Sei sulla pagina 1di 55

PART A TOPOLOGY COURSE: HT 2013

Contents
1. Metric spaces 2
1.1. What is Topology about? 2
1.2. Denitions and examples 3
1.3. Balls, open, closed and bounded sets 4
1.4. Maps between metric spaces 7
1.5. Convergent sequences 10
1.6. Complete metric spaces 12
1.7. The space of bounded continuous functions 14
1.8. Various strong versions of continuity 15
2. Topological spaces 17
2.1. Denitions and examples 17
2.2. Closure of a set 20
2.3. Interior of a set 23
2.4. Boundary of a set 24
2.5. Separation axioms 24
2.6. Subspace of a topological space 25
2.7. Basis for a topology 27
2.8. Product of topological spaces 29
3. Connected spaces 32
3.1. Denition and properties of connected spaces 32
3.2. Path-connected spaces 34
4. Compact spaces 37
4.1. Denition and properties of compact sets 38
4.2. Compact spaces and continuous maps 41
4.3. Sequential compactness 43
5. Quotient spaces, quotient topology 46
5.1. Denition and examples 46
5.2. Separation axioms 48
5.3. Quotient spaces and continuous maps 50
Recommended books in Topology:
(1) J. Dugundji, Topology;
(2) J.R. Munkres, Topology, A First Course;
(3) W. Sutherland, Introduction to Metric and Topological Spaces (particularly recommended);
(4) O. Viro, O. Ivanov, V. Kharlamov, N. Netsvetaev, Elementary Topology,
http://www.math.uu.se/oleg/educ-texts.html
For further reading see the list at the end of the synopsis on the Maths Institute web pages.
These notes are adapted from a set of lecture notes written by Cornelia Drutu and Wilson Suther-
land. I wish to thank them both for having allowed me to use them.
2 PART A TOPOLOGY COURSE: HT 2013
1. Metric spaces
1.1. What is Topology about? You have already seen and used some topological notions:
(1) In Analysis, several concepts like
convergent sequences;
bounded sequences;
continuous functions;
need a way of dening nearby points as well as a way of dening points not too far apart.
This can be done via a distance (or metric).
In order to dierentiate, one needs to consider functions dened on open sets.
(2) In Linear Algebra, you have seen the notion of an inner product space. In such a space one can
dene
the length of a vector (hence of distance between two points);
the angle between vectors: in particular one may speak of orthogonality, and use trigonometry.
In what follows we weaken that structure, and we only require the notion of the distance between
two points. In particular we give up the notion of angle.
This notion of distance is encapuslated in the following denition.
PART A TOPOLOGY COURSE: HT 2013 3
1.2. Denitions and examples.
Denition 1.1. A metric space is a non-empty set X endowed with a function d : X X R with
the following properties:
(M1) d(x, y) 0 for all x, y X; d(x, y) = 0 if and only if x = y;
(M2) (Symmetry) for all x, y X, d(y, x) = d(x, y);
(M3) (Triangle inequality) for all x, y, z X, d(x, z) d(x, y) + d(y, z).
The function d is called a metric or distance.
e
e
e
e
e
e
e
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
rj
y z
x
d(x, z)
A rst example of a metric space is provided by the following familiar structure.
Denition 1.2. A normed vector space (NVS) is a vector space V over R or C together with a
function V R called a norm and written v [[v[[ such that
(N1) for all v V, [[v[[ 0; [[v[[ = 0 if and only if v = 0;
(N2) for all v V and any scalar , [[v[[ = [[ [[v[[;
(N3) (subadditivity) for all u, v V, [[u +v[[ [[u[[ +[[v[[.
Proposition 1.3. A norm [[ [[ denes a metric on V , d(x, y) = [[x y[[ , for every x, y V .
Proof. Property (M1) follows from (N1). Property (M2) follows from (N2) applied to = 1 and to
v = x y.
Property (M3) follows from (N3) applied to u = x y and v = y z .
Remarks 1.4. (1) A vector subspace U of a normed vector space (V, | |) is naturally endowed
with a norm: the restriction of | | to U.
(2) A Cartesian product V W of two normed vector spaces (V, | |
V
) and (W, | |
W
) can be
endowed with several norms: | |
V
+| |
W
, max (| |
V
, | |
W
) and
_
| |
2
V
+| |
2
W
.
Example 1.5. On R
n
several norms can be dened. For x = (x
1
, x
2
, . . . , x
n
), the l
2
, l
1
and l

norms are
[[x[[
2
=

_
n

i=1
x
2
i
, [[x[[
1
=
n

i=1
[x
i
[ , [[x[[

= max
1in
[x
i
[ .
See Sheet 1, Exercise 5.
4 PART A TOPOLOGY COURSE: HT 2013
Remark 1.6. The norm [[ [[
2
is dened by the standard inner product (or dot product) on R
n
, that
is [[x[[
2
=
_
x, x, where
x, y = x
1
y
1
+x
2
y
2
+ +x
n
y
n
.
The space R
n
endowed with the above inner product is sometimes called Euclidean space, therefore
[[ [[
2
is called the Euclidean norm and the metric it denes the Euclidean metric on R
n
.
The other two norms [[ [[
1
and [[ [[

are not dened by any inner product on R


n
.
A more complicated example is the following.
Example 1.7. (Function spaces) Let S be any non-empty set and let X be the set B(S, R) of all
bounded functions from S to R. Then X is a real vector space, where for f, g X and R we
dene f +g and f by:
(f +g)(s) = f(s) +g(s), (f)(s) = f(s) .
Moreover [[f[[

= sup
sS
[f(s)[ is a norm on B(S, R).
Proof. Property (N1) is immediate.
For any scalar and function f X, [[f[[

= sup
sS
[f(s)[ = [[ sup
sS
[f(s)[ = [[[[f[[

.
For any two functions f, g X and any s S, [f(s) + g(s)[ [f(s)[ + [g(s)[ [[f[[

+ [[g[[

.
Whence [[f +g[[

[[f[[

+[[g[[

.
There are many metric spaces with a metric that is not dened by a norm, as in the following
example. This is also a rather counter-intuitive example: a metric space in which all distinct points
are at the same distance.
Example 1.8 (Discrete metric spaces). The discrete metric on any non-empty set X is dened by
d(x, y) =
_
0 if x = y,
1 if x ,= y.
This space (X, d) is called a discrete metric space.
Proof. Properties (M1) and (M2) are clearly satised.
In (M3) it is enough to study the case when x ,= z . Then either x ,= y or z ,= y. Thus d(x, z) = 1
and d(x, y) + d(y, z) 1.
Examples 1.9. (1) A non-empty subset A of a metric space (X, d) is naturally endowed with a
metric, written d
A
: the restriction of d to AA.
(2) A Cartesian product X
1
X
2
of two metric spaces can be endowed with a metric (see Sheet
1, Exercise 4).
1.3. Balls, open, closed and bounded sets.
Denition 1.10. Let (X, d) be a metric space, x
0
a point in X and r > 0.
The open ball in X of radius r centred at x
0
is the set B(x
0
, r) = x X : d(x, x
0
) < r.
The closed ball in X of radius r centred at x
0
is the set B

(x
0
, r) = x X : d(x, x
0
) r.
The sphere in X of radius r centred at x
0
is the set S(x
0
, r) = x X : d(x, x
0
) = r.
If more than one metric is dened on X then we write B
d
(x
0
, r), B

d
(x
0
, r), S
d
(x
0
, r).
Examples 1.11. (a) X = R
2
, d = d
2
: B
d2
(x
0
, r) is the open disc of radius r centred at x
0
.
(b) X = R
2
, d = d
1
: B
d1
(x
0
, r) is the inside of the square centred on x
0
with diagonals of
length 2r parallel to the axes, as in the diagram below.
PART A TOPOLOGY COURSE: HT 2013 5
(c) X = R
2
, d the discrete metric: B
d
(x
0
, r) =
_
x
0
if r 1
R
2
if r > 1.
(d) X the set B([0, 1], R) of all bounded real-valued functions on [0, 1] , d = d

the sup metric:


for f
0
X and r > 0, B
d
(f
0
, r) is the set of all f X whose graphs lie inside a ribbon of
vertical width 2r centred on the graph of f
0
.
d
d
d
d

d
d
d
d
(r, 0)
Denition 1.12. A subset A of a metric space X is bounded if A B(x
0
, r) for some x
0
X and
r > 0.
Denition 1.13. Let (X, d) be a metric space and U X. We say U is open in X if for every
x U , there exists
x
> 0 such that B(x,
x
) U.
Examples 1.14. (a) X = R
2
, d = d
2
: open sets are as in the MT Analysis course.
(b) X = R
2
, d = d
1
or d

: open sets exactly as in (a) above (see Sheet 1, Exercise 6).


(c) X endowed with the discrete metric: any subset U X is open.
(d) Any open ball B(x
0
, r) in a metric space (X, d) is open in X.
Let us explain why (d) is true: take an arbitrary point x B(x
0
, r) and
x
= r d(x, x
0
). Then
for every y B(x,
x
), d(x
0
, y) d(x
0
, x) + d(x, y) < r. Hence, B(x,
x
) B(x
0
, r).
Proposition 1.15. A subset in a metric space is open if and only if it is a union of open balls.
Proof. Consider a union of open balls A =

iI
B(x
i
, r
i
). For every x A there exists j I such
that x B(x
j
, r
j
). According to the argument justifying Example (d) above, for
x
= r
j
d(x, x
j
)
the ball B(x,
x
) is contained in B(x
j
, r
j
), hence in A.
Conversely, let U be an open subset in a metric space (X, d). For every x U , there exists
x
> 0
such that B(x,
x
) U.
Then

xU
B(x,
x
) U . Since x B(x,
x
), it follows that U

xU
B(x,
x
). Whence the
equality.
Corollary 1.16. A subset in (R, [ [) is open if and only if it is a union of open intervals.
Proposition 1.17. Suppose that X is a metric space. Then
(a) X and are open in X;
(b) if U and V are open in X, then so is U V ;
(c) if U
i
is open in X for all i I , then so is
_
iI
U
i
.
Remark 1.18. Note that in Proposition 1.17, (c), I might be innite, even uncountable.
6 PART A TOPOLOGY COURSE: HT 2013
Proof. (b) Consider an arbitrary point x U V .
As U is open there exists
x
> 0 such that B(x,
x
) U . Likewise, as x V and V is open
there exists
x
such that B(x,
x
) V .
Then for
x
= min(
x
,
x
), B(x,
x
) B(x,
x
) U and B(x,
x
) B(x,
x
) V . Therefore
B(x,
x
) U V .
(c) For every x

iI
U
i
there exists k I such that x U
k
. Since U
k
is open there exists

x
> 0 such that B(x,
x
) U
k

iI
U
i
.
Remark 1.19. Property (b) in Proposition 1.17 is not true for innite intersections:

nN,n1
_

1
n
,
1
n
_
= 0 .
Denition 1.20. A subset V of a metric space X is closed in X if X V is open in X.
Examples 1.21. (a) In Euclidean spaces, the closed sets are exactly as you saw last term.
(b) In a discrete metric space X, any subset of X is closed in X (since any subset is open in X).
(c) Any closed ball B

d
(x
0
, r) in a metric space (X, d) is closed in X.
Remark 1.22. (1) In a metric space there may exist subsets that are neither open nor
closed. For instance intervals [a, b) in (R, [ [).
(2) On the other hand in a discrete metric space X all subsets are both open and closed.
Proposition 1.23. Suppose that X is a metric space. Then
(a) X and are closed in X;
(b) if A and B are closed in X, then so is A B;
(c) if A
i
is closed in X for all i I , then so is

iI
A
i
.
Proof. This follows from the fact that A is closed if and only if X A is open, from Proposition 1.17,
and from the De Morgan laws (see the Appendix).
Remark 1.24. Property (b) in Proposition 1.23 is not true for innite unions:
_
nN,n1
_
1
n
, 1
1
n
_
= (0, 1) .
Proposition 1.25. Let (X, d) be a metric space and let A X be a non-empty subset endowed with
the restricted metric d
A
.
Then a subset Y A is open (respectively, closed) in (A, d
A
) if and only if Y = A U , where U
is open (respectively, closed) in X.
PART A TOPOLOGY COURSE: HT 2013 7
Proof. We begin by noting that if a A and r > 0 then B
d
A
(a, r) = B(a, r) A.
Assume that Y A is open in (A, d
A
). Then for every y Y there exists
y
> 0 such that
B
d
A
(y,
y
) Y . As in the proof of Proposition 1.15 it follows that Y =

yY
B
d
A
(y,
y
).
Take
U =
_
yY
B(y,
y
) .
Then U A =

yY
[B(y,
y
) A] = Y .
Conversely, let U be an open set in X and Y = U A. For every y Y U there exists
y
> 0
so that B(y,
y
) U . Then B(y,
y
) A = B
d
A
(y,
y
) U A = Y .
The statement on closed sets follows immediately from the statement on open sets, and from the
fact that A (A U) = A (X U).
1.4. Maps between metric spaces.
Denition 1.26. Let (X, d
X
) and (Y, d
Y
) be metric spaces. A function f : X Y is an isometric
embedding if, for every x
1
, x
2
X,
d
Y
(f(x
1
), f(x
2
)) = d
X
(x
1
, x
2
).
If, in addition, f is surjective, f is said to be an isometry.
Example 1.27. The function i : (R, d) (R
2
, d
2
), x (x, 0) is an isometric embedding, but not an
isometry.
Remarks 1.28. (1) An isometric embedding is necessarily injective.
(2) If f is an isometry, then f
1
is dened and is also an isometry.
(3) The composition of two isometric embeddings is an isometric embedding. The composition of
two isometries is an isometry.
For most purposes, isometric embeddings and isometries are too rigid. So, we come to a key
denition.
Denition 1.29. Let (X, d
X
) and (Y, d
Y
) be metric spaces and let f : X Y be a function.
(a) We say that f is continuous at x
0
X if given any > 0, there exists > 0 such that
whenever d
X
(x, x
0
) <
d
Y
(f(x), f(x
0
)) < ;
equivalently, such that f(B
d
X
(x
0
, )) B
d
Y
(f(x
0
), ).
(b) We say that f is continuous if f is continuous at every x
0
X.
Examples 1.30. (1) The projection map p: (R
2
, d
2
) (R, d), (x
1
, x
2
) x
1
is continuous. For
in the above denition, we may set to be .
(2) If (Y, d
Y
) is any metric space and (X, d
disc
) is a discrete metric space, then any function
f : (X, d
disc
) (Y, d
Y
) is continuous.
To see this, x
0
be any point of X, and let > 0. Setting = 1/2 gives B(x
0
, ) = x
0
.
So, f(B
d
disc
(x
0
, )) B
d
Y
(f(x
0
), ), as required.
Now we come to another key denition.
Denition 1.31. Let (X, d
X
) and (Y, d
Y
) be metric spaces. A function f : X Y is a homeomor-
phism if it is a bijection, and both f and f
1
are continuous. We then say that (X, d
X
) and (Y, d
Y
)
are homeomorphic.
Examples 1.32. (1) (R, d) and ((0, 1), d) are homeomorphic. (See Sheet 1, Exercise 3.)
8 PART A TOPOLOGY COURSE: HT 2013
(2) The surface of a cylinder (without the ends)
X = (x, y, z) R
3
[ x
2
+y
2
= 1 , 1 < z < 2
and an annulus lying in the xy-plane, without the boundary,
Y = (x, y, 0) [ 1 < x
2
+y
2
< 4
are homeomorphic.
Indeed, the maps
f : X Y , f(x, y, z) = (zx, zy, 0)
and
f
1
: Y X , f
1
(x, y, 0) =
_
x
_
x
2
+y
2
,
y
_
x
2
+y
2
,
_
x
2
+y
2
_
are both continuous.
We leave it as an exercise to check that f(X) Y , f
1
(Y ) X and that f f
1
=
id
Y
, f
1
f = id
X
.
(3) Let S
1
be the unit circle in R
2
. Give it the metric that it inherits as a subset of R
2
. Let Q
be the hollow square with side length 2 centred at the origin. Also give it the metric that it
inherits as a subset of R
2
. Dene
r : Q S
1
(x, y)
_
x
_
x
2
+y
2
,
x
_
x
2
+y
2
_
which is radial projection onto S
1
. In Sheet 1, Exercise 4, you prove that r is a homeomor-
phism.
Proposition 1.33. If A is a non-empty subset of the metric space (X, d) then the inclusion map
i : A X is continuous, when A is endowed with the restricted metric d
A
.
Proof. Let x
0
A and let > 0. Then, set = and note that i(B
d
A
(x
0
, )) B
d
(x
0
, ).
Remarks 1.34. (1) Every isometric embedding is continuous.
(2) Hence, every isometry is a homeomorphism.
(3) Not every homeomorphism is an isometry. Example: id : (R
n
, | |
2
) (R
n
, | |
1
) (or id :
(R
n
, | |
2
) (R
n
, | |

