Elementary Topology - Full
Elementary Topology - Full
Elementary Topology - Full
ALEX GONZALEZ
1. Introduction 3
2. Metric spaces 4
2.1. Continuous functions 8
2.2. Limits 10
2.3. Open subsets and closed subsets 12
2.4. Continuity, convergence, and open/closed subsets 15
3. Topological spaces 19
3.1. Interior, closure and boundary 21
3.2. Continuous functions 23
3.3. Basis of a topology 25
3.4. The subspace topology 26
3.5. The product topology 28
3.6. The quotient topology 31
3.7. Other constructions 33
4. Connectedness 34
4.1. Path-connected spaces 38
4.2. Connected components 40
5. Compactness 41
5.1. Compact subspaces of Rn 44
6. Countability and separation axioms 47
6.1. The countability axioms 47
6.2. The separation axioms 49
REFERENCES 54
Contents
1
2 ALEX GONZALEZ
ELEMENTARY TOPOLOGY I 3
1. Introduction
When we consider properties of a “reasonable” function, probably the first thing that
comes to mind is that it exhibits continuity: the behavior of the function at a certain
point is similar to the behavior of the function in a small neighborhood of the point.
What’s more, the composition of two continuous functions is also continuous.
Usually, when we think of a continuous functions, the first examples that come to
mind are maps f : R → R:
• the identity function, f (x) = x for all x ∈ R;
• a constant function f (x) = k ;
• polynomial functions, for instance f (x) = xn , for some n ∈ N;
• the exponential function g(x) = ex ;
• trigonometric functions, for instance h(x) = cos(x).
The set of real numbers R is a natural choice of domain to begin to study more
general properties of continuous functions. After all, we are familiar with many of
the properties of the real line, and it is (relatively) easy to draw the graphs of functions
R → R.
So, let’s review the definition of continuity for a function f : R → R: the function f is
continuous at the point a ∈ R if
lim f (x) = f (a)
x→a
(in particular we are assuming that this limit exists!). Another way of putting the
above definition is the following: for each ε > 0 there exists some δ > 0 such that
|f (x) − f (a)| < ε for all x ∈ R such that |x − a| < δ.
The function f is then said to be continuous on all R if it is continuous for all a ∈ R.
So, basically, the definition of continuity depends only on the absolute value, which
is essentially a rule to measure the distance between any two numbers in R. It would
seem that continuity relies on the existence of a rule to measure distances, or more
precisely, a metric.
4 ALEX GONZALEZ
2. Metric spaces
Let X be a set. Roughly speaking, a metric on the set X is just a rule to measure the
distance between any two elements of X .
Definition 2.1. A metric on the set X is a function d : X × X → [0, ∞) such that the
following conditions are satisfied for all x, y, z ∈ X :
A note of waning! The same set can be given different ways of measuring distances.
Strange as it may seem, the set R2 (the plane) is one of these sets. We will see
different metrics for R2 pretty soon.
Example 2.2. The set X = R with d(x, y) = |x−y|, the absolute value of the difference
x − y , for each x, y ∈ R. Properties (M1) and (M2) are obvious, and
d(x, y) = |x − y| = |(x − z) + (z − y)| ≤ |x − z| + |z − y| = d(x, z) + d(z, y).
x
d(x, y)
d(x, z) y
d(z, y)
are also metrics on R2 (the details will be checked in Examples 2.5 and 2.8 below).
We have just see three different metrics on R2 . But why stopping at R2 ? We can
extend these metrics to Rn for all n ≥ 1.
for each (x1 , x2 , . . . , xn ), (y1 , y2 , . . . , yn ) ∈ Rn . As usual (M1) and (M2) are easy to
check, but (M3) is not trivial at all.
Notice that both sides of the inequality are positive. Thus, by squaring the above, it
is equivalent to prove that
v v
n
X n
X n
X
u n
X
u n
u uX
(ri + si )2 ≤ ri2 + i
2
s +2 t r 2t
s2 . i i (2)
i=1 i=1 i=1 i=1 i=1
Replacing this in (2) and simplifying, we deduce that (1) holds if and only if
n
X 2 n
X n
X
ri si ≤ ri2 s2i .
i=1 i=1 i=1
Pn Pn
Proof. Consider the expression i=1 j=1 (ai bj − aj bi )2 . By expanding the brackets,
n X
X n n
X n
X n
X n
X n
X n
X
2 2 2 2 2
(ai bj − aj bi ) = ai bj + aj bi − 2 ai b i aj b j .
i=1 j=1 i=1 j=1 j=1 i=1 i=1 j=1
Since the left part of the equality is positive, this proves the statement.
Example 2.8. Finally, the box metric on Rn is defined by
d (x1 , . . . , xn ), (y1 , . . . , yn ) = max{|xi − yi | i = 1, . . . , n}
Properties (M1) and (M2) are easily seen to hold. Let’s check property (M3). Let x =
(x1 , . . . , xn ), y = (y1 , . . . , yn ) and z = (z1 , . . . , zn ) ∈ Rn . Then, for each i = 1, . . . , n,
|xi − yi | = |(xi − zi ) + (zi − yi )| ≤ |xi − zi | + |zi − yi | ≤ d(x, z) + d(z, y).
Thus, the triangle inequality holds. As an exercise, consider the set R2 with this
metric. Fix then a point a ∈ R2 and draw the set
Da = {x ∈ R2 d(a, x) ≤ 1}.
ELEMENTARY TOPOLOGY I 7
All these examples should serve as a warning of how flexible the notion of metric is:
we should not be surprised to see that statements that go against all intuition are,
in fact, true. The following is a good examples of this: it is a process to define a
metric on every set. This procedure is not rather descriptive, but it is nonetheless
important.
Example 2.9. Let X be any set. The discrete metric on X is defined by
1, if x 6= y
d(x, y) =
0, if x = y
Properties (M1) and (M2) are obvious, so let’s check property (M3). Let x, y, z ∈ X .
If x = y, then d(x, y) = 0, and there is nothing to check. Suppose then that x 6= y.
Then, either z = x 6= y, or z = y 6= x, or z 6= x, y. Regardless the situation, we have
then
1 = d(x, y) and 1 ≤ d(x, z) + d(z, y) ≤ 2,
and the triangle inequality holds.
One could say that this metric is rather bad: if we think of a metric as a rule to
measure the distance between points x, y in the set X , then in this case all points
are far (if x 6= y) or really close (if x = y), but we cannot distinguish anything beyond
this. However, this metric is not useless: it becomes a nice tool to check if geometric
intuition really works and provides counterexamples.
Example 2.10. Consider the set of integer numbers Z, and let p be a prime number.
The p-adic valuation of n ∈ Z is the maximum non-negative integer νp (n) such that
pνp (n) divides n.
We can now define the p-adic metric on Z by
0, if a = b;
d(a, b) = 1
, otherwise.
pνp (a−b)
Property (M1) holds by definition of d, and property (M2) holds by definition of the
p-adic valuation: νp (a − b) = νp (b − a). Let’s prove the triangle inequality. Let then
a, b, c ∈ Z. We can assume that these are all different numbers. In this case, property
(M3) looks like
1 1 1
≤ + ,
pνp (a−b) pνp (a−c) pνp (c−b)
and this is a consequence of the following property of the p-adic valuation:
• νp (a − b) ≥ min{νp (a − c), νp (c − b)}.
Let’s prove the above property. Let then k = νp (a − b), m = νp (a − c), n = νp (c − b)}
and l = min{νp (a − c), νp (c − b). By definition of the p-adic valuation, there exist
x, y, z ∈ Z such that p does not divide any of them, and such that
pk x = a − b = a − c + c − b = pm y + pn z = pl (pm−l y + pn−l z),
and now it is clear that k ≥ l.
8 ALEX GONZALEZ
Fix then some (a, b) ∈ R2 . To prove that µ is continuous at (a, b) we have to check
that for each ε > 0 there exists some δ > 0 such that
if d2 ((a, b), (u, v)) < δ then d1 (µ(a, b), µ(u, v)) = d1 (ab, uv) < ε.
Let’s ignore ε for a second, and suppose that
p
d2 ((a, b), (u, v)) = (a − u)2 + (b − v)2 < δ
for some δ > 0. Then, we have |a − u|, |b − v| < δ , and
d1 (ab, uv) = |ab − uv| = |b(a − u) + a(b − v) + (a − u)(v − b)| ≤
≤ |b| · |a − u| + |a| · |b − v| + |a − u| · |b − v| < δ 2 + δ(|a| + |b|).
Let’s go back now to the question of the continuity of µ at (a, b) ∈ R2 . Let then ε > 0,
and consider the quadratic equation on δ :
δ 2 + (|a| + |b|)δ − ε = 0.
√
−(|a|+|b|)+ (|a|+|b|)2 +4ε
This equation has always a positive solutions, namely δ = 2
, and
the above discussion implies that, for this particular choice of δ , if d2 ((a, b), (u, v)) < δ
then d1 (ab, uv) < ε as required.
Exercise 2.15. Let (X, d) be a metric space and let z ∈ X . We can then define a map
δz
X / [0, ∞)
x / δz (x) = d(z, x)
Prove that the map δz is continuous with respect to the metric d on X and the
Euclidean metric on [0, ∞) ⊆ R.
Example 2.16. Let d be the p-adic metric on the set of integers Z, and fix some
α ∈ Z. Then, the maps
f g
(Z, d) / (Z, d) (Z, d) / (Z, d)
k / α·k k / k+α
are continuous.
