , r1 , . . . , rn , rn+1 in order
0 = r0 < < r` < rn+1 < rm < r1 = 1
the immediate predecessor of rn+1 is r` and rm is the immediate successor. Then U r` Urm by
(26.2). By normality (25.2.(2)) there is an open set Urn+1 such that U r` Urn+1 U rn+1 Urm .
The sets Uri , i n + 1, still satisfy (26.2).
Consider the function f : X [0, 1] given by
(
inf{r Q [0, 1] | x Ur } x U1
f (x) =
1
x X U1
Then f (B) = 1 by definition and f (A) = 0 since A U0 . But why is f continuous? It suffices
(15.2) to show that the subbasis intervals (11.1) [0, a), a > 0, and (b, 1], b < 1, have open preimages.
Since
[
f (x) < a r < a : x Ur x
Ur
r<a
0

f (x) > b r > b : x 6 U

r0

r > b : x 6 U r x

(X U r )

r>b

the sets
f 1 ([0, a)) =

Ur

and f 1 ((b, 1]) =

r<a

(X U r )

r>b

are open.

Example 26.3. Let X = R and let A = (, 1] and B = [2, ). If we let Ur = (, r) for


r Q [0, 1] then

0 x 0
f (x) = x 0 < x < 1

1 1x
is the Urysohn function with f (A) = 0 and f (B) = 1.
Theorem 26.4 (Tietze extension theorem). Every continuous map from a closed subspace A of a
normal space X into (0, 1), [0, 1) or [0, 1] can be extended to X.
Lemma 26.5. Let X be a normal space and A X a closed subspace. For r > 0 and any
continuous map f0 : A [r, r] there exists a continuous map g : X R such that
(26.6)

x X : |g(x)|

1
r
3

and

a A : |f0 (a) g(a)|

2
r
3

56

GENERAL TOPOLOGY

Proof. Since the sets f01 ([r, 13 r]) A and f01 ([ 13 r, r]) A are closed and disjoint in A they
are also closed and disjoint in X. Choose (26.1) a Urysohn function g : X [ 13 r, 13 r] such that
g(x) = 13 r on f01 ([r, 13 r]) and g(x) = 13 r on f01 ([ 13 r, r]). From
1
1 1
1
f01 ([r, r]) = f01 ([r, r]) f01 ([ r, r]) f01 ([ r, r])
3
3
3
3
|
{z
} |
{z
} |
{z
}
g= 31 r

|g| 31 r

g= 31 r

we see that the second inequality of (26.6) is satisfied.

Proof of 26.4. We shall first prove the theorem for functions from A to [0, 1] or rather to the
homeomorphic interval [1, 1] which is more convenient for reasons of notation.
Given the continuous function f : A [1, 1] we shall now recursively define a sequence g1 , g2 ,
of continuous functions X R such that
 n1
 n
n
X
1 2
2
(26.7)
x X : |gn (x)|
and a A : |f (a)
gi (a)|
3 3
3
i=1
To obtain g1 we apply 26.5 with f0 = f and r = 1. Suppose that g1 , . . . , gn1 have been defined. If

Pn1 
2 n1
then we obtain a function gn satisfying
we apply 26.5 with f0 = f
i=1 gi |A and r = 3
(26.7).
P
ByP
the first inequality in (26.7), the series n=1 gn (x) converges uniformly. The sum function

F = n=1 gn is continuous by the uniform limit theorem (17.15). By the second inequality in
(26.7), F (a) = f (a) for all a A.
Assume now that f : A (1, 1) maps A into the open unit interval. We know that we can
extend f to a continuous function F1 : X [1, 1] into the closed interval. We want to modify F1
so that it does not take the values 1 and 1 and stays the same on A. The closed sets F11 ({1})
and A are disjoint. There exists (26.1) a Urysohn function U : X [0, 1] such that U = 0 on
F11 ({1}) and U = 1 on A. Then F = U F1 is an extension of f that maps X into (1, 1).
A similar procedure applies to functions f : A [1, 1) into the half-open interval.