)).
PART A TOPOLOGY COURSE: HT 2013 9
It is not true that if f : X Y is a continuous bijection between metric spaces, then f
1
is
necessarily also continuous.
Examples 1.35. (1) Let f be the identity map (R, d
disc
) (R, d) where d
disc
is the discrete
metric. By Example 1.30 (2), f is continuous. But f
1
is not continuous at any x R. To see
this, set = 1/2. Then, there is no > 0 such that f
1
(B
d
(x, )) B
d
disc
(f
1
(x), ) = x.
(2) Consider f : [0, 1) S
1
, t (cos(2t), sin(2t)). It is not hard to prove that f is a
continuous bijection. But f
1
is not continuous at (1, 0).
Continuity of maps can also be dened using either open sets or closed sets.
Recall that if f : X Y is a function, then for any subset A of Y , its inverse image (or pre-image)
is
f
1
(A) = x X : f(x) A.
So, f
1
(A) is dened even when f is not invertible.
Proposition 1.36. Let X, Y be metric spaces and let f : X Y be a map. Then f is continuous
if and only if the inverse image of every open set is open (i.e. whenever U is open in Y , the
inverse image f
1
(U) is open in X).
Proof. Assume that f is continuous in the sense of Denition 1.29. Let U be an open set in Y . We
prove that f
1
(U) is open in X. Let x
0
be an arbitrary point in f
1
(U). Then f(x
0
) U , and
since U is open there exists > 0 such that B(f(x
0
, )) U .
The map f is continuous at x
0
therefore for > 0 given above there exists > 0 such that
f(B(x
0
, )) B(f(x
0
, )) U . This implies that B(x
0
, ) f
1
(U).
Conversely, assume that f
1
(U) is open in X whenever U is open in Y . Let us prove that f is
continuous in the sense of Denition 1.29.
Let x
0
be an arbitrary point in X, and let > 0 be an arbitrary positive number. The set
U = B(f(x
0
), ) is open, therefore by hypothesis f
1
(U) is an open set, moreover containing x
0
. It
follows that there exists > 0 such that B(x
0
, ) f
1
(U). This implies that f (B(x
0
, )) U =
B(f(x
0
), ).
We have thus proved that f is continuous at x
0
, and x
0
was arbitrary.
Proposition 1.37. Let X, Y be metric spaces and let f : X Y be a map.
Then f is continuous if and only if the inverse image of every closed set is closed (i.e.
whenever V is closed in Y the inverse image f
1
(V ) is closed in X).
Proof. This follows from Proposition 1.36 and the formula for complements of inverse images:
f
1
(Y V ) = X f
1
(V )

Above, we saw that closed and open sets can used to verify continuity. Below, we show that some
sets can be shown to be open or closed by using continuous maps.
Proposition 1.38 (dening open/closed sets via continuous functions). Let X be a metric space and
let f : X R be a continuous function. Then for every R
(1) the sets x X [ f(x) > and x X [ f(x) < are open;
(2) the sets x X [ f(x) , x X [ f(x) and x X [ f(x) = are closed.
10 PART A TOPOLOGY COURSE: HT 2013
Proof. (1) follows from Proposition 1.36 applied to U = (, +), and to U = (, ), open sets in
R.
(2) follows from Proposition 1.37 applied to F = [, +), to F = (, ] , and to F = ,
closed sets in R.

Examples 1.39. (1) The set x R : e


x
> cos x is open in R with the Euclidean metric.
(2) The planar set (x, y) R
2
: x
3
y
3
+y is closed in R
2
with the Euclidean metric (or the
metric d
1
or the metric d

).
We investigate the behaviour of continuity with respect to composition of maps.
Proposition 1.40. If f : X Y and g : Y Z are continuous maps, where X, Y, Z are metric
spaces, then so is g f .
Proof. Let U be an open set in Z.
Since g is continuous, g
1
(U) is an open subset of Y .
Since f is continuous, f
1
_
g
1
(U)
_
is an open subset of X.
But the latter is the same as (g f)
1
(U).
1.5. Convergent sequences. In the setting of metric spaces, convergent sequences can be used to
dene most of the important topological notions (closed sets, continuous functions etc).
Denition 1.41. A sequence (x
n
) in a metric space (X, d) converges to a point a X if for every
> 0, there exists N N such that whenever n N
d(x
n
, a) < :
equivalently, whenever n N , x
n
B(a, ).
The sequence (x
n
) is called a convergent sequence; the element a is called the limit of the sequence
(x
n
).
Remark 1.42. In the above one can replace the inequality d(x
n
, a) < by the inequality d(x
n
, a) .
Indeed, consider the properties:
(strict) for every > 0, there exists N N such that whenever n N
d(x
n
, a) < .
(non-strict) for every > 0, there exists N N such that whenever n N
d(x
n
, a) .
We prove that these properties are equivalent.
Clearly (strict) implies (non-strict). Conversely, assume that (non-strict) is satised, and
let us prove (strict). Consider an arbitrary > 0. Property (non-strict) applied to

2
implies
that there exists N such that for every n N, d(x
n
, a)

2
< .
Proposition 1.43. The limit of a convergent sequence in a metric space is unique.
PART A TOPOLOGY COURSE: HT 2013 11
Proof. Indeed, assume that (x
n
) converges to a, and that (x
n
) converges to b. Take > 0 arbitrary.
There exists N
1
such that for every n N
1
, d(x
n
, a) < , and N
2
such that for every n N
2
,
d(x
n
, b) < . Then for any n max(N
1
, N
2
),
d(a, b) d(x
n
, a) + d(x
n
, b) < 2 .
We have thus found that for every > 0, d(a, b) < 2. It follows that d(a, b) = 0, whence
a = b.
Examples 1.44. (a) X = R with the Euclidean metric: convergence is the usual convergence of
(x
n
) to a.
(b) X with a discrete metric: a sequence (x
n
) is convergent if and only if there exists N such
that x
n
= x
N
for every n N.
(c) X = B(A, R), where A R, endowed with the metric
d

(f, g) = |f g|

= sup
aA
[f(a) g(a)[ .
The sequence (f
n
) of functions converges to f in (X, d

) if and only if (f
n
) converges to
f uniformly on A.
We can also rephrase convergence in terms of open sets.
Proposition 1.45. Let (x
n
) be a sequence of points in a metric space (X, d). Then the following
are equivalent
(1) (x
n
) converges to a point a X;
(2) for every open set U containing a, there exists a natural number N such that
n N x
n
U.
Proof. (1) (2): Suppose (x
n
) converges to a, and that U is an open set containing a. Then there
is an > 0 such that B(a, ) U . By denition of convergence, there is an N N such that
n N x
n
B(a, ) U,
as required.
(2) (1): For any > 0, B(a, ) is open. So, assuming (2), there is an N N such that
n N x
n
B(a, ),
as required.
One can use convergent sequences to check that a function between metric spaces is continuous.
Proposition 1.46. A map f : X Y between two metric spaces is continuous if and only if for
every x X and every sequence (x
n
) in X converging to x, the sequence (f(x
n
)) converges to f(x).
Proof. Assume that f is continuous. Let (x
n
) be a sequence that converges to x X. We want to
prove that (f(x
n
)) converges to f(x). More precisely, we want to show that for every > 0 there
exists N such that f(x
n
) B(f(x), ) whenever n N.
The map f is continuous at x, therefore there exists > 0 such that f(B(x, )) B(f(x), ).
As (x
n
) converges to x, there exists N such that for any n N, x
n
B(x, ), therefore f(x
n
)
f(B(x, )) B(f(x), ).
Conversely, assume that for every x X and every sequence (x
n
) in X converging to x, the
sequence (f(x
n
)) converges to f(x). We want to prove that f is continuous.
Assume that f is not continuous at some point x. Then there exists > 0 such that for every > 0
the set f(B(x, )) has some point outside B(f(x), ). In particular for =
1
n
there exists x
n
B(x,
1
n
)
such that d(f(x
n
), f(x)) . The sequence (x
n
) then converges to x while d(f(x
n
), f(x)) .
12 PART A TOPOLOGY COURSE: HT 2013
On the other hand, since (x
n
) converges to x by hypothesis f(x
n
) converges to f(x). This
contradicts the inequality d(f(x
n
), f(x)) .
We may then conclude that f is continuous.
The notion of convergent sequence allows one to dene closed subsets in metric spaces without
having to refer to their complementary subsets.
Proposition 1.47. A subset F in a metric space (X, d) is closed if and only if for every convergent
sequence contained in F its limit is also in F .
Proof. Assume that F is closed and let (x
n
) be a convergent sequence in F with limit a.
If a X F , as X F is open, there exists > 0 such that B(a, ) X F .
Since (x
n
) converges to a it follows that there exists N such that for every n N, x
n
B(a, )
X F . This contradicts the fact that the entire sequence (x
n
) is contained in F .
Therefore a cannot be in X F , it must be in F .
Conversely, assume that F has the property that for every convergent sequence contained in F its
limit is also in F . We want to prove that X F is open.
Assume, on the contrary, that X F is not open. Then there exists a X F such that no ball
centred in a is contained in X F . Thus, for every > 0, B(a, ) F ,= .
In particular for every integer n 1 we can pick a point x
n
in B
_
a,
1
n
_
F . It is easy to check
that the sequence (x
n
) converges to a. Since (x
n
) is contained in F , by hypothesis a should be
likewise in F . This contradicts the fact that a X F .
We conclude that X F must be open.
1.6. Complete metric spaces. Complete metric spaces are spaces in which:
convergent sequences can be dened without a priori knowing their limit (using the Cauchy
property).
the Contraction Map Theorem is true. This theorem is an important tool in Analysis, in
particular in solving various types of equations.
We notice a property satised by convergent sequences.
Denition 1.48. A sequence (x
n
) in a metric space (X, d) is Cauchy if for every > 0 there exists
N N such that whenever m n N
d(x
m
, x
n
) < .
Remark 1.49. Inequality d(x
m
, x
n
) < may be replaced by d(x
m
, x
n
) without changing the
property (see Remark 1.42).
Proposition 1.50. Any convergent sequence is Cauchy.
Proof. Let (x
n
) be a sequence in a metric space (X, d) that converges to a point a X. Let be
an arbitrary positive number. There exists N such that for every n N, d(x
n
, a) <

2
. It follows
that for every m n N,
d(x
n
, x
m
) d(x
n
, a) + d(x
m
, a) <

2
+

2
= .

Denition 1.51. A metric space X is complete if every Cauchy sequence in X converges (to a point
in X).
PART A TOPOLOGY COURSE: HT 2013 13
Examples 1.52. (a) R is complete with the usual metric; so is Z, with the same metric;
(b) R 0 with the restricted Euclidean metric is not complete:
_
1
n
_
is a Cauchy sequence in
R 0, but it is not convergent in R 0.
Proposition 1.53. If (X, d
X
), (Y, d
Y
) are two isometric metric spaces and X is complete then Y
is complete as well.
Proof. Let f : Y X be an isometry. If (y
n
) is a Cauchy sequence then so is (f(y
n
)). Since
X is complete (f(y
n
)) converges to some point x X. Since f
1
is continuous, y
n
converges to
f
1
(x).
Remarks 1.54. In Proposition 1.53 the hypothesis cannot be weakened to there exists a homeomor-
phism f : Y X. See for instance the function tan :
_

2
,

2
_
R.
Proposition 1.55. A subset A of a complete metric space is complete (when endowed with the
restriction of the metric) if and only if A is closed.
Proof. Let (X, d) be a complete metric space and let A be a subset in X with the restricted metric
d
A
.
Assume that (A, d
A
) is complete. We shall prove that A is closed using Proposition 1.47.
Let (x
n
) be a sequence in A convergent to a point x X. Then (x
n
) is Cauchy in (X, d), hence
it is Cauchy in (A, d
A
). As (A, d
A
) is complete, it follows that (x
n
) converges to some a A.
The uniqueness of the limit of a convergent sequence (Proposition 1.43) implies that x = a A.
Conversely, assume that A is closed, and let us prove that (A, d
A
) is complete.
Let (x
n
) be a Cauchy sequence in (A, d
A
). Then (x
n
) is a Cauchy sequence in (X, d), which is a
complete metric space, therefore (x
n
) converges to a point x X.
The fact that A is closed and Proposition 1.47 imply that x A. Thus, (x
n
) converges in (A, d
A
)
to a point x A.
Examples 1.56. (a) the sets Q and RQ with the restricted Euclidean metric are not complete;
(b) the subspace (0, 1) of R with the usual metric is not complete.
Proposition 1.57. A Cartesian product XY of two complete metric spaces is complete (whatever
the product metric chosen on X Y ).
Proof. Let (X, d
X
) and (Y, d
Y
) be two metric spaces. On X Y the following three metrics can be
dened (see Exercise 4, Sheet 1):
d
2
((x
1
, y
1
), (x
2
, y
2
)) =
_
d
X
(x
1
, x
2
)
2
+ d
Y
(y
1
, y
2
)
2
,
d
1
((x
1
, y
1
), (x
2
, y
2
)) = d
X
(x
1
, x
2
) + d
Y
(y
1
, y
2
) ,
d

((x
1
, y
1
), (x
2
, y
2
)) = max (d
X
(x
1
, x
2
) , d
Y
(y
1
, y
2
)) .
It is easy to prove that
(1) d

d
2
d
1
2d

.
It suces to show that X Y is complete when endowed with one of the three metrics, for the
other two metrics completeness will be deduced without diculty via the sequence of inequalities (1).
If (x
n
, y
n
) is a Cauchy sequence in (X Y, d
1
), then (x
n
) is a Cauchy sequence in (X, d
X
) and
(y
n
) is a Cauchy sequence in (Y, d
Y
). Since X and Y are both complete, (x
n
) converges to a point
14 PART A TOPOLOGY COURSE: HT 2013
x X and (y
n
) converges to a point y Y . It follows that for every > 0 there exists N
1
such that
d
X
(x
n
, x)

2
whenever n N
1
, and there exists N
2
such that d
Y
(y
n
, y)

2
whenever n N
2
.
Then for every n max (N
1
, N
2
),
d
1
((x
n
, y
n
), (x, y)) = d
X
(x
n
, x) + d
Y
(y
n
, y) < .

Theorem 1.58. For any non-empty set S, the function space (B(S, R), d

) of bounded real-valued
functions on S with the sup metric is complete.
Remark 1.59. If S is nite, say S = 1, 2, .., n, the map f (f(1), ..., f(n)) is a bijection from
B(S, R) to R
n
, and via this map (B(S, R), d

) is identied with (R
n
, d

). The latter metric space


is known to be complete. Therefore what remains to be proved is the case when S is innite. This
can be seen as an innite dimensional version of the statement (R
n
, d

) is complete.
Proof. Let (f
n
) be a Cauchy sequence in (B(S, R), d

). Then for an arbitrary > 0 there exists N


such that for any m n N,
(2) |f
n
f
m
|

= sup
sS
[f
n
(s) f
m
(s)[ < .
For an arbitrary s S, [f
n
(s) f
m
(s)[ |f
n
f
m
|

. This and the above imply that for an


arbitrary s S, (f
n
(s)) is a Cauchy sequence in (R, [ [). As (R, [ [) is a complete metric space, it
follows that (f
n
(s)) converges to a real number
s
.
We dene the function f : S R, f(s) =
s
.
Consider an arbitrary > 0 and N such that whenever m n N, (2) holds.
Fix > 0 and n N. We have that for any m n, [f
n
(s) f
m
(s)[ < for every s S.
In the inequality corresponding to each s S we let m , and obtain that [f
n
(s) f(s)[ .
Then for each s S, [f(s)[ [f
n
(s) f(s)[ + [f
n
(s)[ + |f
n
|

. This implies that f is a


bounded function.
The fact that [f
n
(s) f(s)[ for every s S implies that |f
n
f|

.
We have thus obtained that for any > 0 there exists N such that for every n N, |f
n
f|

.
This implies that (f
n
) converges to f in (B(S, R), d

).
1.7. The space of bounded continuous functions. Continuous functions form a very interesting
metric space.
Let (X, d) be a metric space. Consider the set (
b
(X, R) of bounded continuous real-valued func-
tions on X. It is a subset of B(X, R), the space of bounded real-valued functions on X. Recall that
the latter is endowed with the sup-norm, inducing the sup-metric
d

(f, g) = sup
xX
[f(x) g(x)[.
Theorem 1.60. The subspace (
b
(X, R) is a closed subset in B(X, R).
Proof. We prove that (
b
(X, R) is closed in B(X, R) by proving that whenever it contains a convergent
sequence it contains its limit as well.
Let (f
n
) be a sequence in (
b
(X, R) convergent to f in B(X, R). Then for every > 0 there exists
N such that for every n N,
|f
n
f|

= sup
xX
[f
n
(x) f(x)[ < .
We want to prove that f is continuous at an arbitrary point x
0
X.
PART A TOPOLOGY COURSE: HT 2013 15
Let > 0. There exists N such that
|f
N
f|

<

3
.
By hypothesis f
N
is continuous at x
0
, therefore there exists > 0 such that d(x, x
0
) < implies
[f
N
(x) f
N
(x
0
)[ <

3
.
Assume that d(x, x
0
) < . Then
[f(x)f(x
0
)[ [f(x)f
N
(x)[ +[f
N
(x)f
N
(x
0
)[ +[f
N
(x
0
)f(x
0
)[ < 2|f
N
f|

+

3
<
2
3
+

3
= .