1
Since pνp (α)
≤ 1, it follows that, for each a ∈ Z and each ε > 0,
if d(a, b) < ε then d(f (a), f (b)) ≤ d(a, b) < ε
and hence f is a continuous map. In particular notice that if p does not divide α
then f is an isometry.
Let’s check now that the map g is continuous. In this case, for each a, b ∈ Z we have
1 1
d(f (a), f (b)) = d(a + α, b + α) = = = d(a, b),
pνp (a+b−(b+α)) pνp (a−b)
so in fact this map is always an isometry (in particular, it is continuous).
2.2. Limits. Convergence of sequences is another feature that depends on the exis-
tence of a metric.
Definition 2.18. Let (X, d) be a metric space and let {xn }n∈N be a sequence of
elements of X . Then, the the sequence {xn }n∈N converges to the element x ∈ X if, for
every ε > 0 there exists some N ∈ N such that
d(x, xn ) < ε for all n ≥ N.
When this is the case, the element x is called the limit of the sequence {xn }n∈N , and
we write x = limn→∞ xn .
ELEMENTARY TOPOLOGY I 11
Theorem 2.19. Let (X, d) be a metric space. If the sequence {xn }n∈N in X has a limit,
then it is unique.
Proof. Suppose the sequence {xn }n∈N converges to the points x, y ∈ X , and suppose
x 6= y . This means that, for each ε > 0 there exist Nx , Ny ∈ N such that
d(x, xn ) < ε, for all n ≥ Nx and d(y, xn ) < ε, for all n ≥ Ny .
d(x,y)
In particular, let ε = 2
> 0, and let N = max{Nx , Ny }, where Nx , Ny are as
above (with respect to this particular choice of ε). Then, for each n ≥ N , we have
d(x, xn ), d(y, xn ) < ε. On the other hand, by the triangle inequality,
d(x, y)
d(x, y) ≤ d(x, xn ) + d(xn , y) < 2 · ε = 2 · = d(x, y),
2
which is impossible. Thus, x = y .
Example 2.20. Let X be a set with the discrete metric. Then, a sequence {xn }n∈N is
convergent if and only if there exists some N ∈ N such that xn = xN for all n ≥ N .
Studying the convergence of sequences in a set with the discrete metric is not really
interesting.
Example 2.21. Let X be the set of all continuous functions f : [0, 1] → R, and
consider the following two metrics on X :
Z 1
dint (f, g) = |f (x) − g(x)|dx d∞ (f, g) = max{|f (x) − g(x)| x ∈ [0, 1]}.
0
Now fix some α > 0 and let {fn }n∈N be the sequence where fn is the function with
graph
(0, α)
(0, 0) (1, 0)
( 12 − n1 , 0) ( 12 , 0) ( 12 + n1 , 0)
Then, the sequence {fn }n∈N converges to the function g = 0 (constant on [0, 1] with
image 0) with respect to the metric dint . Indeed, for each n
Z 1 Z 1
1 1
dint (0, fn ) = |0 − fn (x)|dx = fn (x) dx = and lim = 0.
0 0 n n→∞ n
However, the same sequence {fn } does not converge to the constant function g = 0
with respect to the metric d∞ , since for any n
d∞ (0, fn ) = max{|0 − fn (x)| x ∈ [0, 1]} = α > 0.
12 ALEX GONZALEZ
In fact, this sequence does not converge to any continuous function with respect to
the metric d∞ , but we are not ready yet to prove this.
2.3. Open subsets and closed subsets. The idea of open/closed subset again relies
in the existence of a metric.
Definition 2.22. Let (X, d) be a metric space. For each element a ∈ X and each
ε ∈ (0, ∞) ⊆ R, the open ball with center a and radius ε is the subset
Ba,ε = {x ∈ X d(a, x) < ε} ⊆ X.
A subset U ⊆ X is open with respect to the metric d if for each y ∈ U there is some
ε(y) ∈ (0, ∞) such that
By,ε(y) ⊆ U.
A subset U ⊆ X is closed with respect to the metric d if the set X \ U is open with
respect to the metric d.
The notation ε(y) above is there to stress the idea that the radius of the ball By,ε
depends on the element y . Note also that we have to specify with respect to which
metric the subset U is open in X . The following example illustrates the reason.
Example 2.23. Let dEuc and ddis be, respectively, the Euclidean and the discrete
metrics on R2 , and let a ∈ R2 be any point. Then, U = {a} ⊆ R2 is open with respect
to ddis but not open with respect to dEuc .
Lemma 2.24. Let (X, d) be a metric space. Then for each x ∈ X and for each ε > 0,
(1) Bx,ε = {y ∈ X d(x, y) < ε} is an open subset of X with respect to d; and
(2) Dx,ε = {y ∈ X d(x, y) ≤ ε} is a closed subset of X with respect to d.
Proof. To prove part (1), we have to show that, for each y ∈ Bx,ε there exists some
δ > 0 such that By,δ ⊆ Bx,ε . The following sketch already hints which is the right
choice of δ .
ELEMENTARY TOPOLOGY I 13
To prove (2), notice that X \ Dx,ε = {y ∈ X d(x, y) > ε}, and we have to prove that
this subset is open. Let then y ∈ X \ Dx,ε , as in the following picture.
The following is the most important and basic result about open subsets in a metric
space.
Theorem 2.25. Let then (X, d) be a metric space, and let O be the collection of all open
subsets of X with respect to d. Then,
14 ALEX GONZALEZ
(1) the total subset, X , and the empty set, ∅, are elements of O;
(2) if {Uα }α∈Γ ⊆ O is any family (possibly infinite) of open subsets, then
[
Uα ∈ O;
α∈Γ
Proof. Property (1) is easy: since X contains all possible open balls, it is open. Also,
the empty set ∅ contains no points at all, and thus it also satisfies the condition to
be an open subset.
Let now U = {Uα }α∈Γ ⊆ O be a family (possibly infinite) of open subsets of X , and
let V = ∪Γ Uα ⊆ X . For each u ∈ V , there is at least one Uα ∈ U which contains u.
Furthermore, since Uα is open, there exists some ε > 0 such that
d
Bu,ε ⊆ Uα ⊆ V.
Hence V is open with respect to d.
Proposition 2.26. Let (X, d) be a metric space. Then, the following holds:
(2) if {U1 , . . . , Un } ⊆ O is any finite family of closed subsets, then the union
Sn
i=1 Ui
is also a closed subset;
Example 2.27. Let X be a set with the discrete metric. Then, for each x ∈ X the
one-element subset {x} ⊆ X is both open and closed. To check that it is open just
ELEMENTARY TOPOLOGY I 15
notice that {x} = Bx,1 (open ball of radius 1). To check that it is closed, notice that
the complement [ [
X \ {x} = {y} = By,1
y∈X\{x} y∈X\{x}
is a (possibly infinite) union of open subsets, and hence is open.
Example 2.28. Consider R2 with the Euclidean metric. Then, for each u ∈ R2 , the
subset {u} ⊆ R2 is closed but not open. Indeed, it is clear that {u} is not open. To
check that it is closed, consider the complement U = R2 \ {u}. We can describe U as
U = {v ∈ R2 d(u, v) > 0}.
d(u,v)
Fix some v ∈ U and set ε = 2
> 0. Then, for each w ∈ Bv,ε we have
d(u, v)
d(u, w) ≥ |d(u, v) − d(v, w)| = d(u, v) − d(v, w) > > 0,
2
and thus w ∈ U (in the above list of (in)equalities we are using some result from the
first exercise list).
Exercise 2.29. Let (X, d) be a metric space and let x ∈ X be any element. Prove
that the subset {x} ⊆ X is closed with respect to d (in other words, regardless of the
metric, a point is always a closed subset).
Proof. Suppose first that (2) is true. We prove by contradiction that V c is open.
Assume then that V c is not open, and let u ∈ V c be an element for which Bu,ε 6⊆ V c ,
for any ε > 0. This means, in particular, that for any n ∈ N the ball Bu, 1 contains
n
(at least) an element xn of V . We can form then the sequence {xn }n∈N ⊆ V , but this
sequence clearly converges to the element u ∈ / V , hence the contradiction.
Suppose now that (1) is true. We have to show that every convergent sequence of
elements of V has its limit in V . Let then {xn }n∈N ⊆ V be a convergent sequence,
and suppose v = lim xn ∈ V c . Since V c is open by hypothesis, there exists some
ε > 0 such that Bv,ε ⊆ V c . On the other hand, since v = lim xn , there exists some
ε0 < ε and some N ∈ N such that d(v, xn ) < ε0 for all n ≥ N , and this implies that
xn ∈ Bv,ε0 ⊆ Bv,ε ⊆ V c ,
which contradicts the assumption that xn ∈ V for all n.
16 ALEX GONZALEZ
Let’s analyze the definition of continuity. Let then f : (X, dX ) → (Y, dY ) be a contin-
uous function, and let a ∈ X . The map f is continuous at a if, for each ε > 0 there
is some δ > 0 such that
if dX (a, x) < δ then dY (f (a), f (x)) < ε.
The left part above says that x ∈ Ba,δ , while the right part says that f (x) ∈ Bf (a),ε .
This suggests some relation between continuity and open subsets, which we formu-
late properly as the following result. You can take it as a set of equivalent definitions
of the notion of continuity.