The properties mentioned in 26.1 and 26.4 characterize normal spaces among T1 -spaces.
Corollary 26.8. Let X be a T1 -space. Then the following conditions are equivalent:
(1) X is normal
(2) For any pair A, B of disjoint closed subspaces of X there exists a continuous function
f : X [0, 1] such that f (A) = 0 and f (B) = 1
(3) For any closed subset A of X and any continuous function f : A [0, 1] there exists a
continuous function F : X [0, 1] such that F |A = f .
Proof. The Urysohn lemma (26.1) says that (1) = (2) and the converse is clear. The Tietze
extension theorem (26.4) says that (1) = (3). Only (3) = (1) remains. Assume that X is a T1 space with property (3). Let A, B be two disjoint closed subsets. The function f : A B [0, 1]
given by f (A) = 0 and f (B) = 1 is continuous (15.5.(6)). Let F : X [0, 1] be a continuous
extension of f . Then A f 1 [0, 41 ) and B f 1 ( 34 , 1] where f 1 [0, 14 ) and f 1 ( 34 , 1] are open and
disjoint. Hence X is normal.

Many familiar classes of topological spaces are normal.
Theorem 26.9. Compact Hausdorff spaces are normal.
Proof. See 20.8.(2).

In fact, every regular Lindel


of space is normal [6, Ex 32.4] [4, 3.8.2].
Theorem 26.10. Metrizable spaces are normal.
Proof. Let X be a metric space with metric d and let A, B be disjoint closed sets. The continuous
function f : X [0, 1] given by
d(x, A)
f (x) =
d(x, A) + d(x, B)
is a Urysohn function with f (A) = {0} and f (B) = {1}. This shows that X is normal (26.8). 

GENERAL TOPOLOGY

57

Theorem 26.11. [4, Problem 1.7.4]. Linearly ordered spaces are normal.
Proof. We shall only prove the special case that every well-ordered space is normal. The half-open
intervals (a, b], a < b, are (closed and) open (14.2.(5)). Let A and B be two disjoint closed subsets
and let a0 denote the smallest element of X. Suppose that neither A nor B contain a0 . For any
point a A there exists a point xa < a such that (xa , a] is disjoint from B. Similarly, for any point
b B there exists a point xb < b such that (xb , b] is disjoint from A. The proof now proceeds as
the proof (25.4) for normality of R` . Suppose next that a0 A B, say a0 A. The one-point set
{a0 } = [a0 , a+
0 ) is open and closed (as X is Hausdorff). By the above, we can find disjoint open
sets U , V such that A {a0 } U and B V . Then A U {a0 } and B V {a0 } where the
open sets U {a0 } and V {a0 } are disjoint.

In particular, the well-ordered spaces S and S are normal (the latter space is even compact
Hausdorff).
27. Second countable regular spaces and the Urysohn metrization theorem
We investigate closer the class of regular spaces. We start with an embedding theorem that is
used in other contexts as well.
27.1. An embedding theorem. We discuss embeddings into product spaces.
Definition 27.2. A set {fj : X Yj | J} of continuous functions is said to separate points
and closed sets if for any point x X and any closed subset C X we have
x 6 C = j J : fj (x) 6 f (C)
Lemma 27.3. Let f : X Y be a map that separates points and closed sets. Assume that X is
T1 or that f is injective. Then f is an embedding.
Proof. The map separates points since points are closed in the T1 -space X. In particular, f is
injective. The assumption is that
f (x) f (C) = x C
when x X and C X is closed. This implies that f (X) f (C) f (C). Then actually
f (X) f (C) = f (C) for the other inclusion is obvious. But this equality says that f (C) is closed
in f (X). Hence the continuous map f : X f (X) is a homeomorphism.

Theorem 27.4 (Evaluation embedding theorem). Let {fj : X Yj | J} be a family of continuous functions that separates points and closed sets. Assume that X Q
is T1 or that at least one
of the functions fj is injective. Then the evaluation map f = (fj ) : X jJ Yj is an embedding.
Proof. Also the map f separates points and closed sets because
15.3.(6)

f (x) f (C) = j J : j f (x) j (f (C)) = j J : j f (x) j (f (C))


j J : fj (x) fj (C)

(fj ) separates

for all points x X and all closed subsets C X. The theorem now follows from 27.3.

xC


In particular, if one of the functions fj is injective and separates points and closed sets then
f = (fj ) is an embedding. For instance, the graph X X Y : x (x, g(x)) is an embedding
for any continuous map g : X Y .
27.5. A universal second countable regular space. We say that a space X is universal for
some property if X has this property and any space that has this property embeds into X.
Theorem 27.6 (Urysohn metrization theorem). The following conditions are equivalent for a
second countable space X:
(1) X is regular
(2) X is normal
(3) X is metrizable
The Hilbert cube [0, 1] is a universal second countable metrizable (or normal or regular) space.