Corollary 1.61. The space ((


b
(X, R), d

) with the sup metric is complete.


Proof. This follows from:
the fact that (B(X, R), d

) is complete (Theorem 1.58);


the fact that a closed subset in a complete metric space is complete (Proposition 1.55);
Theorem 1.60.

Corollary 1.62. The space ([a, b] of continuous real-valued functions on [a, b] (where a < b are
real numbers) with the sup metric is complete.
Proof. We only need to recall that any continuous real-valued function on [a, b] is bounded, and apply
Corollary 1.61.
1.8. Various strong versions of continuity. We now dene some strong versions of continuity.
Denition 1.63. Let (X, d
X
), (Y, d
Y
) be metric spaces.
A map f : X Y is uniformly continuous if for every > 0 there exists > 0 such that
d
X
(x, x

) < d
Y
(f(x), f(x

)) < .
Remark 1.64. Uniform continuity is stronger than continuity.
Indeed f continuous means that for every x
0
X and for every > 0 there exists
x0
> 0
(depending on x
0
) such that d
X
(x
0
, y) <
x0
d
Y
(f(x
0
), f(y)) < .
Uniform continuity requires the existence of a uniform , the same for all points x
0
.
A particular case of a uniformly continuous map is the following.
Denition 1.65. Let (X, d
X
), (Y, d
Y
) be metric spaces. A map f : X Y is KLipschitz (or
simply Lipschitz ), for some constant K (0, ), if
d
Y
(f(x), f(x

)) Kd
X
(x, x

) , x, x

X .
Proposition 1.66. Any KLipschitz map is uniformly continuous.
Proof. For every > 0 one can take =

K
.
Examples 1.67. (a) Let f : I R be a dierentiable function, where I is an open interval in R,
such that f

is bounded, that is [f

(x)[ M for every x I . Then f is MLipschitz.


Indeed, according to the Mean Value theorem, for every x < y in I , f(x) f(y) =
f

(c)(x y), where c (x, y).


(b) The function x x

with [0, 1] is Lipschitz on [1, +).


16 PART A TOPOLOGY COURSE: HT 2013
(c) The function x 2
x
is Lipschitz on any interval [a, b] but not on R.
Indeed if the map were KLipschitz then for any x R, 2
x+1
2
x
= 2
x
would have to
be at most K, which is impossible.
(d) Given a metric space (X, d) and a point a X, the function f : X R, f(x) = d(x, a) is
1Lipschitz. This is an immediate consequence of the Triangle Inequality.
Example (d) and Proposition 1.38 imply that open balls and the complements of closed balls are
open sets. Indeed B(a, r) = f
1
(1, r) and X B

(a, r) = f
1
(r, +).
Remark 1.68. Any isometric embedding f : X Y between metric spaces is 1-Lipschitz and hence
uniformly continuous.
Denition 1.69. A map f : X X of a metric space (X, d) to itself is called a Kcontraction (or
simply a contraction) if it is KLipschitz for some constant K < 1.
Example 1.70. The cosine function on [0, 1].
According to the Mean Value Theorem, for every x, y [0, 1],
[ cos x cos y[ sin 1 [x y[ .
Theorem 1.71 (Contraction Map Theorem). Suppose that X is a complete metric space and
that f : X X is a Kcontraction, for some constant K < 1. Then
(a) f has a unique xed point, i.e. there exists a unique p X such that f(p) = p;
(b) for any x
0
X, the sequence (x
n
) dened inductively by x
n
= f(x
n1
) for all n 1
converges to p;
(c) (error estimate) d(x
n
, p)
K
n
1 K
d(x
0
, x
1
) for any n 1.
Remarks 1.72. (1) If X is not complete then Theorem 1.71 is not true. For instance the
function f : (0, 1] (0, 1] , f(x) =
x
2
is a
1
2
contraction but it has no xed point in (0, 1] .
(2) The hypothesis f is a contraction cannot be weakened to f satises the inequal-
ity:
(3) d(f(x
1
), f(x
2
)) < d(x
1
, x
2
) , for all x
1
,= x
2
in X .
For instance the function f : R R, f(x) =

x
2
+ 1 satises (3), due to the Mean Value
Theorem, and R is complete, but f has no xed point.
Remark 1.73. Theorem 1.71 is useful when trying to solve an equation x = f(x). When f is a
contraction:
Theorem 1.71, (a), implies that the equation admits a unique solution;
Theorem 1.71, (b) and (c) give a way of approximating the solution of the equation x = f(x),
and of estimating the error in the approximation.
See Exercise 9 on Problem Sheet 1, where an integral equation is solved using the contraction
mapping theorem.
PART A TOPOLOGY COURSE: HT 2013 17
Proof of Theorem 1.71.
Step 1: Existence of a xed point. Let x
0
be an arbitrary point in X, and dene the sequence
(x
n
) inductively by x
n+1
= f(x
n
).
We prove by induction that d(x
n+1
, x
n
) K
n
d(x
1
, x
0
). For n = 0 it is an equality.
Assume that the inequality is satised for n. Then
d(x
n+2
, x
n+1
) = d(f(x
n+1
), f(x
n
)) Kd(x
n+1
, x
n
) K
n+1
d(x
1
, x
0
) .
For m n N we have that
d(x
n
, x
m
) d(x
n
, x
n+1
) +d(x
n+1
, x
n+2
) + +d(x
m1
, x
m
) (K
n
+K
n+1
+ +K
m1
)d(x
1
, x
0
) .
The last term is equal to K
n 1K
mn
1K

K
n
1K

K
N
1K
.
For any > 0 there exists N large enough so that
K
N
1K
< . It follows that (x
n
) is a Cauchy
sequence. Since X is a complete metric space, (x
n
) converges to some point p X.
The function f is continuous. In the equality f(x
n
) = x
n+1
by letting n we obtain that
f(p) = p.
Step 2: Uniqueness of the xed point. Let q be another point such that f(q) = q . Then d(p, q) =
d(f(p), f(q)) Kd(p, q). If p ,= q this yields a contradiction. It follows that p = q .
This in particular implies that any sequence constructed as in Step 1 converges to the same point
p, the unique xed point of f .
Moreover, in Step 1 it was proved that for m n, d(x
n
, x
m
)
K
n
1K
. If in the last inequality we
x n and let m , we obtain d(x
n
, p)
K
n
1K
.
2. Topological spaces
2.1. Denitions and examples.
Denition 2.1. A topological space (X, T ) consists of a non-empty set X together with a family T
of subsets of X satisfying:
(T1) X, T ;
(T2) U, V T U V T ;
(T3) U
i
T for all i I
_
iI
U
i
T .
The family T is called a topology for X. The sets in T are called the open sets of X. When T is
understood we talk about the topological space X.
There are several good reasons to generalise the setting from metric spaces to topological spaces:
it is clear that two isometric metric spaces are identical from the metrical point of view; a
natural question to ask is what are the common properties of two homeomorphic metric spaces
(like the cylinder and the annulus in (2) of Examples 1.32) ?
some types of convergence cannot be dened in the setting of metric spaces, but can be dened
in the setting of topological spaces; this is the case for point-wise convergence of a sequence
of functions;
some constructions, like quotient spaces (dened in Section 5), may be endowed much more
easily with a topology than with a metric;
18 PART A TOPOLOGY COURSE: HT 2013
an important part of the theory of convergence and continuity, as well as other important
notions, may be recovered using the three axioms (T1), (T2), (T3) only. This setting may
simplify some arguments and emphasize what is really essential to them.
Examples 2.2. (1) Any metric space (X, d) gives rise to a topological space (X, T
d
), where T
d
consists of those subsets of X which are open in X with respect to the metric d.
(2) (Discrete spaces) Let X be any non-empty set. The discrete topology on X is the set of all
subsets of X.
(3) (Indiscrete spaces) Let X be any non-empty set. The indiscrete topology on X is the family
of subsets X, .
(4) Let X be any non-empty set. The co-nite topology on X consists of the empty set together
with every subset U of X such that X U is nite.
Denition 2.3. (1) We call a topological space (X, T ) metrizable if it arises as in Example 2.2,
(1), from (at least one) metric space (X, d) i.e. there is at least one metric d on X such that
T = T
d
.
(2) Two metrics on a set are topologically equivalent if they give rise to the same topology.
Example 2.4. The metrics d
1
, d
2
, d

on R
n
are all topologically equivalent - see Sheet 1, Ex. 6, (3).
We shall call the topology dened by the above three metrics the standard (or canonical ) topology
on R
n
.
Denition 2.5. Given two topologies T
1
, T
2
on the same set, we say T
1
is coarser than T
2
if T
1
T
2
.
Remark 2.6. For any space (X, T ), the indiscrete topology on X is coarser than T which in turn is
coarser than the discrete topology on X.
Denition 2.7. Let (X, T ) be a topological space.
A subset V of X is closed in X if X V is open in X (i.e. X V T ).
Examples 2.8. (a) You have to be careful which space youre saying a set is closed in.
For example in the space [0, 1) with the usual topology coming from the Euclidean metric,
[1/2, 1) is closed.
(b) In a discrete space, all subsets are closed since their complements are open.
(c) In the co-nite topology on a set X, a subset is closed if and only if it is nite or all of X.
Proposition 2.9. Let X be a topological space. Then
(C1) X, are closed in X;
(C2) if V
1
, V
2
are closed in X then V
1
V
2
is closed in X;
(C3) if V
i
is closed in X for all i I then

iI
V
i
is closed in X.
Proof. Properties (C1),(C2), (C3) follow from (T1), (T2), (T3) and from the De Morgan laws (see
the Appendix).
Denition 2.10. A sequence (x
n
) in a topological space X converges to a point x X if given any
open set U containing x there exists an integer N such that x
n
U for all n N.
PART A TOPOLOGY COURSE: HT 2013 19
Examples 2.11. (a) In a metric space this is equivalent to the metric denition of convergence by
Proposition 1.45.
(b) In a (non-empty) indiscrete topological space X any sequence converges to any point x X.
(c) In an innite space X with the co-nite topology any sequence (x
n
) of pairwise distinct
elements (i.e. such that x
n
,= x
m
when n ,= m) converges to any point x X.
Remark 2.12. Note that the uniqueness of the limit for a convergent sequence is not granted in a
topological space.
In particular (b) and (c) show that both the indiscrete topology and the co-nite topology on an
innite set are not metrizable, as in a metric space any convergent sequence has only one limit.
In order to have the uniqueness of the limit for a convergent sequence, one has to add an extra
axiom (Hausdor). We shall discuss this axiom later on.
Proposition 2.13. If a subset F in a topological space X is closed then for any convergent sequence
contained in F any limit of it is also in F .
Proof. Let (x
n
) be a convergent sequence, contained in F , and let x X be a limit of (x
n
). If
x X F then, as X F is open, it follows that there exists N such that x
n
X F for every
n N. This contradicts the hypothesis that (x
n
) is contained in F .
Remark 2.14. Unlike in the case of metric spaces, the converse of the statement in Proposition 2.13
might not be true.
Loose Remark. We shall see many implications of the following form:
topological property property in terms of sequences.
In metric spaces the converse implication may sometime be true (i.e. the topological property may
be characterized in terms of sequences).
In the general topological setting the converse implication is never true.
Denition 2.15. Suppose that (X, T
X
) and (Y, T
Y
) are topological spaces and that f : X Y is
a map. We say that f is continuous if U T
Y
f
1
(U) T
X
i.e. inverse images of open sets are
open.
If there are other topologies around we may say f is continuous with respect to T
X
, T
Y
or f is
(T
X
, T
Y
)-continuous.
Proposition 2.16. A map f : X Y of topological spaces is continuous if and only if f
1
(V ) is
closed in X whenever V is closed in Y .
Proof. This follows from the denition of a continuous map and from the formula
f
1
(Y V ) = X f
1
(V ) .

Proposition 2.17. If f : X Y and g : Y Z are continuous maps, X, Y, Z topological spaces,


then so is g f .
The proof is identical to the one of Proposition 1.40.
20 PART A TOPOLOGY COURSE: HT 2013
Denition 2.18. A homeomorphism between topological spaces X and Y is a bijection f : X Y
such that f and f
1
are continuous. The spaces X, Y are said to be homeomorphic.
Proposition 2.19. If a map f : X Y is a continuous map between two topological spaces then for
any sequence (x
n
) in X converging to a point x, (f(x
n
)) converges to f(x).
Proof. Let U be an open set containing f(x). Then f
1
(U) is an open set (f is continuous) con-
taining x.
Since (x
n
) converges to x there exists N such that for n N, x
n
f
1
(U). It follows that
f(x
n
) U .
Remark 2.20. The converse of Proposition 2.19 might not be true.
The following simple result is often a really useful way of showing that a set is open.
Lemma 2.21. Let U be a subset of a topological space. Then the following are equivalent:
(1) U is open;
(2) for every x in U , there is an open set U
x
containing x such that that U
x
U .
Proof. (1) (2): Suppose that U is open. For each x U , set U
x
= U .
(2) (1): Suppose that for each x U , there is an open set U
x
as in (2). Then clearly

xU
U
x
= U . So, U is a union of open sets, and hence is open.
2.2. Closure of a set. Property (C3) of closed sets, that is:
if V
i
is closed in X for all i I then

iI
V
i
is closed in X.
suggests the idea of approximating from above an arbitrary set by a closed set.
Denition 2.22. Consider a topological space X and a subset A X.
The closure of A in X is the set
A =

Fclosed,AF
F .
A point in A is sometimes called an adherent point of A.
Proposition 2.23. The closure A is the smallest (with respect to inclusion) closed subset
of X containing A, that is:
(a) A is closed in X and it contains A;
(b) if A V where V is closed in X then A V .
Proof. (a) is immediate from the denition of the closure and property (C3) of closed sets.
A closed subset V as in (b) appears in the intersection dening A, hence A V .
Proposition 2.24. Let A, B be subsets of a topological space X. Then
(a) A B implies that A B;
(b) A is closed in X if and only if A = A;
(c) A = A.
PART A TOPOLOGY COURSE: HT 2013 21
Proof. (a) We have A B B, we apply Proposition 2.23, (b).
(b) A A. If A is closed then Proposition 2.23, (b), with V = A implies A A, whence equality.
If A = A then A is closed.
(c) We apply (b) to A.
The following is a way of characterising points in A that is often useful.
Proposition 2.25.
A = x X : for every open U X with x U , U A ,= .
Proof. Consider A
1
= x X : for every open U X with x U , U A ,= .
We rst prove that A A
1
using Proposition 2.23, (b). Clearly A
1
contains A. It remains to
prove that A
1
is closed. We prove that X A
1
= B
1
is open, by using Lemma 2.21. Note that
B
1
= x X : there exists an open set U
x
such that x U
x
X A .
For every element y in some U
x
one can take U
y
= U
x
and have y U
y
X A. It follows that
each U
x
is entirely contained in B
1
. So, by Lemma 2.21, B
1
is open.
We have thus proved that A
1
is closed, which implies that A A
1
.
Assume that there exists x A
1
A. Since x , A it follows that there exists a closed set F
containing A such that x , F . Then x X F , and X F is an open set which does not intersect
A. This contradicts the fact that x A
1
.
Examples 2.26. (1) The closure of (a, b) in (R, [ [) is [a, b] .
(2) The closure of Q in (R, [ [) is R. Same for R Q.
(3) The closure of any set A in a space X endowed with the co-nite topology is either X if A
is innite, or A if A is nite.
Denition 2.27. A subset A in X is called dense if A = X.
We shall see in what follows that the closure of a set A contains two types of points:
points in A that are isolated;
points in A or outside A but near which points in A accumulate (i.e. accumulation
points).
Denition 2.28. A point x X such that for any open U X with x U , (U x) A ,= is
called an accumulation point (or limit point) of A.
The set of accumulation points of A is sometimes denoted by A

.
Note that A = A A

. The points in A A

are points x A such that for some open set U


containing x, U A = x. Such points are called isolated points of A.
Examples 2.29. In the metric space (R, [ [)
(1) if A = (a, b) then A

= A = [a, b] ;
(2) if A = Q or R Q then A

= A = R;
(3) if A = Z then A = Z while A

= .
(4) if A = (0, 1) 9, 10 then A = [0, 1] 9, 10, A

= [0, 1] and 9, 10 are isolated points.