Theorem 2.31. Let (X, dX ) and (Y, dY ) be metric spaces, and let f : X → Y be a map.
Then, the following statements are equivalent:
(4) for any sequence {xn }n∈N in X , if {xn } converges to z ∈ X , then the sequence
{f (xn )}n∈N in Y converges to f (z).
We are not assuming the existence of any inverse function! For instance consider
the function f : R → R defined by f (x) = x2 . Clearly, this function does not have an
inverse, but the pre-image of the subset {1} ∈ R makes perfect sense:
Remark 2.32. With this set of equivalent definitions of continuity, we do not need to
define continuity at a point, we can define continuous functions directly: a function
between metric spaces f : X → Y is continuous if and only if for each open subset
U ⊆ Y the pre-image f −1 (U ) is an open subset of X .
Be careful with the following: condition (2) above concerns only the pre-image of
open subsets of Y , but it says nothing about the image of open subsets of X . The
following is, in general, false:
if U ⊆ X is an open subset, then f (U ) ⊆ Y is an open subset.
Proof of Theorem 2.31. Since this proof is rather long, let’s split it in smaller parts,
each proving an equivalence of statements.
First we prove that (1) implies (2). Suppose then that f is a continuous map, and
let U ⊆ Y . Let also a ∈ f −1 (U ), and let u = f (a). Since U is open, there exists
some ε(u) > 0 such that Bu,ε(u) ⊆ U . Also, since f is continuous, there exists some
δ(a) > 0 such that
if dX (a, x) < δ(a) then dY (f (a), f (x)) = dY (u, f (x)) < ε(u),
which implies that f (Ba,δ(a) ) ⊆ U , and hence Ba,δ(a) ⊆ f −1 (U ). This proves that
f −1 (U ) is open with respect to dX .
Let’s see now that (2) implies (1). Suppose then that f −1 (U ) is an open subset of X
for each open subset U ⊆ Y , and fix a ∈ X . For each ε > 0, the open ball Bf (a),ε is
an open subset of Y , and thus f −1 (Bf (a),ε ) is an open subset of X . Furthermore,
a ∈ f −1 (Bf (a),ε ),
and this implies that there exists some δ > 0 such that
Ba,δ ⊆ f −1 (Bf (a),ε ).
The above inclusion is equivalent to say that f is continuous at a ∈ X .
• Statements (2) and (3) are equivalent.
Let V ⊆ Y be a closed subset, and let U = V c , which is an open subset by hypothesis.
By (2), f −1 (U ) is an open subset of X . Furthermore, we have
f −1 (U ) = f −1 (V c ) = f −1 (Y \ V ) = X \ f −1 (V ),
and thus f −1 (V ) is a closed subset, and this proves that (2) implies (3). Replacing
“open” by “closed” proves the converse.
• Statements (1) and (4) are equivalent.
First we prove that (1) implies (4). Let then {xn }n∈N be a sequence in X which
converges to z ∈ X . Since f is continuous (at z ), for each ε > 0 there exists some
δ > 0 such that
if dX (z, x) < δ then dY (f (z), f (x)) < ε.
Also, since the sequence is convergent, we know that there exists some N ∈ N such
that dX (z, xn ) < δ for all n ≥ N . Thus, for all n ≥ N , we have
dY (f (z), f (xn )) < ε,
which implies (4).
Finally, we have to check that (4) implies (1). Suppose otherwise that (1) is not true
on a fixed element a ∈ X : there exists some ε > 0 such that for each δ > 0 there is
some x ∈ X such that
dX (a, x) < δ but dY (f (a), f (x)) ≥ ε.
By fixing such ε, we can then define the sets
1
An = {x ∈ X dX (a, x) < and dY (f (a), f (x)) ≥ ε},
n
18 ALEX GONZALEZ
which is non-empty for all n ∈ N. Next choose for each n an element xn ∈ An : the
sequence {xn }n∈N clearly converges to a ∈ X , while the sequence {f (xn )}n∈N does
not converge to f (a), contradicting (4).
The above result implies a crucial consequence: it is not the metric(s) what matters,
but the collection(s) of open/closed subsets!
ELEMENTARY TOPOLOGY I 19
3. Topological spaces
We finished the previous chapter realizing that, when we are talking about continuity
of functions, metrics are not so relevant as open subsets. We thus shift the focus to
collections of open subsets instead of metrics.
(T1) the total set, X , and the empty set, ∅, are elements of T ;
Example 3.2. Let X be any set, and let T1 be the collection of all the subsets of X ,
and let T2 = {∅, X}. Then, both T1 and T2 are topologies on X . More precisely,
The topology T1 is induced by the discrete metric, whereas the T2 is as far from the
discrete topology as it can be, hence the name.
Example 3.3. Let (X, d) be a metric space, and let T be the collection of open subsets
of X with respect to the metric d. Then, T is a topology on X by Theorem 2.25, and
(X, T) is a topological space. In other words, every metric on X defines a topology
on X .
Notice that different metrics on the same set X could give rise to the same topology.
For example, let X = Rn . The Manhattan metric (Example 2.5), the Euclidean metric
(Example 2.6) and the box metric (Example 2.8) all define the same topology on X .
Another example of this situation is the following: all metrics on a given finite set
define the same topology, the discrete metric.
Example 3.4. Let X = {a, b, c} and let T = {∅, {a}, {a, b}, X}. Then, T is a topology
on X which is not induced by any metric.
20 ALEX GONZALEZ
a b c
Let’s see another example of all the unexpected situations that can happen when
dealing with topological spaces.
Example 3.6. This example is usually called the line with two origins. Let X = R∪{z}
(here z denotes simply an extra element), and let
(1) T1 = {U ⊆ R U is open with the Euclidean metric};
3.1. Interior, closure and boundary. This section reviews some basic concepts re-
lated to open/closed subgroups.
(1) the interior of A is the union of all the open subsets which are contained in A:
[
A◦ = U;
U ∈T, U ⊆A
(2) the closure of A is the intersection of all the closed subsets which contain A:
\
A= V.
V c ∈T, A⊆V
Proof. We prove part (i) and leave part (ii) as an exercise. Let W = ∪U ∈T, U ⊆A U . Since
T is a topology, W ∈ T and W ⊆ A. This means that W ⊆ A◦ . On the other hand,
A◦ ∈ T and A◦ ⊆ A. By definition, A◦ ⊆ W .
Lemma 3.10. Let (X, T) be a topological space. Then the following holds.
(3) If A◦ ⊆ B ⊆ A then A◦ = B ◦ .
22 ALEX GONZALEZ
Tn
(5) Let {U1 , . . . , Un } be a finite collection of subsets of X and let V = i=1 Ui . Then,
n
\
◦
V = (Ui )◦ .
i=1
Proof. Properties (1) and (2) are obvious by definition. Property (3) is proved as
follows. If A◦ ⊆ B , then by definition A◦ ⊆ B ◦ ⊆ B ⊆ A. Also, since B ◦ is an open
subset of A, it follows that B ◦ ⊆ A◦ . Property (4) follows from the fact that, for each
γ ∈ Γ, we have ∩Γ Uα ⊆ Uγ ⊆ ∪Γ Uα . Finally, let’s prove property (5). The intersection
I = ∩ni=1 (Ui )◦ is an open subset of ∩ni=1 Ui , and hence I ⊆ V . This combined with
property (4) finishes the proof.
A similar list of properties exists for the closure of a subset. We state it below without
proof (the proof is similar to the result above, so it is left as an exercise).
Lemma 3.11. Let (X, T) be a topological space. Then the following holds.
Sn
(5) Let {U1 , . . . , Un } be a finite collection of subsets of X and let V = i=1 Ui . Then,
n
[
V = Ui .
i=1
Finally, let’s see some interesting properties of the interaction of interior, closure and
boundary.
Lemma 3.12. Let (X, T) be a topological space. Then, for any subsets A, B ⊆ X ,
(1) the closure of the complement of A is the complement of the interior of A:
c
Ac = X \ A = X \ (A◦ ) = A◦ ;
ELEMENTARY TOPOLOGY I 23
(4) ∂ B ◦ ∩ B ◦ = ∅.
c
Proof. To prove (1), notice that A◦ = A \ ∂(A) = A ∩ ∂(A) . Thus,
c c c
A◦ = X \ (A◦ ) = X \ A ∩ ∂(A) = (X \ A) ∪ X \ ∂(A) =
c c
= Ac ∪ ∂(A) = Ac ∪ ∂(A) = Ac ∪ ∂(Ac ) = Ac
To prove (2) just replace A by B c everywhere above. To check property (3), let B = A.
Then
∂ B = b ∩ X \ B = B ∩ X \ B ⊆ B.
To prove
(4),◦ notice
c that Y ∩ Y c = ∅ for any Y ⊆ X . Thus, we just need to check that
◦
∂ B ⊆ B . Indeed,
∂ B ◦ = ∂ (B ◦ )c = ∂ B c ⊆ B c = (B ◦ )c ,
where (from left to right) the first equality holds by definition, the second equality
holds by (2), the inequality holds by (3) and the last equality holds by (1).
Exercise 3.13. Let (X, T) be a topological space. Prove that, for any subset A ⊆ X ,
def
∂(A) = A \ A◦ = {x ∈ A x ∈
/ A◦ }.