58

GENERAL TOPOLOGY

Proof. (1) = (2): Let X be a regular space with a countable basis B. We claim that X is normal.
Let A and B be disjoint closed sets in X. By regularity (25.2.(1)), each point a A X B has
a basis neighborhood Ua B such that U a is disjoint from B. Let U1 , U2 , . . . be the elements in
the image of the map A B : a 7 Ua . Then

[
A
Un and U n B = for n = 1, 2 . . .
n=1

Similarly, there is a sequence V1 , V2 , . . . of basis open sets such that

[
B
Vn and V n A = for n = 1, 2 . . .
n=1

The open sets

n=1 Un and

U10
U20

n=1

Vn may not be disjoint. Consider instead the open sets

= U1 V 1

V10 = V1 U 1

= (V 1 V 2 )

V20 = V2 (U 1 U 2 )

..
.

..
.

Un0 = Un (V1 V n )

Vn0 = Vn (U1 U n )

Since Un0 covers as much of A as does Un and Vn0 covers as much of B as does Vn we have

[
[
A
Un0 and B
Vn0
n=1

where

[
n=1

Un0

Vn0 =

 [

n=1

mn

n=1

  [

0
0
Um
Vn0
Um
Vn0 =
m>n

0
X Vn does not intersect
X Um if m n and Um
Um does not intersect
because
0
Vn Vn if m > n.
(2) = (3): Let X be a normal space with a countable basis B. We show that there is a countable
set {fU V } of continuous functions X [0, 1] that separate points and closed sets.
For any closed subset C and any point x 6 C, or x X C, there are (25.2) basis open sets
U , V such that x U U V X C (choose V first). Let fU V : X [0, 1] be a Urysohn
function (26.1) such that fU V (U ) = 0 and fU V (X V ) = 1. Then fU V (x) = 0 and fU V (C) = 1 so
that fU V (x) 6 fU V (C). Since there are only countably many basis sets there are only countable
many of these Urysohn functions. According to the Diagonal embedding theorem (27.4) there is
an embedding X [0, 1] with the fU V as coordinate functions. Since [0, 1] is metrizable, the
subspace X is metrizable (17.5).
(3) = (1): Any metrizable space is regular, even normal (26.10).

0
Um

Vn0

The point is that in a second countable regular hence normal space there is a countable set of
(Urysohn) functions that separate points and closed sets. Therefore such a space embeds in the
Hilbert cube.
We have (27.6, 24.2) the following identities

 
 

2nd countable
2nd countable
2nd countable
=
=
regular
normal
metrizable

 
 

Countable dense subset
Lindelof
Subspace of
=
=
=
metrizable
metrizable
Hilbert cube
between classes of topological spaces.

GENERAL TOPOLOGY

59

28. Completely regular spaces and the StoneCech


compactification
There is no version of the Tietze extension theorem (26.4) for regular spaces. Instead we have
a new class of spaces.
Definition 28.1. A space X is completely regular if points are closed in X and for any closed
subset C of X and any point x
6 C there exists a continuous function f : X [0, 1] such that
f (x) = 0 and f (C) = 1.
Clearly
normal = completely regular = regular = Hausdorff = T1
and none of these arrows reverse (25.5, [6, Ex 33.11]).
In a completely regular space there are enough continuous functions X [0, 1] to separate
points and closed sets.
Theorem 28.2. The following conditions are equivalent for a topological space X:
(1) X is completely regular
(2) X is homeomorphic to a subspace of [0, 1]J for some set J
(3) X is homeomorphic to a subspace of a compact Hausdorff space
Proof. (1) = (2): If X is completely regular then the set C(X) of continuous maps X [0, 1]
separates points and closed sets. The evaluation map
Q
: X jC(X) [0, 1], j ((x)) = j(x), j C(X), x X,
is therefore an embedding (27.4).
(2) = (3): [0, 1]J is compact Hausdorff by the Tychonoff theorem (20.17).
(3) = (1): A compact Hausdorff space is normal (26.9) and hence completely regular. It is easy
to see that any subspace of a completely regular space is completely regular.