22 PART A TOPOLOGY COURSE: HT 2013
Proposition 2.30. Let (X, d) be a metric space and let A X.
(1) The closure A is the set of limits of convergent sequences (a
n
) in A.
(2) The set of accumulation points A

is the set of limits of convergent sequences (a


n
) of pairwise
distinct elements in A (i.e. such that a
n
,= a
m
if n ,= m).
Proof. (1) Let A

be the set of all limits of convergent sequences (a


n
) in A.
Since A is closed and it contains A, it contains all the limits of convergent sequences (a
n
) in A,
due to Proposition 2.13. Thus A

A.
On the other hand, let x be a point in A. According to Proposition 2.25, for every integer n 1,
the open ball B
_
x,
1
n
_
intersects A in a point a
n
. The sequence (a
n
) converges to x. Hence x A

.
We have thus proved that A A

.
(2) Let A

be the set of all limits of convergent sequences (a


n
) in A such that a
n
,= a
m
whenever
n ,= m.
Let x be a point in A

, the limit of a sequence (a


n
). At most one element in the sequence can be
equal to x, thus without loss of generality we may assume that a
n
,= x for all n.
For every open set U containing x there exists N such that a
n
U for all n N. Moreover
a
n
U x. In particular (U x) A ,= . Thus x A

.
We conclude that A

.
Conversely let x be a point in A

. We construct inductively a sequence (a


n
) in A such that
d(x, a
n+1
) < d(x, a
n
) and 0 < d(x, a
n
) <
1
n
.
Since x A

, B(x, 1) x intersects A in a point a


1
.
Assume that we found the elements in the sequence a
1
, ..., a
n
satisfying the required hypotheses.
Let r
n+1
= min
_
1
n+1
, d(x, a
n
)
_
.
Since x A

, B(x, r
n+1
) x intersects A in a point a
n+1
.
The sequence (a
n
) thus constructed converges to x, and for n < m, d(x, a
m
) < d(x, a
n
), whence
a
m
,= a
n
. Therefore x A

.
We have proved that A

, which together with the converse inclusion implies that A

= A

.
Proposition 2.31. A map f : X Y of topological spaces is continuous if and only if f(A) f(A)
for every A X.
Proof. Assume that f is continuous. Since f(A) is closed, it follows that f
1
_
f(A)
_
is closed. The
latter subset also contains A, therefore it contains A. We have thus proved that A f
1
_
f(A)
_
,
which implies that f(A) f(A).
Conversely, assume that f(A) f(A) for every A X.
Let F be a closed subset in Y and let A = f
1
(F). Then f(A) F , whence f(A) f(A) F .
It follows that A f
1
(F) = A, therefore A = A. This implies that A is closed.
PART A TOPOLOGY COURSE: HT 2013 23
2.3. Interior of a set. Property (T3) of open sets, that is:
if U
i
are open for all i I then

iI
U
i
is open
suggests the idea of approximating from below an arbitrary set by an open set.
Denition 2.32. Let X be a topological space X and A a subset in X. The interior of A in X is
the set

A =
_
Uopen,UA
U .
Note that

A may be empty even if A is not. For instance consider the subset A = [0, 1] 0 in
R
2
. Clearly A contains no open ball B((x, y), ), therefore A contains no non-empty open subset.
Proposition 2.33. The interior

A is the largest (with respect to inclusion) open subset of
X contained in A, that is:
(a)

A is open and contained in A;
(b) every open subset of X contained in A lies in

A.
Both (a) and (b) follow easily from the denition of the interior.
Proposition 2.34. Let A, B be subsets of a topological space X. Then
(a) A B implies that

A

B;
(b) A is open in X if and only if

A = A;
(c)

A =

A;
Proof. (a)

A A B, and

A is open. Therefore

A

B.
(b) If A is open, since A A it follows that A

A, whence equality. The converse implication is
immediate.
(c) follows from (b) applied to

A.

Examples 2.35. The interior of [a, b] (or (a, b] ) in (R, [ [) is (a, b);
Indeed (a, b) is open and contained in [a, b] , therefore in its interior.
If a was in the interior then there would exist > 0 such that (a , a + ) is in the
interior, therefore in [a, b] . This is impossible, hence a is not in the interior. Likewise for b.
It follows that the interior is in (a, b), therefore it is equal to it.
The interiors of Z, Q, R Q in (R, [ [) are .
None of the sets above contains an open interval, therefore none contains a non-empty open
subset.
Let X be a space endowed with the co-nite topology, and let A X. The interior

A is
either equal to A if X A is nite (i.e. A is open), or it is empty.
Indeed if

A would be non-empty, then X

A would be nite.
As

A is contained in A, X A is contained in X

A, therefore it is also nite.
Next we note that the operation of taking the interior and of taking the closure are in some sense
complementary to each other.
24 PART A TOPOLOGY COURSE: HT 2013
Proposition 2.36. Let A be a subset in a topological space X.
The closure of the complement of A is the complement of the interior, i.e.
X A = X

A.
Proof. Denote X A by B.
As

A is open, and

A A, its complement X

A is closed, and it contains B. Therefore B X

A.
On the other hand B is closed and it contains B = X A. It follows that X B is open and
contained in A. Therefore X B

A. This is equivalent to X

A B.
From the two opposite inclusions above we deduce that B = X

A.

If in Proposition 2.36 we denote X A by B we obtain:


Proposition 2.37. Let B be a subset in a topological space X.
The interior of the complement of B is the complement of the closure, i.e.

X B = X B.
Note that the notation

X B simply means the interior of X B.


2.4. Boundary of a set.
Denition 2.38. The set A = A

A is called the boundary of A.
Proposition 2.39. The boundary of A equals the intersection AX A, and it equals the boundary
of X A.
Proof. Indeed A = A

A = A (X

A) = A (X A). The last inequality is due to Proposition
2.36.
Then by the above (X A) = (X A) A.
We conclude that A = (X A).
Example 2.40. Let A = [a, b), where a < b. Then A = [a, b] .
On the other hand R A = (, a) [b, +), and its closure is R A = (, a] [b, +).
It follows that A = a, b.
2.5. Separation axioms.
Denition 2.41. A topological space satises the rst separation axiom if for any two distinct points
a ,= b there exists an open set U containing a and not b.
Proposition 2.42. A topological space satises the rst separation axiom if and only if singletons
are closed subsets in it.
Proof. If every singleton x is closed in X then for any two distinct points a ,= b consider U =
X b.
Conversely, assume that X satises the rst separation axiom. We prove that V = X x is
open, for any x X.
Let y V arbitrary. Then y ,= x, hence by the rst separation axiom there exists U
y
open,
containing y, not containing x, hence contained in V .
So, by Lemma 2.21, V is open.
PART A TOPOLOGY COURSE: HT 2013 25
Denition 2.43. A topological space X is said to be Hausdor (or to satisfy the second separation
axiom) if given any two distinct points x, y in X, there exist disjoint open sets U, V with x U, y
V.
Examples 2.44. (a) Any metric space is Hausdor.
(b) An indiscrete space X with more than one point does not satisfy the rst separation axiom.
(c) An innite set with the co-nite topology is not Hausdor (consequently it is not a metrizable
topological space), but it satises the rst separation axiom because singletons are closed in
it.
Proposition 2.45. If f : X Y is an injective continuous map of topological spaces X, Y and Y
satises the rst (respectively, second) separation axiom then so does X.
Proof. Assume that Y satises the rst separation axiom. Let a, b be two distinct points in X.
Since f is injective, f(a) ,= f(b). By the rst separation axiom in Y there exists U open such that
f(a) U and f(b) , U . This implies that a f
1
(U), and b , f
1
(U). As f is continuous, f
1
(U)
is open. We have thus proved that X satises the rst separation axiom.
A similar argument shows that if Y is Hausdor then X is Hausdor. Note that we must use the
fact that U V = f
1
(U) f
1
(V ) = . See the Appendix.
Corollary 2.46. If spaces X, Y are homeomorphic then X satises the rst (second) separation
axiom if and only if Y is satises the rst (respectively the second) separation axiom.
Proposition 2.47. In a Hausdor space, a sequence converges to at most one point.
Proof. Indeed assume that a convergent sequence (x
n
) has two distinct limits a ,= b. Then by the
Hausdor property there exist U, V open and disjoint (i.e. UV = ) such that a U and b U . As
(x
n
) converges to a there exists N such that x
n
U for every n N. Similarly, (x
n
) convergent to
b implies that there exists N

such that x
n
V for every n N

. Then for every n max(N, N

),
x
n
U V , contradicting U V = .

Remark 2.48. Note that the rst separation axiom does not suce to ensure the uniqueness of the
limit of a convergent sequence. Indeed, an innite space with the co-nite topology satises the rst
separation axiom (see above), but contains sequences with several limits (see Example 2.11, (c)).
2.6. Subspace of a topological space. We saw that a subspace of a metric space is naturally
endowed with a metric structure.
Now we shall see that the same thing happens for topological structures.
Denition 2.49. Let (X, T ) be a topological space and let A be a non-empty subset of X.
The subspace or induced topology on A is T
A
= A U : U T .
The three properties (T1), (T2), (T3) that a topology must satisfy are easily checked:
(T1) = A T
A
; A = X A T
A
.
(T2) Given V
1
, V
2
in T
A
, V
1
= A U
1
and V
2
= A U
2
, where U
1
and U
2
are both in T .
Then V
1
V
2
= U
1
U
2
A T
A
.
26 PART A TOPOLOGY COURSE: HT 2013
(T3) Let (V
i
)
iI
be such that V
i
T
A
for every i I . Then for every i I there exists U
i
open
subset in X such that V
i
= A U
i
.
It follows that

iI
V
i
= A

iI
U
i
which is in T
A
because

iI
U
i
is in T .
The following is easy but useful.
Lemma 2.50. Let (X, T ) be a topological space, and let A be a non-empty subset of X. Give A
the subspace topology. Then the inclusion map i : A X is continuous.
Proof. Let U be an open subset of X. Then i
1
(X) = A X, which is open in A. So, i is
continuous.
Proposition 2.51. A subset W in A is closed in T
A
if and only if there exists F closed in (X, T )
such that W = A F .
Proof. Assume that W A is closed in T
A
. Then AW is open, i.e. there exists U open in X such
that A W = A U .
The last equality may be re-written as W = A AU = A U = A(X U). The set F = X U
is closed in X.
Conversely, let F be closed in (X, T ), and let W = A F .
Then A W = A F = A (X F). Since F is closed in X, X F is open in X, hence A W
is open in A. It follows that W is closed in A.
Proposition 2.52. Let (A, T
A
) be a subspace of (X, T ), and let B A.
(a) The closure of B in (A, T
A
), B
A
, coincides with the intersection between A and the closure
of B in X, B
X
.
(b) The interior of B in (A, T
A
),

B
A
, contains the interior of B in X,

B
X
.
Proof. (a) The set B
X
is closed in X, therefore B
X
A is closed in A, and it contains B. Therefore
B
A
B
X
A.
On the other hand B
A
is a closed set in A.
It follows by Proposition 2.51 that B
A
= F A, where F is closed in X. Since B B
A
F it
follows that B
X
F , hence B
X
A F A = B
A
.
We conclude that B
X
A = B
A
.
(b) Since

B
X
is open in X and contained in B A, it is an open set in A. Therefore

B
X

B
A
.
Remarks 2.53. (1) The inclusion

B
X


B
A
may be strict.
Consider for instance
B = (0, 1) 0 A = R 0 X = R
2
.
As B contains no open ball in R
2
,

B
X
is empty.
But B is an open set in A, therefore

B
A
= B.
PART A TOPOLOGY COURSE: HT 2013 27
(2) If A is open then

B
X
=

B
A
for every B A.
Indeed

B
A
is open in A, hence

B
A
= U A for some U open in X.
Since A is also open, it follows that

B
A
is open in X. It is also contained in B, therefore

B
A


B
X
.
This and Proposition 2.52, (b), imply equality.
We nish our discussion of subspaces of topological spaces with some easy to see but important
remarks.
Remarks 2.54. (1) If A B X and (X, T ) is a topological space then the topology induced on
A by T coincides with the topology induced on A by T
B
.
(2) If A X and (X, d) is a metric space then the topology on A induced by T
d
coincides with
the topology on A induced by the restricted metric d
A
.
This follows immediately from the fact that for any a A, the ball centred in a and of
radius r > 0 in (A, d
A
) coincides with B(a, r) A, where B(a, r) is the ball centred in a and
of radius r > 0 in (X, d).
(3) If A is open in X then T
A
is contained in T .
(4) If A is closed in X then any set closed in A is closed in X.
2.7. Basis for a topology. As in the case of vector spaces, we can consider a minimal amount of
data (a basis) via which we may recover the whole space endowed with the considered structure.
This notion is particularly useful in two settings:
that of topologies induced by metrics;
that of products of topological spaces (to be studied further on).
Denition 2.55. Given a topological space (X, T ), a subfamily B of T is called a basis for T if
every set in T can be expressed as a union of sets in B.
Example 2.56. In a metric space (X, d) the family of open balls
B = B(x, r) : x X, r > 0
is a basis for T
d
.
In particular in R the family of open intervals B = (a, b) [ a, b R, a < b is a basis for the
standard topology.
Moreover, the family of open intervals with rational endpoints B
Q
= (a, b) [ a, b Q, a < b is a
basis for the standard topology in R (Ex. 9, (2), Sheet 2).
Criterion 2.57. Let (X, T ) be a topological space, and B a basis for T .
A subset U is open in (X, T ) if and only if for every x U there exists B B such that
x B U .
Proof. Indeed, if U is open then there exists (B
i
)
iI
collection of subsets in the basis B such that
U =

iI
B
i
. This accounts for the direct implication.
Conversely assume that for every x U there exists B
x
B such that x B
x
U . Then
U

xU
B
x
U , whence U =

xU
B
x
.
28 PART A TOPOLOGY COURSE: HT 2013
Proposition 2.58. Let X, Y be topological spaces, and let B be a basis for the topology on Y .
A map f : X Y is continuous if and only if for every B in B its inverse image f
1
(B) is open
in X.
Proof. The necessary part follows from the fact that B T , i.e. a basis is composed of open sets.
The sucient part follows from:
the fact that every open set in Y can be written as

iI
B
i
, with B
i
B for every i ;
the fact that f
1
_
iI
B
i
_
=

iI
f
1
(B
i
).