Exercise 3.16. Let X = {1, 2, 3, 4} and let Y = {1, 2, . . . , 8}. Consider also the
collections
TX = {∅, {1}, {4}, {1, 3}, {2, 4}, {1, 2, 4}, {1, 3, 4}, X}
TY = {∅, {1}, {1, 2}, {1, 2, 3}, . . . , X}
Finally, let f : X → Y be the map defined by
2n, if n is even
f (n) =
n + 1, if n is odd
Prove that TX and TY are topologies for X and Y respectively, and show that f is a
continuous function.
Theorem 3.17. Let (X, TX ) and (Y, TY ) be topological spaces, and let f : X → Y be
a map. Then the following are equivalent.
Proof. We prove that (1) implies (2), (2) implies (3), and (3) implies (1). Afterwards we
show that (4) is equivalent to (1).
Assume first that (1) holds, and let A ⊆ X be a subset. For each x ∈ A we have to
check that f (x) ∈ f (A). Fix such x and let V ∈ TY be any open subset containing
f (x). Since f is continuous, the subset U = f −1 (V ) is an open subset that contains
the element x. Note that U ∩A 6= ∅, hence there exists y ∈ A∩U , and f (y) ∈ V ∩f (A).
Since every open subset containing f (x) intersects f (A) nontrivially, this means that
f (A) ⊆ f (A).
Assume now that (2) holds. Let B ⊆ Y be a closed subset, and let A = f −1 (B). We
have to show that A = A. We have f (A) ⊆ B , and thus if x ∈ A then
Suppose next that (3) holds, and let V ⊆ Y be an open subset. Then W = V c = Y \ V
is closed, and
f −1 (W ) = f −1 (Y \ V ) = X \ f −1 (V )
By hypothesis f −1 (W ) is closed, and thus f −1 (V ) is open.
Next we prove that (1) implies (4). Le x ∈ X and let V ∈ TY containing f (x). Then the
set U = f −1 (V ) is an open subset containing x. Conversely, assume that (4) holds.
Let V ∈ TY and let x ∈ f −1 (V ). Then f (x) ∈ V , and by hypothesis there is some
ELEMENTARY TOPOLOGY I 25
3.3. Basis of a topology. Very often we will not be able to specify all the open subsets
of a given topology, but rather we will specify a subcollection that generates the whole
topology.
Example 3.19. Consider R with the topology induced by the Euclidean metric, and
let
B1 = {(x0 , x1 ) ⊆ R x0 , x1 ∈ Q} B2 {(x0 , x1 ) ⊆ R x0 , x1 ∈ R \ Q}
Both collections above are bases for the same topology, their intersection is empty,
and while B1 has countably many elements, B2 does not!
Lemma 3.20. Let (X, T) be a topological space, and let B be a subset of T . Then, B
is a basis for T if and only if it satisfies the following properties.
(B1) For each U ∈ T and each x ∈ U there exists some V ∈ B such that x ∈ V ⊆ U .
(B2) Let V1 , V2 ∈ B and let x ∈ V1 ∩ V2 . Then there exists V ∈ B such that x ∈ V ⊆
V1 ∩ V2 .
Proof. Suppose first that B is a basis for T . In order to check condition (B1), let U ∈ T
and x ∈ U . By definition of basis, there exists some collection {Vγ }γ∈Γ ⊆ B such that
[
U= Vγ .
γ∈Γ
We can use Lemma 3.20 to formalize this idea of a topology generated by a basis.
26 ALEX GONZALEZ
Given a basis B for a topology on X , we can define the following collection of subsets:
TB = {U ⊆ X for each x ∈ U there exists V ∈ B such that x ∈ V ⊆ U }.
We call TB the topology generated by B. Again, keep in mind that the same topology
may be generated by more than one basis.
3.4. The subspace topology. Let (X, T) be a topological space, and let A ⊆ X be a
subset. We can then define
TA = {U ∩ A U ∈ T}.
Proof. (T1) ∅ = ∅ ∩ A, A = X ∩ A ∈ TA .
(T2) Let {Uγ }γ∈Γ ⊆ TA . By definition of TA , for each γ ∈ Γ there
S exists some Wγ ∈ T
such that Uγ = A ∩ Wγ . Since T is a topology, we have W = γ∈Γ Wγ ∈ T , and hence
[ [ [
Wγ ∩ A = W ∩ A ∈ TA .
U= Uγ = (Wγ ∩ A) =
γ∈Γ γ∈Γ γ∈Γ
(T3) Let {U1 , . . . , Un } ⊆ TA . Again, for each i there exists some Wi ∈ T such that
Ui = Wi ∩ A. Since T is a topology, we have
n
\ n
\ n
\
Wi ∩ A ∈ T.
U= Ui = (Wi ∩ A) =
i=1 i=1 i=1
Lemma 3.24. Let (X, T) be a topological space and let A ⊆ X , with the subspace
topology. Then, V ⊆ A is open if and only if there is some U ∈ T such that V = U ∩ A.
Theorem 3.26. Let (X, T) be a topological space, let A ⊆ X be a subspace, and let
incl : A → X be the inclusion map. The subspace topology TA on A is the only topology
satisfying the following property.
(S) A map f : (Z, TZ ) → (A, TA ) is continuous if and only if the composition map
incl ◦f : (Z, TZ ) → (X, T) is continuous.
Proof. Suppose first that f is a continuous map. The inclusion map incl : A → X
is continuous by Lemma 3.25, and it follows that incl ◦f is a continuous map by
Proposition 3.15. Conversely, suppose that incl ◦f is continuous. We have to check
that f is continuous: for each V ∈ TA , f −1 (V ) ∈ TZ . Let U ∈ TX be such that
V = U ∩ A. Then,
(incl ◦f )−1 (U ) = f −1 (U ∩ A) = f −1 (V )
is open, and f is continuous.
Theorem 3.27. Let (X, TX ) and (Y, TY ) be topological spaces,
S and suppose that there
exist open (closed) subsets A, B ⊆ X such that X = A B . Consider A and B as
topological subspaces of X . Then, the following holds.
28 ALEX GONZALEZ
Proof. Part (1) is an immediate consequence of Proposition 3.15 and Lemma 3.25:
the map f (respectively g ) is the composition of the inclusion map incl : A → X
(respectively incl : B → X ) and h.
To prove (2), let V ∈ TY . Then,
h−1 (V ) = f −1 (V ) ∪ g −1 (V ).
Since f and g are continuous maps, f −1 (V ) ∈ TA and g −1 ∈ TB . Since A and B are
open subsets of X , it follows that f −1 (V ) and g −1 (V ) are also open subsets of X , and
thus h−1 (V ) ∈ TX .
3.5. The product topology. Let now (X, TX ) and (Y, TY ) be topological spaces, let
X × Y be the product set, and consider the collection of subsets
B = {U × V U ∈ TX and V ∈ TY }.
Proof. We have to check the conditions in Definition 3.21. Condition (B1’) states
that for each (x, y) ∈ X × Y there must exist some W ∈ B such that (x, y) ∈ W .
Let U ∈ TX and V ∈ TY be such that x ∈ U and y ∈ V . Then W = U × V ∈ B
satisfies condition (B1’). Condition (B2) states that, given U1 × V1 , U2 × V2 ∈ B and
(x, y) ∈ (U1 × V1 ) ∩ (U2 × V2 ), there exists some U × V ∈ B such that
(x, y) ∈ U × V ⊆ (U1 × V1 ) ∩ (U2 × V2 ).
Choose U = U1 ∩ U2 ∈ TX , V = V1 ∩ V2 . Then U × V clearly satisfies the condition.
Let TX×Y be the topology generated by the above basis BX×Y , and let πX : X ×Y → X
and πY : X × Y → Y be the obvious projection functions. The product topology is
determined by the following universal property.
Theorem 3.29. The collection TX×Y is the only topology on X × Y satisfying the
following condition.
(P) A map f : (Z, TZ ) → (X × Y, TX×Y ) is continuous if and only if the compositions
πX ◦ f : (Z, TZ ) → (X, TX ) and πY ◦ f : (Z, TZ ) → (Y, TY ) are continuous.
Proof. The collection B is contained in TX×Y , and clearly satisfies the conditions (B1)
and (B2) in Lemma 3.20, and thus TX×Y is the topology generated by B. We have to
check then that it satisfies property (P), and that it is unique satisfying this condition.
The rest of the proof is divided into several smaller parts.
ELEMENTARY TOPOLOGY I 29
We have to check that, for each W ∈ TX×Y , f −1 (W ) ∈ TZ . Let’s assume first that
W = U × V , for some U ∈ TX and V ∈ TY . Then we have
f −1 (W ) = f −1 (U × V ) = f −1 ((U × Y ) ∩ (X × V )) = f −1 (U × Y ) ∩ f −1 (X × V ) =
−1
= f −1 (πX (U )) ∩ f −1 (πY−1 (V )) = (πX ◦ f )−1 (U ) ∩ (πY ◦ f )−1 (V ) ∈ TZ .
This means that the preimage of any element of the basis BX×Y is in TZ . Now, each
W ∈ TX×Y is a union of elements of BX×Y , and this shows that f is continuous.
4. The topology TX×Y is unique satisfying property (P).
Let T be another topology on X × Y satisfying property (P), and consider the identity
map Id : (X × Y, T) → (X × Y, T). This is clearly continuous, and hence πX = πX ◦ Id
and πY = πY ◦ Id are continuous too. This means that, for each U ∈ TX and V ∈ TY ,
we have
−1
πX (U ) = U × Y ∈ T πY−1 (V ) = X × V ∈ T
and hence U × V = (U × Y ) ∩ (X × V ) ∈ T . In other words, T contains the basis
BX×Y , and thus TX×Y ⊆ T .