Corollary 28.3. (Cf 14.19, 25.3) Any subspace of a completely regular space is completely regular.
Any product of completely regular spaces is completely regular.
Proof. The first part is easily proved (and we used it above). The second part follows from (28.2)
because (15.9) the product of embeddings is an embedding.

Corollary 28.4. Any locally compact Hausdorff space is completely regular.
Proof. Locally compact Hausdorff spaces are open subspaces of compact spaces (23.4). Completely
regular spaces are subspaces of compact spaces (28.2).

Example 28.5 (A normal space with a non-normal subspace). Take any completely regular but
not normal space (for instance (25.5) R` R` ) and embed it into [0, 1]J for some set J (28.2). Then
you have an example of a normal (even compact Hausdorff) space with a non-normal subspace.

28.6. The StoneCech


construction. For any topological space X let C(X) denote the set of
continuous maps j : X I of X to the unit interval I = [0, 1] and let
Q
: X jC(X) I = map(C(X), I)
be the continuous evaluation map given by (x)(j) = j(x) or j = j for all j C(X). (The space
map(C(X), I) is a kind of double-dual of X). This construction is natural: For any continuous
map f : X Y of X into a space Y , there is an induced map C(X) C(Y ) : f of sets and yet
an induced continuous map
Y
Y
f
(28.7)
map(C(X), I) =
I
I = map(C(Y ), I)
jC(X)

kC(Y )

given by k f = f (k) = kf for all k C(Y ) (a kind of projection map). The map f is an
extension of f in the sense that the diagram
(28.8)

jC(X)

/Y

kC(Y )

60

GENERAL TOPOLOGY

is commutative: k f X = kf X = kf = k Y f .
Q
Put X = (X) and define : X X to be the corestriction of : X jC(X) I to X.
Then Q
X is a compact Hausdorff space (it is a closed subspace of the compact (20.17) Hausdorff
space I) and is, by design, a continuous map with a dense image X in X. This construction
is natural: For any continuous map f : X Y , the induced map f (28.7) takes X into Y for
(28.8)

f (X) = f (X X) f X X = Y f (X) Y Y = Y
and thus we obtain from (28.8) a new commutative diagram of continuous maps
(28.9)


X

/Y
Y


/ Y

where f : X Y is the corestriction to Y of the restriction of f to X. (This makes into


a functor with a natural transformation from the identity functor from the category of topological
spaces to the category of compact Hausdorff spaces.)
The next result says that X X is the universal map of X into a compact Hausdorff space.
Theorem 28.10. Let X be a topological space.
(1) The map X X is a continuous map of X into a compact Hausdorff space.
(2) For any continuous map f : X Y of X into a compact Hausdorff space Y there exists a
unique continuous map f : X X such that the diagram
(28.11)

/Y
XB
BB
|>
|
BB
||
BB
||
B!
|| f
X
f

commutes. (We say that f : X Y factors uniquely through X.)


(3) Suppose X X is a map of X to a compact Hausdorff space X such that any map
of X to a compact Hausdorff space factors uniquely through X. Then there exists a
homeomorphism X X such that
XB
BB
||
BB
|
|
BB
|
|
B!
|
|}
'
/ X
X
commutes.
(4) If X is completely regular then X X is an embedding. If X is compact Hausdorff then
X X is a homeomorphism.
Q
Proof. (4): Since the corestriction of an embedding is an embedding (15.9) and X I is an
embedding when X is completely regular (28.2), also X X is an embedding. If X is compact
Hausdorff, X is normal (26.9), hence completely regular, so we have just seen that : X X is
an embedding. The image of this embedding is closed (20.9.(1)) and dense. Thus the embedding
is bijective so it is a homeomorphism.
(1) and (2): If Y is compact Hausdorff then by item (4) Y is a homeomorphism in diagram (28.9)
and hence f = 1
Y f is a possibility in (28.11). It is the only possibility [6, Ex 18.13], for the
image of X is dense in X.
(3) Let X X be a map to a compact Hausdorff space satisfying the above universal property.
/ X that, by uniqueness, are inverse to each other. 
Then there exist continuous maps X o
Definition 28.12. A compactification of the space X is an embedding c : X cX of X into a
compact Hausdorff space cX such that c(X) is dense in cX.