Proposition 2.59. Let (X, T ) be a topological space, B a basis for T and A X. Then
(4) A = x X : for every B B with x B, B A ,= ,
(5)

A = x X : there exists B B such that x B A.
Proof. (4) Denote by A
B
the set
x X : B A ,= for any B B with x B .
Recall that according to Proposition 2.25,
A = x X : U A ,= for any U T with x U .
This and the fact that B T imply that A A
B
.
Let x be an arbitrary point in A
B
. Let U be an open set containing x. According to the Criterion
2.57, there exists B B such that x B U .
Since x A
B
, it follows that B A ,= , whence U A ,= .
We conclude that x A.
(5) We denote by I
B
the set
x X : there exists B B such that x B A.
Let x be an arbitrary point in I
B
. Then there exists B B such that x B A. Note that the
whole of B is contained in I
B
. Thus x B I
B
. This, according to the Criterion 2.57, implies that
I
B
is open.
Also I
B
is a subset of A by denition, therefore I
B


A.
Let x be an arbitrary point in

A, which is an open set. The Criterion 2.57 implies that there exists
B B such that x B

A A. It follows that x I
B
.
We may therefore conclude that

A I
B
, whence the two sets are equal.
Remark 2.60. In a metric space, (5) can be made even more precise:
(6)

A = x X : there exists > 0 such that B(x, ) A.
Proof. Let A
d
be the set
x X : there exists > 0 such that B(x, ) A.
According to (5), A
d
is contained in

A.
The converse inclusion follows from the fact that if a point x is contained in some ball B(y, r) then
B(x, ) B(y, r) for = r d(x, y).

PART A TOPOLOGY COURSE: HT 2013 29


Now we shall see:
how to recognise when a family of subsets is a basis of some topology;
how to re-construct the topology when a basis is given.
Proposition 2.61. Let X be a set and B a family of subsets of X such that
(B1) X is a union of sets in B;
(B2) the intersection B B

of any B, B

B can be expressed as a union of sets in B.


Then the family T
B
of all unions of sets in B is a topology for X. Note that B is a basis for T
B
.
Proof. (T1) The set X is contained in T
B
due to (B1).
The empty set is contained in T
B
as the union of no sets from B.
(T2) Let U, V be two sets in T
B
. Then U =

iI
B
i
and V =

jJ
B

j
, where B
j
, B

j
B.
The intersection U V is equal to

iI, jJ
B
i
B

j
. According to (B2), B
i
B

j
=

kKij
B

k
,
for some collection (B

k
)
kKij
in B. It follows that
U V =
_
iI, jJ
_
kKij
B

k
,
therefore U V is in T
B
.
Property (T3) is immediate from the denition of T
B
.
2.8. Product of topological spaces. One of the main ways of constructing new topological spaces
out of given ones is by taking Cartesian products. This is what we shall discuss now.
Proposition 2.62. Let (X, T
X
), (Y, T
Y
) be topological spaces. The family of subsets
B
XY
= U V : U T
X
, V T
Y
,
satises the conditions (B1) and (B2) in Proposition 2.61.
Proof. Indeed X Y is itself in B
XY
. This accounts for (B1).
Also, consider U V and U

, with U, U

T
X
and V, V

T
Y
. Then
(U V ) (U

) = (U U

) (V V

)
is in B
XY
. This accounts for (B2).
Denition 2.63. The family of all unions of sets from B
XY
is a topology T
XY
for XY (according
to Proposition 2.61).
We call this topology the product topology.
The space (X Y, T
XY
) is called the topological product of (X, T
X
) and (Y, T
Y
).
NB One of the commonest errors about products is to assume that any open set in the product is a
rectangular open set such as U V .
Criterion 2.64. A subset W XY is open in the product topology if and only if for any (x, y) W
there exist U T
X
, V T
Y
such that (x, y) U V W.
Proposition 2.65. Let (X Y, T
XY
) be the topological product of (X, T
X
) and (Y, T
Y
).
If A is closed in X and B is closed in Y then AB is closed in X Y .
30 PART A TOPOLOGY COURSE: HT 2013
Proof. Since A is closed, its complement U = X A is open.
Likewise B closed implies V = Y B open.
The complement (X Y ) (AB) is not U V , but
(x, y) X Y ; x , A or y , B = (U Y ) (X V ) .
Therefore it is open.
Remark 2.66. For nitely many topological spaces X
1
, ..., X
n
one can dene a topology on X
1

X
n
either by considering a basis composed of products of open sets U
1
U
n
, or by induction
on n and by identifying X
1
X
n
with (X
1
X
n1
) X
n
.
Proposition 2.67. Let (X, d
X
) and (Y, d
Y
) be two metric spaces.
The product topology of the two metric topologies T
d
X
and T
d
Y
coincides with the topology induced
by any of the product metrics (recall that they are all topologically equivalent, by Ex. 4, Sheet 1).
Proof. Consider the product metric
d

((x, y), (x

, y

)) = max [d
X
(x, x

) , d
Y
(y, y

)] .
For every U T
d
X
and V T
d
Y
, U V is in T
d
. It follows that B
XY
T
d
, whence
T
XY
T
d
, due to property (T3) of T
d
.
Conversely, let W be a set in T
d
. We prove that W is in T
XY
using the Criterion 2.64.
Let (x, y) W . Then there exists such that B
d
((x, y), ) W . But
B
d
((x, y), ) = B
d
X
(x, ) B
d
Y
(y, ) ,
and the latter set is in B
XY
.
Proposition 2.68. Let (X Y, T
XY
) be the topological product of (X, T
X
) and (Y, T
Y
), and let
AB be a subset of X Y . Then
(i) AB = AB ;
(ii)

AB =

A

B .
Proof. (i) According to Proposition 2.65, AB is a closed set. Since it also contains AB, we have
that AB AB.
Conversely, let (x, y) be a point in AB. Then for every U open in X containing x, U A ,= .
Likewise for every V open in Y containing y, V B ,= . Then (U V ) (AB) ,= .
It follows that (x, y) satises the property described in Proposition 2.59, (4), for the set AB with
respect to the basis B
XY
. Therefore (x, y) AB. We have thus proved that A B AB,
whence the equality.
(ii) The set

A

B is open and contained in AB, therefore it is contained in

AB.
Conversely, let (x, y) be a point in

AB. Proposition 2.59, (5), applied to the set A B with


respect to the basis B
XY
implies that there exist U open in X and V open in Y such that
(x, y) U V A B. This implies that x U A and y V B, whence x

A and y

B,
that is (x, y)

A

B.
We have proved that

AB

A

B, which together with the opposite inclusion yields the
equality.
PART A TOPOLOGY COURSE: HT 2013 31
Proposition 2.69. (1) If X and Y satisfy the rst separation axiom then so does their topolog-
ical product X Y .
(2) If X and Y are Hausdor spaces, so is their topological product X Y .
Proof. (1) Let (x, y) ,= (x

, y

). Then either x ,= x

or y ,= y

.
Assume that x ,= x

. Then there exists U open in X containing x and not x

. It follows that
U Y contains (x, y) and not (x

, y

). The case y ,= y

is similar.
The proof of (2) is done along the same lines, using the fact that
U V = (U Y ) (V Y ) = .

Proposition 2.70. (1) When XY is endowed with the product topology T


XY
, the projection
maps p
X
: XY X , p
X
(x, y) = x, and p
Y
: XY Y , p
Y
(x, y) = y , are continuous.
(2) ( a second way of dening the product topology) The product topology T
XY
is the coarsest
topology making p
X
and p
Y
continuous: any topology T on X Y with respect to which
p
X
and p
Y
are continuous contains T
XY
.
Proof. (1) Indeed for any open set U in X, p
1
X
(U) = U Y ; likewise for any open set V in Y ,
p
1
Y
(V ) = X V .
(2) According to the above any topology T with respect to which p
X
and p
Y
are continuous
must contain all sets U Y with U T
X
, and all sets X V with V T
Y
.
Therefore, by property (T2), T must contain all sets U V = (U Y )(XV ). Thus B
XY
T
whence, by property (T3) of T , T
XY
T .
Proposition 2.71. Let X, Y be two topological spaces.
(1) Let Z be a topological space, and let f : Z X and g : Z Y be two maps.
The map (f, g) : Z XY is continuous, with XY endowed with the product topology,
if and only if both f and g are continuous.
(2) ( a third way of dening the product topology) The product topology is the only topology for
which (1) holds for all possible Z and f, g.
Remark 2.72. Proposition 2.71, (1) can be reformulated as follows:
A map F : Z XY from a topological space Z into the topological product XY is continuous
if and only if both p
X
F : Z X and p
Y
F : Z Y are continuous.
Proof of Proposition 2.71. (1) Assume that the map (f, g) is continuous. Then f = p
X
(f, g)
and g = p
Y
(f, g) are also continuous, by Proposition 1.40.
Conversely, assume that both f and g are continuous.
For every U open in X and V open in Y ,
(f, g)
1
(U V ) = f
1
(U) g
1
(V )
is open.
This and Proposition 2.58 applied to the map (f, g) and the basis B
XY
allow us to conclude that
(f, g) is continuous.
(2) Let T be another topology on X Y for which (1) holds for all possible Z and f, g.
32 PART A TOPOLOGY COURSE: HT 2013
We take Z = X Y with the product topology T
XY
and the maps f = p
X
and g = p
Y
continuous. Then (p
X
, p
Y
) = id
XY
is a continuous map between (XY , T
XY
) and (XY , T ).
It follows that for every open set U T , its inverse image by id
XY
, which is U , is contained in
T
XY
.
We have thus proved that T T
XY
.
Clearly id
XY
= (p
X
, p
Y
) is a continuous map between (X Y , T ) and (X Y , T ). This and
the property (1) of T implies that p
X
and p
Y
are continuous also when X Y is endowed with the
topology T .
Proposition 2.70, (2), implies that T
XY
T , whence equality.
3. Connected spaces
3.1. Denition and properties of connected spaces. We now dene the important notion of
connected space.
This notion allows one to dene intervals in R in terms of the topology on R (instead of the
order on R). This was seen last term and shall be recalled here.
In various problems in Analysis and Geometry one needs to work in a set A such that one
can move continuously inside A from an arbitrary point of A to another (i.e. a set A that is
path-connected).
In real life too such questions naturally occur. If one can only travel by car for instance, one
is bound to consider the topological space X composed of all the points above the sea level
and make ones travels plans within a path-connected component A of X.
A weaker condition which is also much used in Analysis and Geometry is that A cannot be
written as the disjoint union of two open non-empty sets (i.e. A is connected, as opposed to
disconnected).
We shall see that the two conditions of connectedness and path-connectedness are not equivalent
in general.
Denition 3.1. A topological space X is disconnected if there are disjoint open non-empty subsets
U and V such that U V = X.
If X is not disconnected, it is called connected.
Proposition 3.2. Let X be a topological space. The following properties are equivalent:
(1) the only subsets of X which are both open and closed are X and ;
(2) X is connected;
(3) any continuous map from X to 0, 1 (with the discrete topology) is constant.
Proof. (1) (2) If U, V are open sets in X such that U V = and U V = X, then U is both
open and closed (as X U = V ). It follows by (1) that either U = or U = X, hence V = .
(2) (3) Let f : X 0, 1 be a continuous map. Then U = f
1
(0) and V = f
1
(1) are
open, disjoint subsets of X, and U V = X. So, by denition of connectedness, one of U and V is
empty, and therefore f is not surjective.
(3) (1) Let A be a subset of X both open and closed. Then the map f : X 0, 1
such that f(x) = 1 if x A and f(x) = 0 if x , A is continuous: f
1
(1) = A which is open,
f
1
(0) = X A, which is open.
PART A TOPOLOGY COURSE: HT 2013 33
According to (3) either f is constant 1, which means that A = X, or f is constant 0, which
means that A = .
Denition 3.3. A (non-empty) subset A of a topological space X is connected if A with the subspace
topology is connected.
Conventionally an empty set is assumed to be connected.
Remark 3.4. The connectedness of a subset A can be formulated in terms of the topology on X
(rather than the subspace topology) as follows:
A is connected if and only if, whenever U and V are open subsets of X such that A (U V )
and U V A = , then either U A = or V A = .
The following was proved in the Michaelmas term Analysis course.
Proposition 3.5. Let R be endowed with the standard topology and let A be a subset in R. The
following statements are equivalent:
(1) A is a connected subset of R ;
(2) A is an interval, i.e. a set of the form
(, b), (, b], (a, b), [a, b), (a, b], [a, b], (a, ), [a, ), , and R,
where b > a for intervals with endpoints a, b except for [a, b] , where b a.
Proposition 3.6. If f : X Y is a continuous map of topological spaces X, Y and A X is
connected then so is f(A). (The continuous image of a connected set is connected.)
Proof. Let U and V be open sets in Y such that f(A) U V and f(A) U V = . Then f
1
(U)
and f
1
(V ) are open in X. So, f
1
(U) A and f
1
(V ) A are disjoint and open in A. Since A
is connected, one of f
1
(U) A and f
1
(V ) A is empty. Hence, one of U f(A) and V f(A) is
empty. So, f(A) is connected.
We saw that several topological properties (openness, closedness etc) are preserved under nite
intersections and unions. This is no longer true for connectedness: one can easily draw examples
disproving the statement both for nite intersections and unions.
Still, under some extra conditions, connectedness can be granted for unions of connected sets.
Proposition 3.7. Suppose that A
i
: i I is a family of connected subsets of a topological space X
with A
i
A
j
,= for each pair i, j I .
Then

iI
A
i
is connected.
Proof. Let f :

iI
A
i
0, 1 be a continuous map. Its restriction to A
i
, f[
Ai
, is also continuous.
Since A
i
is connected it follows that f[
Ai
is constant equal to c
i
0, 1.
For each pair i, j I there exists a point x in A
i
A
j
, and f(x) = c
i
= c
j
. Therefore all the
constants c
i
are equal, and f is constant.
Corollary 3.8. If C
i
: i I and B are connected subsets of a topological space X such that
C
i
B ,= for each i I , then B

iI
C
i
is connected.
34 PART A TOPOLOGY COURSE: HT 2013
Proof. Proposition 3.7 implies that A
i
= C
i
B is connected for every i I .
Since A
i
A
j
contains B for every i, j , again by Proposition 3.7 it follows that

iI
A
i
=
B

iI
C
i
is connected.
Theorem 3.9. The topological product XY is connected if and only if both X and Y are connected.
Remark 3.10. An inductive argument allows one to extend Theorem 3.9 to any nite product of
topological spaces: a topological product X
1
X
n
is connected if and only if X
1
, ..., X
n
are
connected.
Proof of Theorem 3.9. If X Y is connected then X = p
X
(X Y ) and Y = p
Y
(X Y ) are
connected by Proposition 3.6.
Assume that X and Y are connected. Then X y is connected for every y Y , and x Y
is connected for every x X.
Fix a point y
0
Y . Corollary 3.8 applied to B = X y
0
and to C
x
= x Y , connected and
such that B C
x
= (x, y
0
) implies that B

xX
C
x
= X Y is connected.
Theorem 3.11. Suppose that A is a connected subset of a topological space X and A B A.
Then B is connected.
Proof. Let U and V be open subsets of X such that B U V and B U V = . We will show
that one of U B and V B is empty.
Now, A U V and AU V BU V = . So, by the connectedness of A, one of U A and
V A is empty. Say that U A is empty. Then, A lies in the closed set XU . So, by Proposition
2.23, A XU . So, U A = . Since B A, we deduce that U B is empty, as required.
3.2. Path-connected spaces.
Denition 3.12. A path connecting two points x, y in a topological space X is a continuous map
p : [0, 1] X with p(0) = x, p(1) = y.
Remark 3.13. If p : [0, 1] X is a path then p ([0, 1]) is a connected set.
Remark 3.14. If p : [0, 1] X is a path connecting x, y and q : [0, 1] X is a path connecting y, z
then we may construct a path connecting x, z as follows: r : [0, 1] X,
r(t) =
_
_
_
p(2t) if t
_
0,
1
2

,
q(2t 1) if t
_
1
2
, 1

.
The restriction of r to
_
0,
1
2

is continuous because p is continuous and the map t 2t is


continuous. We denote this restriction by r
1
.
Likewise, the continuity of q and of the map t 2t 1 implies that the restriction of r to
_
1
2
, 1

is continuous. We denote the latter restriction by r


2
.
For any closed set F in X, r
1
(F) = r
1
1
(F) r
1
2
(F).
We have that r
1
1
(F) is closed in
_
0,
1
2