On the other hand, we know from part 1. that the maps
πX : (X × Y, TX×Y ) → (X, TX ) πY : (X × Y, TX×Y ) → (Y, TY )
are continuous. Notice that the compositions
πX
πX : (X × Y, TX×Y )
Id / (X × Y, T) / (X, TX )
πY
πY : (X × Y, TX×Y )
Id / (X × Y, T) / (Y, TY )
are continuous maps, and thus by property (P) the map Id : (X × Y, TX×Y ) → (X ×
Y, T) is continuous, and this means that T ⊆ TX×Y .
WARNING!. The definition of the product topology TX×Y on X × Y does NOT say that
all the open subsets in TX×Y are of the form U × V for U ∈ TX and V ∈ TY !
30 ALEX GONZALEZ
For instance, let TR be the Euclidean topology (induced by the Euclidean metric) on
R, and let TR×R be the product topology on R × R. The subsets (0, 2), (1, 3) ⊆ R are
open in this topology, and we may consider the subset
The previous results can be stated actually for infinite cartesian products.
S
Definition 3.31. Let {Xα }α∈Γ be a collection of sets, and let X = α∈Γ Xα . The
cartesian product of {Xα }α∈Γ is
Y
Xα = {set functions ω : Γ → X such that ω(α) ∈ Xα ∀α ∈ Γ}.
α∈Γ
Q
The α-th coordinate of a function ω ∈ α∈Γ Xα is the element ω(α) ∈ Xα . The α-th
projection map is the function
Y
πα : Xα → Xα
α∈Γ
ExampleQ3.33. There are other, “natural” topologies that we can consider on the
productQ α∈Γ Xα , but only one satisfies the universal property above. The box topol-
ogy on α∈Γ Xα is the topology generated by the basis
Y
B={ Uα | Uα ∈ Tα }.
α∈Γ
If the index set Γ is finite (i.e., if we are only considering a finite cartesian product)
then the box product is the same as the product topology, but this is not the case
when the index set Γ is infinite. Indeed, an open subset in the product topology has
the form Y
Uα
α∈Γ
where Uα = Xα for all but finitely many values of α. It is clear then that the box
topology has more open subsets than the product topology.
3.6. The quotient topology. Both the subspace topology and the product topol-
ogy are natural generalizations of constructions available for metric spaces. In this
section we study a new construction which is not available for metric spaces. The
idea behind the quotient topology is to create new spaces by “pasting” together other
spaces.
For example, we can construct a cylinder out of a stripe (a subset of R2 ), by gluing
together two non-adjacent sides, as described in the following sketch.
y y
In this case, it is easy to provide the cylinder with a metric (just consider the cylinder
as a subset of R3 with the Euclidean metric), but we can use this example to analyze
a more general construction.
As a set, the cylinder is the quotient set of the stripe by the following equivalence
relation:
(x, y) ∼ (x, y) if x 6= 0, 1 (0, y) ∼ (1, y).
The question is then which topology do we put on this quotient set. Notice that the
quotient set of the stripe by this equivalence relation is not a subset of R3 (although
it can be identified with a subset of R3 ). As happened with the product of topological
spaces, not any topology is acceptable.
32 ALEX GONZALEZ
Let (X, T) be a topological space, and let ∼ be an equivalence relation on the set X .
You may think of the elements of X in the same equivalence class as the elements
that we are going to glue together, just as we did with the stripe above. Let then X
be the set of equivalence classes, and let
π : X −→ X
be the map that sends each x ∈ X to its equivalence class. Notice that the map π is
surjective (as a map of sets). Define also
T = {U ⊆ X π −1 (U ) ∈ T}.
For obvious reasons, the topological space (X, T) is called the quotient space of (X, T)
by the equivalence relation ∼, and the map π : X → X is called the quotient map.
Theorem 3.35. Let (X, T) be a topological space, and let ∼ be an equivalence relation
on X . Let also (X, T) be the quotient space by ∼. Then, T is the unique topology on X
satisfying the following conditions.
(Q1) The map π : (X, T) → (X, T) is continuous.
Proof. First we have to check that the topology T satisfies conditions (Q1) and (Q2).
The map π : X → X is continuous since, by definition, π −1 (U ) ∈ T for all U ∈ T .
Let f : (X, T) → (Y, TY ). If f is continuous, then clearly f ◦ π is continuous because
it is a composition of continuous functions. Suppose that f ◦ π : (X, T) → (Y, TY )
is continuous. We have to show that π −1 (V ) ∈ T for all V ∈ TY . Since f ◦ π is
continuous, we have (f ◦ π)−1 (V ) = f −1 (π −1 (V )) ∈ T , and hence by definition of T it
follows that π −1 (V ) ∈ T .
ELEMENTARY TOPOLOGY I 33
0
Now, suppose that T is another topology on X which also satisfies properties (Q1)
0
and (Q2) with respect to the map π 0 : (X, T) → (X, T). We have to prove that T = T .
0 0
Consider first the identity map Id : (X, T) → (X, T ). Since T satisfies condition (Q1),
the map π 0 = Id ◦ π is continuous and thus Id is continuous by property (Q2). This
0 0
implies that T ⊆ T . Similarly, we can consider the map Id0 : (X, T ) → (X, T). Again,
since π = Id ◦ π 0 is continuous, it follows that Id0 is continuous, and hence T ⊆ T .
3.7. Other constructions. Combining the subspace topology, the product topology
and the quotient topology we can create many new topologies (or topological spaces).
Here are some of the most relevant constructions in topology.
Example 3.36. The mapping cylinder. Let f : (X, TX ) → (Y, TY ) be a map of topolog-
ical spaces. To construct the mapping cylinder of f , let
a
Mf = [0, 1] × X Y / ∼,
where ∼ is the equivalence relation generated by (1, x) ∼ f (x) for all x ∈ X . The
topology of Mf is a combination of the product topology and the quotient topology.
Example 3.37. The mapping cone. Let f : (X, TX ) → (Y, TY ) be a map of topological
spaces. The mapping cone is a slight variation of the mapping cylinder:
a
Cf = [0, 1] × X Y / ≡,
where ≡ is the equivalence relation generated by (1, x) ≡ f (x) and (0, x) ≡ (0, z)
for all x, z ∈ X . Again, the topology of Cf is a combination of the product and the
quotient topologies.
Example 3.38. The suspension of a topological space. The suspension of (X, T) is
SX = ([0, 1] × X)/ ∼,
where ∼ is the equivalence relation generated by (0, x) ∼ (0, z) and (1, x) ∼ (1, z)
for all x, z ∈ X . The topology on SX is again a combination of the product and the
quotient topologies.
Example 3.39. The join of two spaces. The join of (X, TX ) and (Y, TY ) is
X ∗ Y = ([0, 1] × X × Y )/ ≡,
where ≡ is the equivalence relation generated by (0, x, y) ≡ (0, x, u) and (1, x, y) ≡
(1, z, y) for all x, z ∈ X and all y, u ∈ Y .
34 ALEX GONZALEZ
4. Connectedness
Now that we understand the basics of topological spaces, we can start comparing
different topological spaces, trying to find properties that will distinguish one space
from another. This chapter deals with the first of such properties: connectedness.
Essentially, a topological space is connected if it cannot be formed out of smaller,
disjoint pieces.
Example 4.2. Let X be a set with the discrete topology. Then, X is connected if and
only if X is the empty set or a singleton (if X contains more than one element, then
each sub-singleton of X is open and closed).
Example 4.4. Consider R with the topology corresponding to the Euclidean metric,
and let Q be the subspace of rational numbers. Then, Q is not connected. Indeed,
the subset
(−∞, π) ∩ Q
is open and is also the complement of an open subset, hence closed.
The equality on the left means that A is open, while the equality on the right says
that A is the complement of an open subset, hence closed.
As we have seen already several times in this course, it is always useful to have
alternative statements for the same definition, so let’s see some alternative ways of
defining connectedness.
Theorem 4.6. Let (X, TX ) be a topological space. Then, the following are equivalent.
(2) There does not exist open subsets U, V ⊆ X such that U ∪V = X and U ∩V = ∅.
(3) There does not exist any surjective continuous map f : (X, TX ) → ({0, 1}, Tdisc ).
ELEMENTARY TOPOLOGY I 35
Proof. Let’s check first that (1) is equivalent to (3). Suppose first that there is a
surjective, continuous map f : (X, TX ) → ({0, 1}, Tdisc ), and let A = f −1 ({0}) and
B = f −1 ({1}). Since f is surjective, both A and B are non-empty subsets of X .
Furthermore, since {0} (respectively {1}) is and open an closed subset of {0, 1} and
f is continuous, it follows that A (respectively B ) is an open and closed subset of X .
Finally, since both A and B are non-empty, it follows that none of them is either the
empty set or X , and thus X is not connected.
Suppose now that X is not connected, and let U $ X be a non-empty, open and
closed subset. Let also V = U c , and notice that it is again a non-empty, open and
closed subset of X . We can then define a map f : (X, TX ) → ({0, 1}, Tdisc ) by setting
f (U ) = {0} and f (V ) = {1}, and it is clear that f is surjective and continuous.