GENERAL TOPOLOGY

61

If X is completely regular then X X is a compactification (28.10.(4)) and it is called the

StoneCech
compactification of X. Conversely, if X has a compactification then X is homeomorphic to a subspace of a compact space and therefore (28.2) completely regular. We state these
observations in
Lemma 28.13. X has a compactification if and only if X is completely regular.

The StoneCech
compactification X of a completely regular space X is the maximal compactification in the sense that for any other compactification X cX there is a closed quotient map
X cX map such that
XB
BB
||
|
BB
|
|
BB
|
|
B
}|
/ cX
X
commutes. The map X cX is closed because it is continuous map of a compact space into a
Hausdorff space (20.9.(1)) and surjective because its image is closed and dense.
The Alexandroff compactification X (23.2) of a noncompact (!) locally compact Hausdorff
space X is the minimal compactification in the sense that X = cX/(cX X) for any other
compactification X cX. Indeed, there is a closed quotient map cX X, taking cX X to
, such that
XC
CC
||
CC
|
|
CC
|
|
C!
|
}|
/ X
cX
commutes. The fact that X is open (23.9) in cX implies (as in the proof of 23.2) that this map is
continuous. (What is the minimal compactification of a compact Hausdorff space?)
For instance, there are quotient maps
[0, 1] QQ
QQQ
pp8
p
QQ(
p
pp
R OO
R = S 1
OOO
ll5
l
l
l
OO'
lll
C
of compactifications of R. Here
1
C = {0} [1, 1] {(x, sin ) | 0 < x 1 } L
x
is the Warsaw circle, a compactification of R with remainder C R = [1, 1], which is obtained by
closing up the closed topologists sine curve S by (a piece-wise linear) path L from (0, 0) to ( 1 , 0).
In particular, C/[1, 1] = S 1 . More generally, there are quotient maps Rn [0, 1]n Rn = S n
of compactifications of euclidian n-space Rn .

Investigations of the StoneCech


compactification of the integers Z raise several issues of a
very fundamental nature [9, 10, 8] whose answers depend on your chosen model of set theory.
29. Manifolds
Definition 29.1. A manifold is a locally euclidian second countable Hausdorff space.
Manifolds are locally compact and locally path connected since they are locally euclidian. They
are also metrizable, hence they satisfy all countability axioms (24.1) and all separation axioms
(25.1):




locally compact Hausdorff
completely regular
28.4
manifold =
=
2nd countable
2nd countable

metrizable

27.6,24.2
Lindelof
26.10
=
= metrizable = normal
2nd
countable

countable dense subset

62

GENERAL TOPOLOGY

The line with two zeroes is an example of a locally euclidian space that is not Hausdorff. The
long line [6, Ex 24.12] is a locally euclidian space that is not 2nd countable.
The nsphere S n and the real projective nspace RP n (16.13) are examples of manifolds. S n
lies embedded in Rn+1 from birth but what about RP n ? Does RP n embed in some euclidian
space?
Theorem 29.2 (Embeddings of compact manifolds). Any compact manifold embeds in RN for
some N .
The support of a continuous function : X R is the closed set
supp = {x X | f (x) > 0}
The function is constant with value 0 outside its support.
Theorem 29.3 (Partition of unity). Let X = U1 Uk be a finite open covering of the normal
space X. The there exist continuous functions i : X [0, 1], 1 i k, such that
(1) supp i Ui
Pk
(2)
i=1 i (x) = 1 for all x X.
Proof. We show first that there is an open covering {Vi } of X such that V i Ui . Since the closed
set X (U2 Uk ) is contained in the open set U1 there is an open set V1 such that
X (U2 Uk ) V1 V 1 U1
by normality (25.2). Now V1 U2 Uk is an open covering of X. Apply this procedure once
again to find an open set V2 such that V 2 U2 and V1 V2 U3 Uk is still an open covering
of X. After finitely many steps we have an open covering {Vi } such that V i Ui for all i.
Do this one more time to obtain an open covering {Wi } such that Wi W i Vi V i Ui for
all i. Now choose a Urysohn function (26.1) i : X [0, 1] such that i (W i ) = 1 and i (X Vi ) =
P
0. Then supp i = {x | i (x) > 0} V i Ui and (x) = i (x) > 1 for any x X since x Wi
for some i. Hence i = i is a well-defined continuous function on M taking values in the unit
Pk
Pk
Pk
(x)
1
i (x)
interval such that i=1 i (x) = i=1 (x)
= (x)