, which is closed in [0, 1] , therefore r


1
1
(F) is closed in
[0, 1] . A similar argument yields that r
1
2
(F) is closed in [0, 1] .
We conclude that r
1
(F) is closed in [0, 1] .
PART A TOPOLOGY COURSE: HT 2013 35
Denition 3.15. A topological space X is path-connected if any two points in X are connected by
a path in X.
A subset A X is path-connected if with the subspace topology it satises the previous condition.
Equivalently, A is path-connected if any two points in A can be joined by a path p: [0, 1] X with
image in A. Conventionally the empty set is path-connected.
Examples 3.16. (1) Given a continuous function f : I R, I R an interval, the graph
G
f
= (x, f(x)) [ x I is path-connected.
For instance consider the function f : R R,
f(x) =
_
xsin
1
x
if x ,= 0 ,
0 if x = 0 .
(2) Any normed vector space (V , | |) is path-connected.
Indeed, consider u, v two vectors in V . Let p : [0, 1] V be dened by
p(t) = (1 t)u +tv .
Clearly p(0) = u and p(1) = v. Moreover
|p(s) p(t)| = |(t s)u (t s)v| = [t s[|u v| .
It follows that p is KLipschitz, with K = |u v|.
(3) Any ball in a normed vector space is path-connected.
According to Remark 3.14 it suces to prove that any vector u in a ball B(a, R) is
connected by a path to a.
We consider the path p : [0, 1] V dened by p(t) = (1 t)u +ta.
For every t [0, 1] , |p(t) a| = [1 t[ |u a| |u a| < R. Thus the image of the
path p is entirely contained in B(a, R).
Proposition 3.17. Any path-connected space is connected.
Proof. Let X be a path-connected space and let a be a xed point in X. For every x X there
exists a path p
x
: [0, 1] X such that p
x
(0) = a and p
x
(1) = x. By Remark 3.13, T
x
= p
x
[0, 1] is a
connected set.
We apply Corollary 3.8 to the collection of connected sets T
x
: x X and to the connected set
a, and conclude that

xX
T
x
a = X is connected.
Remark 3.18. There exist connected spaces that are not path-connected.
Example 3.19. Let G
f
be the graph of the function f : (0, ) R, f(x) = sin
1
x
, that is the set
G
f
=
__
x, sin
1
x
_
[ x > 0
_
.
Clearly G
f
is path-connected, hence connected.
The closure of G
f
in R
2
with the standard topology, G
f
, is equal to G
f
(0 [1, 1]). It is
connected but not path-connected.
Remark 3.20. The example 3.19 also shows that A path-connected does not in general imply A
path-connected.
36 PART A TOPOLOGY COURSE: HT 2013
In the above statement two things must be proved:
(i) that G
f
= G
f
(0 [1, 1]);
(ii) that G
f
= G
f
(0 [1, 1]) is not path-connected.
(i) First we prove that G
f
G
f
(0 [1, 1]).
Let (a, b) be an arbitrary point in G
f
.
Proposition 2.30, (1), implies that (a, b) is the limit of a sequence
_
x
n
, sin
1
xn
_
with x
n
> 0 (we
are in the metric space R
2
). This implies that (x
n
) converges to a, therefore a must be in [0, +).
If a (0, +) then by the continuity of the map x sin
1
x
we have that the limit b of sin
1
xn
must be sin
1
a
. Thus in this case (a, b) G
f
.
Assume that a = 0. The sequence
_
sin
1
xn
_
converges to b and it is contained in [1, 1] closed in
R, therefore b [1, 1] .
We have thus found that G
f
G
f
(0 [1, 1]).
For the converse inclusion, we take an arbitrary point (0, b) 0 [1, 1] .
The sequence
_
x
n
, sin
1
xn
_
nN
in G
f
with x
n
=
1
arcsin b+2n
converges to (0, b). Therefore (0, b)
G
f
, hence 0 [1, 1] G
f
.
(ii) Assume that G
f
= G
f
(0 [1, 1]) is path-connected. In particular there is a path
p : [0, 1] G connecting (0, 0) and (1, sin 1).
Since p is continuous and F = 0 [1, 1] is closed, we have that p
1
(F) is closed in [0, 1] . Let
T be the supremum of p
1
(F). This in particular implies that p(T, 1] F = .
As p
1
(F) is closed, its supremum T must be contained in it, therefore p(T) = (0, b) for some
b [1, 1] .
The path p is continuous at T hence for =
1
4
there exists > 0 such that
s [T, T +] , |p(s) p(T)| <
1
4
.
We may assume that is such that T + < 1 and we denote T + by S.
We have that p(S) G
f
, thus p(S) =
_
x
0
, sin
1
x0
_
for some x
0
> 0.
Let : R
2
R be the projection (x, y) = x. It is a continuous map, and so is p, therefore p
is continuous. The interval [T, S] is a connected set, therefore p[T, S] is connected.
It follows that p[T, S] is an interval containing 0 and x
0
> 0, thus containing [0, x
0
] .
Then for every x (0, x
0
] there exists a point in p[T, S] contained in
1
(x).
On the other hand p[T, S] is contained in G
f
and the latter intersects
1
(x) in a unique point,
_
x, sin
1
x
_
.
We then conclude that p[T, S] contains
__
x, sin
1
x
_
: x (0, x
0
]
_
.
According to the choice of S, p[T, S] should be contained in the ball B
_
p(T) ,
1
4
_
. In particular
any two points in p[T, S] should be at distance
1
2
, by the triangle inequality.
On the other hand, for n large enough x
n
=
1

2
+2n
and y
n
=
1
3
2
+2n
are both in (0, x
0
] .
PART A TOPOLOGY COURSE: HT 2013 37
It follows that p[T, S] contains the points
_
x
n
, sin
1
xn
_
= (x
n
, 1) and
_
y
n
, sin
1
yn
_
= (y
n
, 1),
which are at distance at least 2. This gives a contradiction.
Nevertheless the implication connected path-connected holds for a particular type of topological
space.
Proposition 3.21. A connected open subset U of a normed vector space is path-connected.
Remark 3.22. The hypothesis U open in Proposition 3.21 cannot be removed, as shown by the
Example 3.19 (U = G
f
is a connected subset of the normed vector space (R
2
, | |)).
Proof. Let a be a xed point in U and let C be the set of elements x U connected to a by a path
in U .
The set C contains a, therefore it is non-empty. We prove that C is both open and closed in U .
First, we show that C is open. Indeed, consider a point x in C. Since x U which is open,
there exists > 0 such that B(x, ) U . All points in B(x, ) can be connected by a path to x, by
Example 3.16, (3). This and Remark 3.14 imply that B(x, ) C. Hence, C is open.
Now we prove that C is closed in U by proving that U C is open.
Let x be a point in U C. Since x is in U , there is some > 0 such that B(x, ) U . We claim
that B(x, ) is disjoint from C. This will show that B(x, ) U C, and hence that U C is open.
Suppose that there were a point y in B(x, )C. Then, x can joined to y by a path in B(x, ) U ,
using Example 3.16, (3). Also, y can be joined to a by a path in U , since y is in C. Hence, Remark
3.14 implies that there is a path in U joining x to a, which contradicts the assumption that x is not
in C.
Thus C is a non-empty, closed and open subset in U connected, therefore C = U . This and
Remark 3.14 imply that U is path-connected.
Proposition 3.23. If f : X Y is a continuous map of topological spaces X, Y and A X is
path-connected then so is f(A). (The continuous image of a path-connected set is path-
connected.)
Proof. For every two elements x, y f(A) let a, b be two elements in A such that f(a) = x and
f(b) = y.
The set A is path-connected, therefore there exists a path p : [0, 1] A connecting a and b. Then
f p is a path in f(A) connecting x and y.
4. Compact spaces
Compact sets are a particular type of bounded closed set, very much used in all branches of
mathematics. Here are two reasons why this notion is interesting:
in Analysis it is important to know whether there exists a certain object minimizing a given
functional or not (for instance in calculus of variations problems);
from the topological point of view, as it will become clear from the theory, compact sets are
the next best thing, after singletons and nite sets.
38 PART A TOPOLOGY COURSE: HT 2013
4.1. Denition and properties of compact sets.
Denition 4.1. A family U
i
: i I of subsets of a space X is called a cover if X =
_
iI
U
i
.
If each U
i
is open in X then | is called an open cover for X.
Denition 4.2. A subcover of a cover U
i
: i I for a space X is a subfamily U
j
: j J for
some subset J I such that U
j
: j J is still a cover for X.
We call it a nite (or countable) subcover if J is nite (or countable).
Denition 4.3. A topological space X is compact if any open cover of X has a nite subcover.
Proposition 4.4. Let X be a topological space. The following are equivalent:
(1) X is compact;
(2) if V
i
: i I is an indexed family of closed subsets of X such that

jJ
V
j
,= for any nite
subset J I then

iI
V
i
,= .
Proof. (1) (2) Assume that V
i
: i I is a family of closed subsets of X such that

jJ
V
j
,=
for any nite subset J I , while

iI
V
i
= . Then

iI
(X V
i
) = X.
According to (1) there exists J nite subset in I such that

jJ
(X V
j
) = X. This is equivalent
to

jJ
V
j
= , contradicting the hypothesis.
(2) (1) Let U
i
: i I be an open cover of X. Then X =

iI
U
i
, whence

iI
(X U
i
) = .
According to property (2) applied to the family of closed sets X U
i
; i I there exists some
nite subset J of I such that

jJ
(X U
j
) = . Equivalently

jJ
U
j
= X.
Denition 4.5. A subset A of a topological space X is compact if it is compact when endowed with
the subspace topology.
Remark 4.6. In view of the way in which the subspace topology is dened, a subset A of a topological
space X is compact if and only if for every family U
i
: i I of open sets in X such that A

iI
U
i
there exists a nite subset J I such that A

jJ
U
j
.
This can be rephrased in terms of open covers of A, as follows.
Denition 4.7. A family U
i
: i I of subsets of a space X is called a cover for A if A
_
iI
U
i
.
We also sometimes say that the sets U
i
, i I , cover the set A.
Thus, a subset A of X is compact if and only if every family of open sets in X which form a cover
for A have a nite subcover.
Examples 4.8. (1) Any nite space X is compact.
PART A TOPOLOGY COURSE: HT 2013 39
(2) If (x
n
)
nN
is a sequence converging to x in a topological space then the set
L = x
n
[ n N x
is compact.
Indeed consider U
i
: i I, an open cover of L. There exists k I such that x U
k
.
The sequence (x
n
) converges to x therefore there exists N such that x
n
U
k
for every
n N. For every n 1, ..., N, x
n
is contained in some U
in
, i
n
I . Therefore L is
contained in U
i1
U
i
N
U
k
.
(3) Any (non-empty) set with the conite topology is compact. Likewise for the indiscrete topology.
Thus any set X can be endowed with a topology T such that (X, T ) is compact.
The statement for the indiscrete topology follows from the fact that there are only two
open sets in that topology.
The statement for the co-nite topology is easier to check with Proposition 4.4, (2). Indeed
consider a family F
i
: i I of closed sets in X such that any intersection of nitely many
of those sets is non-empty.
Pick k I and consider F
k
= x
1
, ..., x
m
. Assume that for every x
j
there exists F
ij
not
containing it. Then F
k
F
i1
F
im
is empty. This contradicts the hypothesis.
Therefore some x
j
is contained in all F
i
with i I . It follows that

iI
F
i
contains x
j
.
The following was proved in the Michaelmas term Analysis course.
Proposition 4.9. (the Heine-Borel Theorem) Any interval [a, b] in R is compact.
Remark 4.10. An open interval (a, b), where a < b, is not compact in R (with the Euclidean topology).
Nor are the intervals [a, ) or (, b] .
Indeed (a, b) =

nN
_
a +
1
n
, b
1
n
_
. Any nite subfamily
_
a +
1
n1
, b
1
n1
_
, ......,
_
a +
1
n
k
, b
1
n
k
_
(for which we may assume that n
1
< ... < n
k
, otherwise we change the order) covers only
_
a +
1
n
k
, b
1
n
k
_
.
A similar argument can be made for [a, ) =

nN
[a, n) and for (, b] =

nN
(n, b] .
Proposition 4.11. Any closed subset A of a compact space X is compact.
Proof. Indeed let V
i
: i I be an indexed family of closed subsets of A such that

jJ
V
j
,= for
any nite subset J I .
As A is closed in X, V
i
are also closed in X, which is compact. Therefore

iI
V
i
,= .
Remark 4.12. The converse of Proposition 4.11 may not be true: a singleton in X is always compact
but may not be closed.
As the example suggest, a separation axiom must be added in order to have the converse of
Proposition 4.11.
Proposition 4.13. Let X be a Hausdor space.
If K is a compact subset of X and x X K then there exists U, V disjoint open sets such that
K U and x V .
40 PART A TOPOLOGY COURSE: HT 2013
Proof. For every y K, according to the Hausdor property, there exists a pair of disjoint open sets
U
y
and V
y
, y U
y
, x V
y
.
The collection U
y
: y K is an open cover for K. Therefore there exist y
1
, ..., y
m
such that
K U
y1
U
ym
. Let U = U
y1
U
ym
.
The set V = V
y1
V
ym
is open, it contains x, and V U = V

m
j=1
U
yj
= .
Corollary 4.14. Any compact subset K of a Hausdor space X is closed in X.
Proof. According to Proposition 4.13, for every x X K, there exists V open such that x V
X K. This implies that X K is open, by Lemma 2.21.
Proposition 4.15. Let X be a topological space.
(1) If K
1
, ..., K
n
are compact subsets in X then K
1
K
n
is compact in X.
(2) If K
i
: i I is a non-empty family of compact subsets and X is Hausdor then

iI
K
i
is compact.
Proof. (1) Let U
i
: i I be an open cover of K
1
K
n
. In particular for every t 1, ..., n,
K
t

iI
U
i
. As K
t
is compact, there exists a nite subset J
t
of I such that K
t

jJt
U
j
.
It follows that K
1
K
n

n
t=1

jJt
U
j
.
(2) Since X is Hausdor, every K
i
is a closed set according to Corollary 4.14. Therefore
L =

iI
K
i
is also closed. It is also contained in some (each) compact space K
i
, hence by Proposition
4.11, L is compact.
Remark 4.16. (1) Property (1) is no longer true for innite unions.
Example:

nN
[0, n] = [0, ).
(2) Property (2) is no longer true if X is not Hausdor (see Ex. 5, (2), Sheet 4).
Proposition 4.17. A compact subset A of a metric space X is bounded.
Proof. Let x be a point in X. We have that A X =

nN
B(x, n).
It follows that there exist n
1
< ... < n
k
such that A B(x, n
1
) B(x, n
k
) = B(x, n
k
).
Corollary 4.18. A compact subset of a metric space Y is bounded and closed in Y .
PART A TOPOLOGY COURSE: HT 2013 41
4.2. Compact spaces and continuous maps.
Proposition 4.19. If f : X Y is a continuous map of topological spaces and A X is compact
then so is f(A). (The continuous image of a compact set is compact.)
Proof. Let U
i
: i I be an open cover of f(A).
Then f
1
(U
i
) : i I is an open cover of A.
The compactness of A implies that there exists a nite subcover f
1
(U
j
) : j J, J I .
It follows that U
j
: j J is a nite subcover for f(A).
Corollary 4.20. Suppose that X is a compact space, Y is a metric space and f : X Y is
continuous.
Then f(X) is bounded and closed in Y . (When f(X) is bounded we say f is bounded.)
Corollary 4.21 (the extreme value theorem). If f : X R is a continuous real-valued function on
a compact space X then f is bounded and attains its bounds, i.e. there exist m, M X such that
f(m) f(x) f(M) for every x X.
Proof. Since X is compact, f(X) is compact, in particular it is bounded and closed.
Let a, A be the highest lower bound, respectively the least upper bound of f(X).
Since f(X) is closed in R, it follows that both a and A are in f(X).
Remark 4.22. Another consequence of Proposition 4.19 is the following: if A X Y , where Y is a
topological space and X is endowed with the subspace topology, then A compact in X A compact
in Y . This follows by using Proposition 4.19 for the inclusion map i : X Y .
The same implication does not hold when compact is replaced by closed: (0, 1] is closed in (0, 2)
but not in R.
The converse implication holds both for closed and for compact, i.e. A compact (closed) in
Y A compact (closed) in X.
Theorem 4.23. The product X Y is compact if and only if both X and Y are compact.
Proof. If the product X Y is compact then X = p
X
(X Y ) and Y = p
Y
(X Y ) are compact,
due to Proposition 4.19.
Conversely, assume that X and Y are compact.
Let W
i
: i I be an open cover of X Y .
Since the rectangular open sets U V compose a basis for the product topology, we may write
X Y =
_
iI
W
i
=
_
jJ
U
j
V
j
,
where each U
j
V
j
is contained in some W
i
.
Therefore it is enough if we prove that the open cover U
j
V
j
: j J has a nite subcover.
For every y Y , the compact set X y is covered by U
j
V
j
: j J. Therefore there exists
a nite subset F
y
J such that
X y
_
jFy
U
j
V
j
.
42 PART A TOPOLOGY COURSE: HT 2013
The set V
y
=

jFy
V
j
is an open set containing y. The family V
y
: y Y is an open cover of
Y , which is compact.
It follows that there exist y
1
, ..., y
m
such that Y = V
y1
V
ym
.
We state that X Y =

m
k=1

jFy
k
U
j
V
j
. Indeed consider an arbitrary element (x, y) in the
product. Then there exists k 1, ..., m such that y V
y
k
. In particular y V
j
for all j F
y
k
.
On the other hand X y
k


jFy
k
U
j
V
j
, whence there exists j
0
F
y
k
such that x U
j0
. It
follows that (x, y) U
j0
V
j0
.
Theorem 4.24. (general Heine-Borel) Any closed bounded subset of (R
n
, | |
i
), where i
1, 2, , is compact.
Proof. Let A be a closed bounded subset of R
n
. Since A is bounded, there exists x = (x
1
, ..., x
n
)
and r > 0 such that A B
d
(x, r).
The ball B
d
(x, r) in (R
n
, | |