Finally let’s prove that (2) is equivalent to (3). If there exists a map f as in (3), then
U = f −1 ({0}) and V = f −1 ({1}) are open subsets of X which clearly satisfy the
conditions of (2). Conversely, if U, V ⊆ X satisfy the conditions of (2) then we just
have to define f by f (x) = 0 if x ∈ U and f (x) = 1 if x ∈ V .
Theorem 4.7. Let f : (X, TX ) → (Y, TY ) be a continuous map of topological spaces,
and suppose X is connected. Then, f (X) ⊆ Y is a connected subspace of Y .
Proof. Exercise.
Corollary 4.8. Let h : (X, TX ) → (Y, TY ) be a homeomorphism of topological spaces.
Then, X is connected if and only if Y is connected.
Theorem 4.9. The set R with the topology associated to the Euclidean metric is con-
nected.
Proof. Suppose R = A∪B , where A, B are non-empty open subsets such that A∩B =
∅. Then, A and B are both closed (they are the complement of each other), and thus
∂(A) = A ∩ R \ A = A ∩ B = A ∩ B = ∅,
and similarly for B .
On the other hand, we will now show that if U $ R is non-empty, then ∂(U ) 6= ∅,
hence contradicting the above. This part of the proof uses the Euclidean metric on
R. Let then a ∈ R \ U . If
U ∩ (a, ∞) = ∅ = U ∩ (−∞, a),
then U = {a}, and hence U = ∂(U ) 6= ∅. Suppose then that U ∩ (a, ∞) 6= ∅ (the other
possible case is similar), and let b = inf{x ∈ U ∩ (a, ∞)}. Then we have
inf{d(b, u) u ∈ U } = 0 inf{d(b, v) v ∈ R \ U } = 0,
and this implies that b ∈ ∂(U ).
Example 4.10. For all a < b ∈ R, the open interval (a, b) ⊆ R is connected. Indeed,
(a, b) is clearly homeomorphic to ( −π , π ) (give the explicit homeomorphism!), and the
2 2
map
−π π
tan : , −→ R
2 2
is a homeomorphism.
Proof. This follows immediately from Proposition 4.11 since [a, b] is the closure of
(a, b), and the latter is connected.
Lemma 4.13. Let (X, TX ) be a topological space, and let {Aγ }γ∈Γ be a collection of
connected subsets of X such that
Aγ ∩ Aω 6= ∅
for all γ, ω ∈ Γ. Then, A = ∪Γ Aγ is a connected subset of X .
Proof. If X × Y is connected, then both X and Y are connected, since the projection
functions πX : X × Y → X and πY : X × Y → Y are continuous maps.
ELEMENTARY TOPOLOGY I 37
Suppose now that X and Y are connected. Fix then some element a ∈ X and let
Γ = Y (as sets). Consider also the sets
• B = {a} × Y ⊆ X × Y ; and
• Cy = X × {y} ⊆ X × Y , where y ∈ Y .
Note that B is homeomorphic to Y , while Cy is homeomorphic to X for all y ∈ Y .
Thus, all the above subsets are connected subspaces of X × Y .
For each y ∈ Y , we have Cy ∩ B = {(a, y)} = 6 ∅, and hence the subset Uy = B ∪ Cy ⊆
X × Y is connected (for all y ∈ Y ) by Lemma 4.13. Now, consider the collection
{Uy }y∈Y . For any y, y 0 ∈ Y , we have Uy ∩ Uy0 6= ∅, and hence
[
Uy = X × Y
y∈Y
basis for the product topology that contains a. By definition there is some N ∈ N
such that Ui = R for all i ≥ N + 1, and
x = (a1 , . . . , aN , 0, 0, . . .) ∈ U ∩ X∞ .
This shows that indeed X is the closure of X∞ . By Proposition 4.11, X is connected.
Proof. Prove that the space Gf is homeomorphic to X (exercise). Then the statement
follows.
Example 4.18. This example is known as the topologist’s sine curve. Let
Ω = {(x, sin(1/x)) ∈ R2 x > 0} ∪ {(0, y) ∈ R2 y ∈ [−1, 1]} ⊆ R2 .
The following is a sketch of Ω.
A topological space (X, TX ) is path-connected if, for any two elements x, y ∈ X there
is a path from x to y .
Theorem 4.20. Let (X, TX ) be a path-connected space. Then (X, TX ) is connected.
Proof. Let M be a Moebius band and C be a cylinder, both seen as subsets of R3 . Let
also ∂(M ) and ∂(C) be the boundaries of M and C are subsets of R3 . Clearly, ∂(C)
is not connected, while ∂(M ) is path-connected (hence connected). Thus, no map
h : C → M can be a homeomorphism.
It turns out that not all connected spaces are path-connected, and we have already
met the first example of this behavior recently.
Example 4.25. Let Ω ⊆ R2 be the subset in Example 4.18, the topologist’s sine
curve, and recall that Ω is connected. Let’s prove now that Ω is not path-connected.
Let then
A = {(0, y) ∈ R2 y ∈ [−1, 1]} Γ = {(x, sin(1/x)) ∈ R2 x > 0}.
If we choose elements a ∈ A and b ∈ Γ, then there is no path in Ω linking a to b! To
prove this, suppose otherwise and let ω : [0, 1] → Ω be a path with ω(0) = (0, 0) and
ω(1) = ( π1 , 0). Let also F : [0, 1] → R be the continuous map
ω π
1
F : [0, 1] −→ Ω −→ R,
where π1 (x, y) = x. Since F (0) = 0 and F (1) = π1 , we can use the Intermediate Value
2
Theorem to find some t1 ∈ (0, 1) such that F (t1 ) = 3π . Using again the Intermediate
2
Value Theorem we can then find some t2 ∈ (0, t1 ) such that F (t2 ) = 5π . Iterating the
process we can find a sequence {tn }n≥1 , such that
2
tn+1 ∈ (0, tn ) and F (tn ) =
(2n + 1)π
40 ALEX GONZALEZ
Since the sequence {tn } ⊆ [0, 1] is monotonic decreasing and bounded below, it is
convergent: there is some t ∈ [0, 1] such that lim tn = t. Finally, since ω is continuous
2
the sequence {ω(tn )} converges to ω(t). On the other hand, ω(tn ) = ( (2n+1)π , (−1)n ),
and this sequence is not convergent, hence a contradiction.
Alternatively, the set π0 (X) can be constructed as follows. Define a relation on the
set X : x ∼ y if there is some connected subset U ⊆ X such that x, y ∈ U . This is
clearly an equivalence relation, and then we can define π0 (X) as the set of equivalence
classes.
Proposition 4.27. Every continuous map f : (X, TX ) → (Y, TY ) induces a map of sets
f∗ : π0 (X) −→ π0 (Y ).
In other words, if X and Y are homeomorphic, they have to be made out of the same
number of pieces.
ELEMENTARY TOPOLOGY I 41
5. Compactness
A subcover of {Uγ }γ∈Γ is a subcollection {Vω }ω∈Ω ⊆ {Uγ }γ∈Γ such that A ⊆ ∪Ω Vω .
Example 5.2. Consider the family of intervals Un = (n, n + 2) ⊆ R (Euclidean topol-
ogy), for n ∈ Z. The collection {Un }n∈Z is an open cover of R. This open cover does
not contain any proper subcover: if we remove Un for some n ∈ Z then the union of
the remaining subsets Um , m 6= n, does not cover R.
Definition 5.3. Let (X, TX ) be a topological space. A subset A ⊆ X is compact if
every open cover of A contains a finite subcover.
Remark 5.4. Let (X, TX ) be a topological space and A ⊆ X . Let also {Uγ }γ∈Γ be an
open cover of A in X . A finite subcover is a finite subcollection {Vω }ω∈Ω ⊆ {Uγ }γ∈Γ
which is also an open cover of A. This is already a hint of the idea that compactness
generalizes finiteness.
WARNING!. As a note of caution, be extremely careful with the above definition.
Specially with the bit every open cover... It is not enough to check that an open cover
contains a finite subcover, we have to check ALL possible open covers! Indeed, notice
that, for any topological space (X, TX ), the collection {X} is a finite open cover of X ,
but this does not make X compact!
The best way to understand this notion is by looking at some examples. Let’s start
with example of spaces which are not compact.
Example 5.5. Consider R with the Euclidean topology. Then, the open cover in
Example 5.2 does not contain any finite subcover, and hence R is not compact.
Example 5.6. Consider R with the Euclidean topology, and let (0, 1) ⊆ R be the open
interval. Then (0, 1) is not compact. Indeed, consider the open cover
( )
1
Un = ,1
n
n∈N
It is easy to see that this open cover does not contain any finite subcover of (0, 1).
42 ALEX GONZALEZ
Example 5.7. Let X = {0} ∪ {1/n | n ∈ N}, with the subspace topology of a subset
of R. Then X is a compact. Indeed, let {Uγ }γ∈Γ be an open cover of X , and let α ∈ Γ
be such that 0 ∈ Uα . Then Uα must contain all but finitely many elements of X . For
each x ∈ X \ Uα , choose Ux ∈ {Uγ }γ∈Γ containing x. This forms a finite subcover.
Proof. Let A ⊆ X be a closed subset, and let {Uγ }γ∈Γ S be an open cover (recall that the
subsets Uγ are open in X ). Then the collection {Ac } {Uγ }γ∈Γ is an open cover of X ,
and there exists some finite subcover, {W1 , . . . , Wn }, since X is compact. But clearly
this subcover also covers A, and upon removing Ac from {W1 , . . . , Wn } if necessary
this constitutes a finite subcover of the original cover {Uγ }γ∈Γ . Thus A is compact.