i=1 i (x) = (x) = 1.
Proof of 29.2. Let M be a compact manifold. Cover M by finitely many open sets Ui , 1 i k,
homeomorphic to open sets in Rn by embeddings gi : Ui Rn . Let {i } be a partition of unity
for the open covering {Ui }. Extend gi to a function g i : M Rn given by
(
i (x)gi (x) x Ui
g i (x) =
0
x 6 Ui
The extension g i is continuous because the restrictions to the open sets Ui and X supp i are
continuous (15.5.(5)). We claim that the continuous map
k

}|
{
z
}|
{ z
F : M R R Rn Rn = Rk(n+1)
x 7 (1 (x), . . . , k (x), g 1 (x), . . . , g k (x))
is an embedding. Since M is compact and Rk(n+1) is Hausdorff it is enough (20.9.(2)) to show
that F is injective: Suppose that F (x) = F (y). Choose i such that i (x) > 0. Then i (y) = i (x)
is also positive so x, y Ui . But g i |Ui = i gi |Ui is injective so x = y.

In particular RP 2 (16.13) embeds into some euclidian space. Which one?
There is up to homeomorphism just one compact manifold of dimension 1, the circle. Compact
manifolds of dimension 2 are described by a number, the genus, and orientability. In particular,
any simply connected 2-dimensional manifold is homeomorphic to S 2 . We do not know (December
2003) if there are any simply connected 3-dimensional manifolds besides S 3 (Poincare conjecture).
Classification of 4-dimensional compact manifolds is logically impossible.

GENERAL TOPOLOGY

63

References
[1] Colin C. Adams, The knot book, W. H. Freeman and Company, New York, 1994, An elementary introduction
to the mathematical theory of knots. MR 94m:57007
[2] Nicolas Bourbaki, General topology. Chapters 14, Elements of Mathematics, Springer-Verlag, Berlin, 1998,
Translated from the French, Reprint of the 1989 English translation. MR 2000h:54001a
[3] Glen E. Bredon, Topology and geometry, Graduate Texts in Mathematics, vol. 139, Springer-Verlag, New York,
1993. MR 94d:55001
[4] Ryszard Engelking, General topology, second ed., Sigma Series in Pure Mathematics, vol. 6, Heldermann Verlag,
Berlin, 1989, Translated from the Polish by the author. MR 91c:54001
[5] Ryszard Engelking and Karol Sieklucki, Topology: a geometric approach, Sigma Series in Pure Mathematics,
vol. 4, Heldermann Verlag, Berlin, 1992, Translated from the Polish original by Adam Ostaszewski. MR
94d:54001
[6] James R. Munkres, Topology. Second edition, Prentice-Hall Inc., Englewood Cliffs, N.J., 2000. MR 57 #4063
[7] Edwin H. Spanier, Algebraic topology, Springer-Verlag, New York, 1981, Corrected reprint. MR 83i:55001

[8] Juris Stepr


ans, The autohomeomorphism group of the Cech-Stone
compactification of the integers, Trans. Amer.
Math. Soc. 355 (2003), no. 10, 42234240 (electronic). MR 1 990 584
[9] Jan van Mill, An introduction to , Handbook of set-theoretic topology, North-Holland, Amsterdam, 1984,
pp. 503567. MR 86f:54027

[10] Russell C. Walker, The Stone-Cech


compactification, Springer-Verlag, New York, 1974, Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 83. MR 52 #1595
[11] W. Hugh Woodin, The continuum hypothesis, Part I, Notices of the American Math. Soc. 48 (2001), no. 6,
567576.
[12]
, The continuum hypothesis, Part II, Notices of the American Math. Soc. 48 (2001), no. 7, 681690.

Вам также может понравиться