) can be rewritten as
B
d
(x, r) = (x
1
r, x
1
+r) (x
n
r, x
n
+r) [x
1
r, x
1
+r] [x
n
r, x
n
+r] .
Each interval [x
i
r, x
i
+r] is compact by Proposition 4.9, and their product is compact by Theorem
4.23.
It follows that A is a closed subset contained in a compact space, therefore it is compact by
Proposition 4.11.
N.B. Theorem 4.24 fails in general when (R
n
, | |
i
) is replaced by an arbitrary metric
space (X, d):
an easy example of this is (0, 1] - its bounded, and closed in (0, 2), but its not compact;
another such example, in which the ambient space is a normed vector space, moreover complete,
will be given in Exercise 2, Sheet 4.
The inverse of a continuous map is not necessarily continuous, as Examples 1.35 shows. Still, the
statement is true under some extra conditions.
Theorem 4.25. Suppose that X is a compact space, that Y is a Hausdor space and that f : X Y
is a continuous bijection. Then f is a homeomorphism (i.e. f
1
is continuous too).
Proof. Let V be a closed subset of X. Then V is compact, by Proposition 4.11. It follows that f(V )
is compact, by Proposition 4.19. Since Y is Hausdor, this implies that f(V ) is closed, according to
Corollary 4.14.
Corollary 4.26. (a) Let X be a space, and let T
H
T
K
be two topologies such that (X, T
H
) is
Hausdor and (X, T
K
) compact. Then T
H
= T
K
.
(b) Let (X, T ) be a Hausdor compact space. Then any strictly coarser topology T

T is no
longer Hausdor, and any strictly ner topology T

T is no longer compact.
Proof. (a) We apply Theorem 4.25 to the map id : (X, T
K
) (X, T
H
).
(b) is a reformulation of (a).
Quirky application 4.27. No topology on [0, 1] which is strictly coarser than the Euclidean topology
can be Hausdor. No topology which is strictly ner than the Euclidean topology can be compact.
PART A TOPOLOGY COURSE: HT 2013 43
Corollary 4.28. Suppose that X is a compact space, that Y is a Hausdor space and that f : X Y
is injective and continuous.
Then f determines a homeomorphism of X onto f(X).
Remark 4.29. This generalises the Mods result: if f : [0, 1] R is continuous and monotonic then
its inverse function f
1
: f([0, 1]) [0, 1] is continuous.
An important property of continuous maps dened on compact metric spaces is the following.
Theorem 4.30. If f : X Y is a continuous map of metric spaces and X is compact then f is
uniformly continuous.
Proof. Let be an arbitrary positive number (which we consider as xed in the sequel).
Let x be an arbitrary point in X. The map f is continuous at x. Therefore, for the given > 0
there exists
x
such that
(7) d(y, x) <
x
d(f(y), f(x)) <

2
.
We have that X =

xX
B
_
x,
x
2
_
. The space X being compact, there exist x
1
, ..., x
n
such that
X =

n
i=1
B
_
x
i
,
x
i
2
_
.
Let = min
_
x
1
2
, ....,
xn
2
_
. We shall prove that d(y, z) < d(f(y), f(z)) < .
Indeed, let y, z be two points such that d(y, z) < . As y

n
i=1
B
_
x
i
,
x
i
2
_
, there exists
k 1, ..., n such that y B
_
x
k
,
x
k
2
_
, that is d(y, x
k
) <
x
k
2
.
Then d(z, x
k
) d(z, y) + d(y, x
k
) < +
x
k
2
<
x
k
.
It follows according to (7) that d(f(y), f(x
k
)) <

2
and d(f(z), f(x
k
)) <

2
, whence
d(f(y), f(z)) d(f(y), f(x
k
)) + d(f(z), f(x
k
)) < .

4.3. Sequential compactness. Loose Remark. We already saw that in a metric space many
topological properties can be characterised in terms of sequences.
In what follows we show that one of these properties is compactness.
Denition 4.31. A topological space X is sequentially compact if every sequence in X has a subse-
quence converging to a point in X.
A (non-empty) subset A of a topological space X is sequentially compact if, with the subspace
topology, A is sequentially compact. (Conventionally, the empty set is sequentially compact.)
Theorem 4.32. (1) (Bolzano-Weierstrass theorem) In a compact topological space X every
innite subset has accumulation points.
(2) Every compact metric space is sequentially compact.
44 PART A TOPOLOGY COURSE: HT 2013
Proof. (1) We prove that if, for a subset A of X, the set of accumulation points A

is empty then A
is nite.
Recall that the closure A is equal to A A

. Therefore if A

is empty then A = A = A A

.
The set AA

is the set of isolated points of A, hence all points of A are isolated: for every a A
there exists U
a
open set such that U
a
A = a.
We have that A = A

aA
U
a
. The set A is a closed set in a compact metric space, therefore
it is compact by Proposition 4.11. The Borel-Lebesgue property of compact sets implies that there
exist a
1
, ..., a
n
in A such that A = A

n
i=1
U
ai
.
We may then write A =

n
i=1
(U
ai
A) =

n
i=1
a
i
= a
1
, ..., a
n
.
(2) Consider (x
n
) a sequence in a compact metric space X. Let A = x
n
: n N.
Assume that A is nite, A = a
1
, ..., a
k
. For every a
i
we consider the set N
i
= n N : x
n
= a
i
.
Since N = N
1
. N
2
. . N
k
, at least one of the sets N
i
is innite. It follows that (x
n
) has a sub-
sequence which is constant, therefore convergent.
Assume that A is innite. Then according to (1), A

,= . Let a be a point in A

. For every open


set U containing a, (U a) A ,= .
In particular there exists
a term x
n1
of the sequence contained in [B(a, 1) a] A,
a term x
n2
in
_
B
_
a,
1
2
_
a, x
1
, ..., x
n1

A, etc.
a term x
n
k
in
_
B
_
a,
1
k
_
a, x
1
, ..., x
n
k1

A.
In other words, we construct inductively a subsequence of (x
n
) converging to a.
Theorem 4.33. Any compact metric space is complete.
Proof. Let (x
n
) be a Cauchy sequence in a compact metric space (X, d).
According to Theorem 4.32, (2), (x
n
) has a subsequence (x
(n)
) convergent to a point a, where
: N N is a one-to-one increasing map. We prove that (x
n
) converges to a.
Consider an arbitrary > 0. The sequence (x
n
) is Cauchy, therefore there exists N such that for
every n, m N, d(x
n
, x
m
) <

2
.
The subsequence (x
(n)
) converges to a, therefore there exists M such that for every n M,
d
_
x
(n)
, a
_
<

2
.
Let k M large enough so that (k) N. Then for every n N,
d(x
n
, a) d
_
x
n
, x
(k)
_
+ d
_
x
(k)
, a
_
<

2
+

2
= .

Theorem 4.34. Any sequentially compact metric space is compact.


N.B. This and Theorem 4.32, (2), imply that a metric space is compact if and only if it is
sequentially compact.
The proof is done in two steps, which we formulate as two separate Propositions.
Proposition 4.35. Let X be a sequentially compact metric space. For any open cover | of X there
exists > 0 such that for every x X, B(x, ) is entirely contained in some single set from | .
PART A TOPOLOGY COURSE: HT 2013 45
Denition 4.36. The number in Proposition 4.35 is called a Lebesgue number for the cover | .
Proof. Assume that for every n N there exists a point x
n
X such that the ball B
_
x
n
,
1
n
_
is
contained in no subset U | .
The sequence (x
n
) has a subsequence (x
(n)
) converging to some point a. There exists a subset
U in the cover | containing a. As U is open, it also contains a ball B(a, 2), for > 0.
The subsequence (x
(n)
) converges to a, hence there exists N such that if n N then x
(n)

B(a, ). This implies that B
_
x
(n)
,
_
B(a, 2) U .
For n large enough
1
(n)
< , whence B
_
x
(n)
,
1
(n)
_
B
_
x
(n)
,
_
U .
The last inclusion contradicts the choice of the sequence (x
n
).
Denition 4.37. Given a real number > 0, an -net for a metric space X is a subset N X
such that B(x, ) : x N is a cover of X.
Example 4.38. For every >

n
2
, Z
n
is an -net for R
n
with the Euclidean norm | |
2
.
Proposition 4.39. Let X be a sequentially compact metric space and let be an arbitrary positive
number.
Then there exists a nite -net for X.
Remark 4.40. Thus compact metric spaces can be approximated by nite sets.
This illustrates once more our remark that from the topological point of view compact sets are the
next best thing, after singletons and nite sets.
Proof. Assume that there exists no nite -net for X.
Let x
1
be a point in X. According to the assumption there exists x
2
X B(x
1
, ).
We argue by induction and assume that there exist x
1
, ..., x
n
points in X such that d(x
i
, x
j
)
for every i ,= j in 1, 2, ..., n.
Since x
1
, ..., x
n
is nite, it cannot be a -net. Therefore there exists a point x
n+1
in X

n
i=1
B(x
i
, ).
In this manner we construct a sequence (x
n
) such that
(8) d(x
i
, x
j
) i ,= j, i, j N.
The space X is sequentially compact, therefore (x
n
) has a convergent subsequence (x
(n)
), where
: N N is a strictly increasing function.
In particular (x
(n)
) is a Cauchy sequence. It follows that there exists N such that if m n N,
d(x
(n)
, x
(m)
) < . This contradicts (8).
We now nish the proof of the fact that a sequentially compact metric space is compact.
Indeed, let | be an open cover of X and let be the Lebesgue number for the cover | provided
by Proposition 4.35.
Proposition 4.39 implies that there exists a nite -net, x
1
, ..., x
n
.
By the denition of the Lebesgue number, each ball B(x
i
, ) is contained in some set U
i
| .
Thus X =

n
i=1
B(x
i
, )

n
i=1
U
i
.
46 PART A TOPOLOGY COURSE: HT 2013
5. Quotient spaces, quotient topology
In many cases, a space X on which an equivalent relation ! is dened is a topological space (e.g.
the group R and the equivalence relation dened by a subgroup, say Z or Q).
Natural questions to ask are:
whether the quotient space X/! is a topological space too;
which of the topological properties of X are inherited by X/! .
We shall see that a quotient space always has a natural topology, but that this topology may not
have good properties - not even the basic separation properties, even though the ambient space X is
a very good space, e.g. R .
5.1. Denition and examples. First we recall some basic notions on equivalence relations and
partitions.
Let X be a space. Recall that an equivalence relation ! on X is a relation that is
reexive: x!x for every x X;
symmetric: x!y y!x;
transitive: if x!y and y!z then x!z .
The graph of the equivalence relation ! is the set
(9) (
R
= (x, y) X X : x!y .
The equivalence class of an element x X is the set
[x] = y X : y!x .
The set of equivalence classes is denoted by X/! , and it is called the quotient space of X with
respect to !.
Example 5.1. (from Mods)
(a) Let (G, ) be a group and H a subgroup of it.
The relation x!y x
1
y H is an equivalence relation.
An equivalence class is a left coset gH, g G.
The quotient space is G/H.
(b) For instance we may consider the group (R, +) and the subgroup Z. Or the subgroup Q.
Denition 5.2. A cover U
i
: i I of a space X is called a partition if
U
i
,= for every i I ;
U
i
U
j
= for every i ,= j , i, j I .
Remark 5.3. (from Mods) Dening an equivalence relation on a space X is equivalent to dening
a partition of X.
PART A TOPOLOGY COURSE: HT 2013 47
Indeed, given an equivalence relation on a space X, the equivalence classes compose a partition
of X.
Conversely, given a partition U
i
: i I of X, one can dene an equivalence relation by
x!y i I such that x, y U
i
.
Let p : X X/! be the map that assigns to each point x in X the equivalence class [x] (the set
in the partition containing x).
Proposition 5.4. Let (X, T ) be a topological space, and ! an equivalence relation on X.
The family

T of subsets

U in X/! such that p
1
(

U) T is a topology for X/!, called the


quotient topology.
Proof. (T1) p
1
() = and p
1
(X/!) = X are both in T .
(T2) and (T3) follow from the fact that
p
1
_

U

V
_
= p
1
_

U
_
p
1
_

V
_
and that
p
1
_
_
iI

U
i
_
=
_
iI
p
1
(

U
i
) .

Proposition 5.5. A subset



V in X/! endowed with the quotient topology is closed if and only if
p
1
(

V ) is closed in X.
Proof. This follows from the fact that
p
1
_
(X/!)

V
_
= X p
1
_

V
_
.

Example 5.6. Consider the quotient space R/Q dened in Example 5.1, (b).
Let U be an open non-empty set in R/Q. Then p
1
(U) is open and non-empty in R.
For every x R, the set x +p
1
(U) is open and non-empty in R. Since Q is dense in R, there
exists q (x +p
1
(U)) Q.
So, x + q p
1
(U). The set p
1
(U) contains any equivalence class that it intersects, therefore
x +Q p
1
(U) p(x) U .
We have thus proved that for every x R, p(x) U . It follows that U = R/Q.
Thus the quotient topology on R/Q is the indiscrete topology.
48 PART A TOPOLOGY COURSE: HT 2013
5.2. Separation axioms.
Proposition 5.7. A quotient topological space satises the rst separation axiom if and only if every
equivalence class is closed.
Proof. The rst separation axiom is equivalent to the statement that every singleton [x] in X/!
is closed.
According to Proposition 5.5, the latter is equivalent to the fact that the inverse image of [x] by
p, which is [x] seen as a subset of X, is closed.
In order to discuss the Hausdor property we need to dene a particular type of sets.
Lemma 5.8. Let ! be an equivalence relation on a space X, let p : X X/! be the map x [x]
and let A be a subset in X. The following are equivalent:
(1) if A intersects an equivalence class [x] (i.e. A [x] ,= ) then A contains [x] ;
(2) whenever x A and y!x the element y is also in A ;
(3) A = p
1
(p(A)).
The proof of the equivalence is easy and left as an exercise.
Denition 5.9. A subset A of X is saturated with respect to the equivalence relation ! if it satises
one of the equivalent conditions in Lemma 5.8.
Remark 5.10. If A is a saturated subset of X with respect to ! then its complement X A is also
saturated with respect to !.
Proof. Let x X A and let y!x . If y A then x A, contradiction.
Therefore y X A, hence X A is saturated.
Example 5.11. For every subset B in X/!, p
1
(B) is saturated.
Example 5.12. (a) A map f : X Y denes the equivalence relation x!
f
y f(x) = f(y).
The equivalence classes are f
1
(y) , y f(X).
For instance, given the map f : R C, f(t) = e
2it
= cos(2t)+i sin(2t) the equivalence
relation !
f
is the same as that in Example 5.1, (b), dened by the subgroup Z of R.
(b) We may say that a set A is saturated with respect to f or f -saturated if it is saturated with
respect to !
f
.
This is equivalent to the fact that A = f
1
(f(A)).
Proposition 5.13. A quotient topological space is Hausdor if and only if any two distinct equivalence
classes are contained in two disjoint open saturated sets.
Proof. The proof is Exercise 4, Sheet 4.
Examples 5.14. (1) Let S =
_
1
n
: n N
_
. Consider the partition of R composed of S and all
the singletons x with x R S, and the corresponding equivalence relation !.
The quotient space X/! does not satisfy the rst separation axiom.
PART A TOPOLOGY COURSE: HT 2013 49
(2) Let X = [0, 1] [0, 1] , and let be the partition composed of
(0, y), (1, 1 y) with y [0, 1], (x, y) with x (0, 1) .
The quotient X/ is called the Mobius band; it is Hausdor.
Indeed, consider two distinct equivalent classes (0, y), (1, 1 y) and (a, b) with a
(0, 1).
For > 0 and > 0 small enough, the open saturated sets
[B((0, y), ) B((1, 1 y), )] X and B((a, b), ) X
are disjoint.
Similar arguments work for any two distinct equivalent classes in X/.
Proposition 5.15. (1) If X/! is Hausdor then the graph of ! is closed.
(2) If the graph of ! is closed and for every U open in X, p(U) is open then X/! is Hausdor.
Proof. (1) Let (x, y) XX be a point outside the graph (
R
. It follows that [x] ,= [y] , hence there
exist