Proof. Let {Uγ }γ∈Γ be an open cover of f (A) in Y . Since f is a continuous map, the
collection {f −1 (Uγ )}γ∈Γ is then an open cover of A. Now, since A is compact we know
that there is some finite subcover {f −1 (Uγ1 ), . . . , f −1 (Uγr )}, and thus {Uγ1 , . . . , Uγr }
is a finite subcover of f (A).
Theorem 5.11. Let (X, TX ) and (Y, TY ) be topological spaces, and let A ⊆ X , B ⊆ Y
be compact subsets. Then, A × B is a compact subset of X × Y .
Step 1. Let a ∈ A and let N ⊆ X × Y be an open subset such that {a} × B ⊆ N . Then
there exists some U ⊆ B such that {a} × B ⊆ U × B ⊆ N .
You can think of U × B as a tube around the ray {a} × B . To show the above point,
let {Wγ }γ∈Γ be an open cover of {a} × B , where, for each γ , the subset Wγ is an
element of the basis for the product topology such that Wγ ⊆ N . In other words, for
each γ ∈ Γ,
Wγ = Uγ × Vγ ⊆ N,
where Uγ ∈ TX and Vγ ∈ TY .
W1
W2
W3
B
W4
W5
W6
Let {Wγ }γ∈Γ be an open cover of A × B . Given a ∈ A, the ray {a} × B is compact and
may be covered by finitely many elements Wa,1 , . . . , Wa,ma of the original cover. The
union Na = Wa,1 ∪ . . . ∪ Wa,ma is an open susbet containing {a} × B , and by Step 1
we know that there exist a tube Ua × B such that
{a} × B ⊆ Ua × B ⊆ Na .
Thus, for each a ∈ A we may choose some Ua ∈ TX containing the element a such
that the tube Ua × B can be covered by finitely many elements of {Wγ }γ∈Γ . Note
that {Ua }a∈A forms an open cover of A, and thus there exist a1 , . . . , an ∈ A such that
{Ua1 , . . . , Uan } is a finite subcover, since A is compact.
Note that A × B ⊆ Ua1 × B ∪ . . . ∪ Uan × B . Thus, the collection
{Wa1 ,1 . . . , Wa1 ,ma1 , Wa2 ,1 , . . . , Wa2 ,ma2 , . . . , Wan ,1 , . . . , Wan ,man }
is a finite subcover of A × B .
44 ALEX GONZALEZ
Note that the above definition does not make much sense for topological spaces! Let
us now characterize compact subsets of R.
Theorem 5.13 (Heine-Borel Theorem for R). Consider R with the Euclidean metric,
and let A ⊆ R. Then, A is compact if and only it is closed and bounded.
Proof. Suppose first that A is compact. To see that it is bounded, consider the
following open cover of A:
{Un = B0,n n ∈ N}.
Since it is an open cover of A and A is compact, there exists some finite subcover,
namely {Un1 , . . . , Unr }. Let N = max{n1 , . . . , nr }. Clearly, Un1 , . . . , Unr ⊆ UN , and
hence A ⊆ UN = B1,N , i.e. A is bounded.
Let’s see also that A is closed. Suppose otherwise, and let x ∈ A \ A. For each n ∈ N,
let c
Un = Bx, 1 = R \ Bx, 1 .
n n
Now, since u − δ < u, the set Au−δ is covered by a finite subcover of {Uγ }, namely
{Uγ1 , . . . , Uγm }, and then {Uω , Uγ1 , . . . , Uγm } is a finite subcover of Au+δ . Again, this
implies that u + δ ∈ B , which contradicts the hypothesis that u = sup B . Hence, B
is not bounded above, and A is compact.
We can use this theorem to deduce that some other subsets of R are compact. The
following example is not intuitive at all.
Example 5.14. Let K ⊆ R be the Cantor set (see Exercise 6 in List 3): start from
I0 = [0, 1] ⊆ R, then form I1 = [0, 13 ] ∪ [ 23 , 1] by removing the middle third ( 13 , 23 ) of I0 ,
and keep iterating the process. We end up with a sequence {In }n∈N of subsets of R,
and
∞
\
K= In .
n=0
Since each In is closed and K is an (infinite) intersection of closed subsets, we know
that K is closed. Also, it is clearly bounded, since K ⊆ I0 , and hence by the Heine-
Borel it is compact.
Proposition 5.15. Let (X, dX ) be a metric space, and let A ⊆ X be a compact sub-
space. Then, A is closed and bounded.
Proof. First we check that A is bounded. Fix some x ∈ X and let Un = Bx,n for each
n ∈ N. The collection {Un }n∈N is thus an open cover of A in X , and since A is compact
there exists some finite subcover, {Un1 , . . . , Unr }. Let then N = max{n1 , . . . , nr }: we
have A ⊆ UN = Bx,N and hence A is bounded.
46 ALEX GONZALEZ
To see that A is also closed, suppose otherwise and let x ∈ A) \ A. For each n ∈ N
define then c
Un = Bx, 1 = X \ Bx, 1 .
n n
Then, Un is open for all n and we have
[ [ \
Un = X \ Bx, 1 = X \ Bx, 1 = X \ {x} ⊆ A.
n n
n∈N n∈N n∈N
Thus, {Un }n∈N is an open cover of A, and since A is compact there is some finite
subcover, {Un1 , . . . , Uns }. Furthermore, by setting N = max{n1 , . . . , ns }, we have
A ⊆ UN , which implies
A ∩ Bx, 1 = ∅,
N
We can use now the Heine-Borel Theorem to prove that a certain subset of Rn is
compact.
Example 5.17. Consider R2 with the Euclidean metric. Then, every closed disk and
every circle is compact. More generally, in the Euclidean space Rn , the sphere S n−1 ,
S n−1 = {(x1 , . . . , xn ) ∈ Rn x21 + . . . + x2n = 1}
is a compact space.
Example 5.18. The cylinder and the Moebius band are both compact subsets of
R3 . Thus compactness is not a good property to distinguish these two subsets.
Fortunately connectivity has done the job for us already.
Example 5.19. Consider the set Z with the discrete metric. Then, Z is closed and
bounded, but not compact. Indeed, Z is closed as a subset of itself, and it is bounded
because Z is contained in the open ball of radius 2 around any element. However,
every singleton {n} in Z is an open subset, and hence the collection {n} n∈Z is an
open cover which does not contain any finite subcover.
ELEMENTARY TOPOLOGY I 47
In this chapter we analyze certain properties of different nature: countability (of bases
for a given topology) and separation (of different subsets in a given topology). The
ultimate goal is to give conditions for a given topological space to be metrizable (i.e.,
the given topology is actually the topology associated to some metric).
6.1. The countability axioms. Recall the definition of basis of a topology, see Sec-
tion 3.3.
Definition 6.1. Let (X, TX ) be a topological space. X has a countable basis at x if
there is a countable collection B of open subsets satisfying the following:
(i) x ∈ U for each U ∈ B; and
(ii) if V ∈ TX is such that x ∈ V , then there exists some U ∈ B such that U ⊆ V .
If X has a countable basis at every element x ∈ X , then X satisfies the first count-
ability axiom. We also say that X is first-countable.
Theorem 6.2. Let (X, dX ) be a metric space, and let TX be the topology associated to
dX . Then (X, TX ) is first-countable.
Proof. For each x ∈ X , the collection B = {Bx , 1/n} satisfies conditions (i) and (ii) in
the definition above.
Theorem 6.3. Let (X, TX ) be a first-countable space. Then the following holds.
(i) Let A ⊆ X and x ∈ X . Then, x ∈ A if and only if there exists a sequence {xn }n∈N
of points of A converging to x.
(ii) Let f : (X, TX ) → (Y, TY ). Then f is continuous if and only if, for every convergent
sequence {xn }n∈N in X with limit x, the sequence {f (xn )}n∈N converges to f (x).
Proof. Part (i) is proved by modifying appropriately the proof of Theorem 2.30. The
necessary modifications in the proof of Theorem 2.31 imply part (ii).
Definition 6.4. A topological space (X, TX ) satisfies the second countability axiom if
it has a countable basis for TX . We also say that X is second-countable.
Remark 6.5. If (X, TX ) is second-countable then it is first-countable, but the con-
verse is not true in general.
Lemma 6.6. Let (X, TX ) be a second-countable space. Then, every discrete subspace
A ⊆ X is countable.
The collection Ω is countable, and also covers A. Indeed, for each a ∈ A there is
some Uα ∈ {Uγ }γ∈Γ such that a ∈ Uα , and thus Uα contains some element of BA by
definition of basis.
Part (ii). From each nonempty basis element Wn ∈ BA , choose some xn ∈ Wn , and
let D = {xn | Wn ∈ BA , Wn 6= ∅}. Then D is dense in A: for each a ∈ A, every basis
element in BA that contains a must intersect D , so x ∈ D .
6.2. The separation axioms. Given a metric space, we have already seen (in the
exercise lists) that every two elements can be separated from each other by two dis-
joint subsets. We have also seen examples of topological spaces where this property
does not hold any more (for instance X = {a, b} with the indiscrete topology). In
this section we analyze different levels of separability in topological spaces and their
consequences.