U,

V open disjoint subsets of X/! such that [x]



U and [y]

V . Then p
1
(

U) and p
1
(

V )
are open and contain x, respectively y.
Assume that p
1
(

U) p
1
(

V ) intersects (
R
. Then there exist a p
1
(

U) and b p
1
(

V ) such
that a!b. It follows that [a] = [b] is in

U

V , contradiction.
We have thus found (x, y) p
1
(

U) p
1
(

V ) (X X) (
R
.
(2) Consider two distinct points [x] ,= [y] in X/!.
The point (x, y) is in (XX) (
R
which is open, therefore there exist U, V open in X such that
(10) (x, y) U V (X X) (
R
.
Then [x] p(U) and [y] p(V ), p(U), p(V ) open.
If p(U) p(V ) contains some element [z] then there exists u U and v V such that u!z and
v!z . By transitivity u!v, therefore (u, v) (U V ) (
R
. This contradicts the inclusion in (10).
Thus p(U) p(V ) = . We have proved that X/! is Hausdor.
Remark 5.16. The converse of Proposition 5.15, (2), is not true.
Indeed, consider the Mobius band X/, a Hausdor quotient.
The graph of the equivalence relation corresponding to the partition is closed, as it is composed
of the diagonal
= (v, v) X X : v X
and of the two graphs
((0, x), (1, 1 x)) : x [0, 1] and ((1, x), (0, 1 x)) : x [0, 1] .
But it is not true that for every U open in X, p(U) is open in X/ (equivalently, p
1
(p(U)) is
open in X).
Indeed for U = B((0, 1/2), 1/4) X, p
1
(p(U)) is
p
1
(p(U)) = [B((0, 1/2), 1/4) X] [1 (1/4, 3/4)] .
The set p
1
(p(U)) is not open, because it does not contain an open ball centred in (1, 1/2).
50 PART A TOPOLOGY COURSE: HT 2013
Examples 5.17. R/Z is Hausdor.
Indeed, the graph of the equivalence relation is the union of parallel lines
_
nZ
(x, y) : y = x +n ,
and it is closed.
For any open set U in R, p
1
(p(U)) =

nZ
(n+U) which is open as union of open sets. It follows
that p(U) is open in R/Z.
Proposition 5.15, (2), implies that R/Z is Hausdor.
A similar argument shows that R
n
/Z
n
is Hausdor.
Denition 5.18. The real n-dimensional projective space PR
n
is the quotient space of R
n+1

0 with respect to the equivalence relation ! where x!y if and only if there exists ,= 0 such that
x = y.
In other words PR
n
is the set of lines in R
n+1
through 0.
Example 5.19. For any n 1, PR
n
is Hausdor (Ex. 6, Sheet 4).
5.3. Quotient spaces and continuous maps.
Proposition 5.20 (a second way of dening the quotient topology). Let p: X X/! be x [x] .
Then quotient topology on X/! is the largest topology on X/! making p continuous.
Proof. Let

T

be a topology on X/! such that p : X


_
X/!,

T

_
is continuous. Then for every
U

T

, p
1
(U) is in T .
It follows that U

T . We have thus proved that

T



T .
Proposition 5.21. A map g from X/! to another space Z is continuous if and only if g p is
continuous.
Proof. If g is continuous then g p is continuous.
Conversely, assume that g p is continuous. For every open subset V of Z, (g p)
1
(V ) =
p
1
_
g
1
(V )
_
is open in X.
According to the denition of the quotient topology, this implies that g
1
(V )

T .
We conclude that g is continuous.
We now dene the notion of a quotient map, which is useful when trying to prove that a
quotient space is homeomorphic to another topological space.
Denition 5.22. A map p: X Y between topological spaces is a quotient map if
(1) p is surjective, and
(2) for every U Y , U is open if and only if p
1
(U) is open.
The following gives an alternative way of viewing quotient maps.
Proposition 5.23. Let p: X Y be a map between topological spaces. The following conditions are
equivalent:
(1) p is a quotient map;
PART A TOPOLOGY COURSE: HT 2013 51
(2) p is continuous and surjective, and the image under p of every p-saturated open set in X is
open in Y .
Proof. The implication (1) (2) follows from the fact that a p-saturated set A satises p
1
(p(A)) =
A.
Indeed if A is an open saturated set then p
1
(p(A)) is also an open set, therefore by (1), p(A) is
an open set.
We prove the implication (2) (1).
Assume that p
1
(U) is open. It is also p-saturated therefore according to (2) p
_
p
1
(U)
_
= U is
open. In the last equality we used the fact that p is surjective.
There is an important relationship between quotient maps and quotient spaces, as explained by the
following proposition.
Proposition 5.24. (1) Let ! be an equivalence relation on a space X, endow X/! with the
quotient topology, and let p : X X/! be the map sending each point of X to its equivalence
class. Then p is a quotient map.
(2) Suppose that p : X Y is a quotient map and that ! is the equivalence relation on X
corresponding to the partition p
1
(y) : y Y . Then X/! and Y are homeomorphic.
Proof. (1) The map p is a quotient map due to the way in which the quotient topology is dened.
(2) Consider the map p : X/! Y , p
_
p
1
(y)
_
= y. It is surjective because p is surjective, and
it is injective by construction.
If : X X/! is the standard quotient map, then by the denition of p we have that p = p.
In particular, as p is continuous, p is continuous according to Proposition 5.21.
Let U be an open set in X/!. Then
1
(U) is an open p-saturated set in X. According to
Proposition 5.23, (2), p(
1
(U)) is open.
The relation p = p implies that p(
1
(U)) = p(U), therefore p(U) is open.
We have thus proved that p is a continuous bijection, with continuous inverse, therefore a homeo-
morphism.
Examples 5.25. (1) R/Z is homeomorphic to the planar unit circle S
1
.
Consider the map f : R S
1
, f(t) = e
2it
= cos(2t) +i sin(2t) .
It denes the same equivalence relation as the subgroup Z.
The map f is surjective and continuous.
We now show that, for every open set U in R, f(U) is open in S
1
. This will imply, in
particular, that f is a quotient map.
For every x U , there exists > 0 such that (x , x + ) U . We may pick <
1
2
.
Hence the image f(x , x +) is the intersection of an open half-plane with the unit circle
S
1
and therefore f(x , x + ) is open. So, for every point in f(U), there is an open set
containing that point, lying within f(U). Hence, f(U) is open.
Proposition 5.24, (2), implies that R/Z is homeomorphic to S
1
.
52 PART A TOPOLOGY COURSE: HT 2013
(2) R
n
/Z
n
is homeomorphic to S
1
S
1
. .
n times
. This space is denoted by T
n
and it is called the
n-dimensional torus. Note that it is a compact space.
To prove the homeomorphism it suces to consider the quotient map
F(t
1
, .., t
n
) =
_
e
2it1
, ..., e
2itn
_
and argue as in (1).
In many cases, there is an easier way to establish that a space is a quotient space, using the following
proposition.
Proposition 5.26. Let f : X Y be a surjective continuous map between topological spaces. Let !
be the equivalence relation on X dened by the partitition f
1
(y) : y Y . If X is compact and Y
is Hausdor, then X/! and Y are homeomorphic.
Proof. Dene f : X/! Y by [x] f(x). This is well-dened by the denition of !. Now, f p = f
where p: X X/! is x [x] . So, Proposition 5.21, f is continuous. It is injective by the denition
of !, and it is surjective because f is. Since X is compact, so is X/!, by Proposition 4.19. So,
f : X/! Y is a continuous bijection from a compact space to a Hausdor space. By Theorem 4.25,
f is a homeomorphism.
Examples 5.27. (1) Let X = [0, 1] [0, 1] , and let be the partition composed of:
(x, 0), (x, 1), with x arbitrary number in the interval (0, 1),
(0, y), (1, y), with y arbitrary in the interval (0, 1),
(0, 0), (1, 0), (0, 1), (1, 1)
and (x, y) with x, y two arbitrary numbers in the interval (0, 1).
The quotient space [0, 1] [0, 1]/ is homeomorphic to the 2-dimensional torus.
Let f : [0, 1] [0, 1] T
2
be (t
1
, t
2
)
_
e
2it1
, e
2it2
_
. The partition f
1
(y) : y T
2
is
precisely . Hence, by Proposition 5.26, [0, 1] [0, 1]/ is homeomorphic to T
2
.
(2) Consider the closed planar unit disk
D
2
= (x, y) R
2
: x
2
+y
2
1 .
Let be the partition composed of:
all the singletons in the open disk (x, y) with x
2
+y
2
< 1
and the boundary circle S
1
.
The quotient space D
2
/ is homeomorphic to the 2-dimensional sphere
S
2
= (x, y, z) : x
2
+y
2
+z
2
= 1 .
This is proved by applying Proposition 5.26 to the map f : D
2
S
2
, f(0, 0) = (0, 0, 1),
and for (x, y) ,= (0, 0)
f(x, y) =
_
x
_
x
2
+y
2
sin
_

_
x
2
+y
2
_
,
y
_
x
2
+y
2
sin
_

_
x
2
+y
2
_
, cos
_

_
x
2
+y
2
_
_
.
Proposition 5.28. There is a (topological) embedding of the 2-dimensional torus T
2
in R
3
, i.e. T
2
is homeomorphic to a subset of R
3
with the subspace topology.
PART A TOPOLOGY COURSE: HT 2013 53
Proof. Consider the map p : R
2
R
3
,
p(x, y) = ((2 + cos(2y)) cos(2x) , (2 + cos(2y)) sin(2x) , sin(2y)) .
One can easily check that p(x, y) = p(x

, y

) if and only if (x, y) (x

, y

) Z
2
.
Indeed (2 + cos ) cos = (2 + cos

) cos

and (2 + cos ) sin = (2 + cos

) sin

imply that
the sums of the squares are equal therefore (2 + cos )
2
= (2 + cos

)
2
.
The latter is equivalent to [2 + cos [ = [2 + cos

[ , but since 2 + cos 0 for every we obtain


that cos = cos

.
Therefore p(x, y) = p(x

, y

) implies that cos(2y) = cos(2y

) and sin(2y) = sin(2y

), whence
y y

Z. Also cos(2x) = cos(2x

) and sin(2x) = sin(2x

), whence x x

Z.
The map p therefore induces a bijective map p : R
2
/Z
2
p(R
2
) dened by p
_
(x, y) +Z
2
_
=
p(x, y). If we denote by : R
2
R
2
/Z
2
the standard quotient map then p = p. Since p is
continuous it follows by Proposition 5.21 that p is continuous.
The quotient R
2
/Z
2
is compact and p(R
2
) is a subspace of R
3
, therefore Hausdor. Theorem 4.25
implies that p is a homeomorphism.

Denition 5.29. The Klein bottle K is the quotient space of the rectangle [0, 2] [0, ] by the
equivalence relation which identies the points (x, 0), (2 x, ) for each x [0, 2] and the points
(0, y), (2, y) for each y [0, ].
Note that, since [0, 2] [0, ] is compact, the Klein bottle is compact.
Proposition 5.30. There is an embedding of the Klein bottle K in R
4
.
Proof. This embedding is dened by the map g : [0, 2] [0, ] R
4
,
g(x, y) = ((2 + cos x) cos 2y , (2 + cos x) sin 2y , sin xcos y , sin xsin y) .
Indeed g is continuous.
Assume that g(x, y) = g(x

, y

).
Then the same calculation as in the proof of Proposition 5.28 implies that cos x = cos x

.
It follows that cos 2y = cos 2y

and sin 2y = sin 2y

whence y = y

modulo .
If y = y

(0, ) then sin x = sin x

hence either x = x

or x = 0, x

= 2 or x = 2, x

= 0.
If y = 0 and y

= then sin xcos y = sin x

cos y

implies that sin x = sin x

. Together with
cos x = cos x

this implies x

= 2 x.
Thus we proved that the equivalence relation on [0, 2] [0, ] dened by g coincides to the one
with quotient space the Klein bottle.
Consider the continuous bijection g : K g(K) R
4
dened by g.
The Klein bottle K is compact and g(K) is a subspace of R
4
, therefore it is Hausdor. Theorem
4.25 implies that g is a homeomorphism.
For the next proposition we use the following notation:
(a) The unit sphere in R
3
is S
2
= (x, y, z) R
3
: x
2
+y
2
+z
2
= 1;
(b) The upper hemisphere in S
2
is D
+
= (x, y, z) S
2
: z 0;
(c) The closed unit disc in R
2
is D
2
= (x, y) R
2
: x
2
+y
2
1.
54 PART A TOPOLOGY COURSE: HT 2013
Proposition 5.31. The following are all homeomorphic to the projective plane PR
2
.
(a) The quotient space S
2
/! where ! identies each pair of antipodal points of S.
(b) The quotient space D
+
/! where ! identies each pair of antipodal points on the boundary
of D
+
.
(c) The quotient space D
2
/! where ! identies each pair of antipodal points on the boundary of
D.
(d) The quotient space of the square [0, 1] [0, 1] by the equivalence relation which identies
(s, 0), (1 s, 1) for each s [0, 1] and (0, t), (1, 1 t) for each t [0, 1].
Remarks 5.32. (1) Proposition 5.31, (a), implies that the projective plane is compact.
(2) A proposition similar to Proposition 5.31 can be formulated for every projective space PR
n
,
n 2.
Proof. To prove (a) and (b) it suces to restrict the quotient map p : R
3
PR
2
to S
2
, respectively
D
+
, and apply Proposition 5.24, (2).
(c) The stereographic projection : R
2
S
2
,
(x) =
_
2x
1 +x
2
+y
2
,
2y
1 +x
2
+y
2
,
1 x
2
y
2
1 +x
2
+y
2
_
,
restricted to the unit disk D
2
denes a homeomorphism between D
2
and the upper hemisphere.
We apply Proposition 5.24, (2) to p[
D
+ [
D
2 .
(d) follows from the fact that the unit disk D
2
and the unit square [0, 1] [0, 1] are homeomorphic.

Proposition 5.33. There is an embedding of the real projective plane PR


2
in R
4
, i.e. PR
2
is
homeomorphic to a subset of R
4
.
Proof. See Exercise 7, Sheet 4.
APPENDIX A: USEFUL IDENTITIES
Let X and Y be sets, U
i

iI
a set of subsets of X and V
j

jJ
a set of subsets of Y .
(1) De Morgan laws
X

iI
U
i
=
_
iI
(X U
i
)
X
_
iI
U
i
=

iI
(X U
i
)
(2) Distributivity of

over

_
_
iI
U
i
_
=
_
iI
_
A

U
i
_
PART A TOPOLOGY COURSE: HT 2013 55
(3) Images and inverse images
Let f : X Y be a map. Recall that for any subset A in Y its inverse image or pre-image
is
f
1
(A) = x X ; f(x) A .
Then
f(U) V if and only if U f
1
(V )
f
1
(Y V ) = X f
1
(V )
f
_
_
iI
U
i
_
=
_
iI
f(U
i
)
f
_

iI
U
i
_

iI
f(U
i
)
f
1
_
_
_
jJ
V
j
_
_
=
_
jJ
f
1
(V
j
)
f
1
_
_

jJ
V
j
_
_
=

jJ
f
1
(V
j
)
If A B = , A, B subsets of Y , then f
1
(A) f
1
(B) = .
Indeed assume that there exists x f
1
(A) f
1
(B). Then f(x) A and f(x) B,
which contradicts A B = .

Potrebbero piacerti anche