Lemma 6.12. Let (X, TX ) be a Hausdorff space. Then every singleton {x} ⊆ X is a
closed subset.
Theorem 6.14. Let (X, dX ) be a metric space, and let TX be the topology associated
to dX . Then (X, TX ) is normal.
Proof. Clearly, a metric space is Hausdorff, and thus singletons are closed subsets.
Let A, B ⊆ X be closed and disjoint subsets. Note that the points in A are not limit
points of B , and reciprocally the points in B are not limit points of A. Thus,
50 ALEX GONZALEZ
(1) for each a ∈ A there is some ε(a) such that Ba,ε(a) ∩ B = ∅; and
Proof. Suppose that {xn }n∈N converges to x, and let y 6= x be a different element of
X . By definition there exist U, V ∈ TX such that x ∈ U , y ∈ V and U ∩ V = ∅. Since
the sequence converges to x, the open subset U contains all but finitely many terms
of the sequence, which implies that V contains at most finitely many terms of the
sequence. Thus {xn }n∈N cannot converge to y .
Example 6.16. Let X = R. Let also K = {1/n | n ∈ N}, and let
B = {(a, b) | a, b ∈ R} ∪ {(a, b) \ K | a, b ∈ R}.
This forms a basis for a topology on X , namely TX . The space (X, TX ) is easily seen
to be Hausdorff. However, it is not regular, as we next show: the set K is a closed
subset of X , but it cannot be separated from {0}.
Example 6.17. Let X = R, and let B be the collection of intervals of the form [a, b),
with a < b. Then B forms a basis for a topology on X , namely TB . In this example
we show that (X, TB ) is a normal space, while the product space X × X is not.
Next we show that X × X is not normal (with the product topology). Notice that
X × X is regular, since X is regular. Now, suppose that X × X is normal, and let
L = {(x, −x) | x ∈ X}. Then the following holds (it is left as an exercise to prove these
properties actually hold):
Let now A ⊆ L be any closed subset. Since L itself is closed, it follows that A is also
closed in X × X . Actually, both A and L \ A are closed in X × X , and there are
disjoint open subsets UA and VA separating A and L \ A.
Here, UA is the subset of X × X fixed in the previous paragraph. We claim that the
map θ is injective. Let A ⊆ L be a proper subset. Then θ(A) = D ∩ UA 6= ∅, D, since
D ∩ UA , D ∩ VA 6= ∅. Now, if A, B ⊆ L are proper, distinct subsets, then one of them
contains an element that is not contained in the other. For simplicity suppose that
(x, −x) ∈ A, (x, −x) ∈/ B . Then (x, −x) ∈ L \ B , and thus
(x, −x) ∈ UA ∩ VB .
Notice that UA ∩ VB is open and nonempty, and thus K = D ∩ (UA ∩ VB ) 6= ∅, since D
is dense. By definition, the elements of K are elements of D ∩ UA , but not of D ∩ UB ,
and thus θ(A) 6= θ(B).
Lemma 6.18. Let (X, TX ) be a topological space whose singletons are closed subsets.
Then the following holds.
Proof. We only prove part (i), part (ii) being an easy exercise. Suppose X is regular,
and let x ∈ X , U ∈ TX be given. Let B = U c , which is a closed subset not containing
x. Since X is regular, there exist disjoint subsets V, W ∈ TX such that x ∈ V and
B ⊆ W . We claim that V ∩ B = ∅: if y ∈ B , then W ∈ TX is such that y ∈ W and
W ∩ V = ∅. Thus, V ⊆ U as desired.
To prove the converse, let x ∈ X and let B ⊆ X be a closed subset not containing x.
We have to show that there are disjoint open subsets separating x and B . Let U =
B c ∈ TX . Then x ∈ U , and there exists some V ∈ TX be such that x ∈ V ⊆ V ⊆ U .
c
Let W = V . Then x ∈ V , B ⊆ W and V ∩ W = ∅. Thus X is regular.
Theorem 6.19. The following holds.
(i) A subset of a Hausdorff space is Hausdorff.
(ii) A product of Hausdorff spaces is Hausdorff.
(iii) A subset of a regular space is regular.
(iv) A product of regular spaces is regular.
Proof. Point (i). Let (X, TX ) be a Hausdorff topological space, let A ⊆ X , and let
x, y ∈ A. Since X is Hausdorff, there exist U, V ∈ TX such that x ∈ U , y ∈ V and
U ∩ B = ∅. Then U ∩ A, V ∩ A ∈ TA , x ∈ U ∩ A, y ∈ V ∩ A and (U ∩ A) ∩ (V ∩ A) = ∅.
Point (ii). Let {Xγ }γ∈Γ be a family of Hausdorff spaces, let X = Πγ∈Γ Xγ be the product
space, and let x = (xγ ), y = (yγ ) be two distinct elements of X . Since x 6= y , there
exists some γ ∈ Γ such that xγ 6= yγ . Choose W, Z ∈ TXγ to be disjoint subsets
separating xγ and yγ , and let U = πγ−1 (W ) and V = πγ−1 (Z), where πγ : X → Xγ is
the projection map. Then U and V are disjoint open subsets of X that separate x
and y .
Point (iii). Let (X, TX ) be a regular space, and let A ⊆ X . By (i), A is Hausdorff, and
thus singletons are closed subsets. Let also x ∈ A be a point and B ⊆ A be a closed
subset (B is closed in A), x ∈ / B . Let C ⊆ X be the closure of B in X , so B = C ∩ A.
Then x ∈ / C , and the regularity of X implies that there exist U, V ∈ TX such that
x ∈ U , C ⊆ V and U ∩ V = ∅. Thus U ∩ A and V ∩ A are disjoint open subsets of A
separating x and B .
Point (iv). Let {Xγ }γ∈Γ be a family of regular spaces, and let X = Πγ∈Γ Xγ be the
product space. By (ii), X is Hausdorff, and thus singletons are closed subsets. Let
x = (xγ ) ∈ X be a point, and let U ⊆ X be an open subset containing x. Choose a
basis open subset W = Πγ Uγ such that x ∈ W ⊆ U . Since each Xγ is regular, we
may choose for each γ an open subset Vγ such that x ∈ Vγ ⊆ Vγ ⊆ Uγ . Furthermore,
whenever Uγ = Xγ , we choose Vγ = Xγ . This way, V = Πγ Vγ satisfies
x ∈ V ⊆ V = Πγ Vγ ⊆ Πγ Uγ ⊆ U.
By Lemma 6.18 (i) it follows that X is regular.
ELEMENTARY TOPOLOGY I 53
Proof. Let B be a countable basis for TX , and let A, B ⊆ X be disjoint closed subsets.
Then for each point a ∈ A there is an open subset Ua ∈ TX such that Ua ∩ B = ∅.
Since X is regular, we may choose Va ∈ TX such that a ∈ Va ⊆ Va S ⊆ Ua , and thus
we may also choose Wa ∈ B such that a ∈ Wa ⊆ Va . This way, W = a∈A Wa is such
that A ⊆ W and W ∩ B = ∅. Furthermore, the collection {Wa }a∈A is countable (after
removing any possible repetitions), so we may write it as {Wn }n∈N . A similar process
S
produces a collection {Zn }n∈N such that B ⊆ Z = n∈N Zn and Z ∩ A = ∅. However,
W and Z may not be disjoint.
For each n ∈ N, define
n
[ n
[
Wn0 = Wn \ ( Zi ) and Zn0 = Zn \ ( Wi ).
i=1 i=1
Wn0 0 0 0
Z 0 = n Zn0 . Then,
S S
Notice
S that and ZS
are open for all n. Set W =
n n Wn and
0 0 0 0
A ⊆ n Wn and B ⊆ n Zn , and, we claim that W ∩ Z = ∅. Suppose otherwise and
let x ∈ W 0 ∩ Z 0 . Then x ∈ Wj0 ∩ Zk0 for some j, k ∈ N. Suppose for simplicity that
j ≤ z . By definition of Wj0 it follows that x ∈ Wj . On the other hand, by definition of
Zk0 we have x ∈/ Wj , which is a contradiction.
Theorem 6.21. Let (X, TX ) be a compact Hausdorff space. Then X is normal.
REFERENCES
The main published references for this course are the following.
1. M. O. Searcoid, Metric Spaces, Springer Undergraduate Mathematics Series,
ISBN 978-1-84628-369-7.
2. W. A. Sutherland, Introduction to Metric and Topological Spaces, 2nd Edition,
Oxford University Press, ISBN 978-0-19-956308-1.
Other general, electronic sources are the following.
1. http://www-history.mcs.st-and.ac.uk/~john/MT4522/index.html
2. http://en.wikibooks.org/wiki/Topology/Metric_Spaces
References about the p-adic metric.
1. http://scientopia.org/blogs/goodmath/2013/01/13/define-distance-differently-
the-p-adic-norm/
2. http://en.wikipedia.org/wiki/P-adic_number
References about the Cantor set.
1. http://planetmath.org/cantorset
2. http://mathworld.wolfram.com/CantorSet.html
3. http://en.wikipedia.org/wiki/Cantor_set
References about the Heine-Borel Theorem.
1. http://en.wikipedia.org/wiki/Heine%E2%80%93Borel_theorem
References about Hausdorff spaces (this has not been explicitly defined in the course,
but the notion has appeared several times in examples).
1. http://en.wikipedia.org/wiki/Hausdorff_space#Examples_and_counterexamples
References about the classification of surfaces.
1. http://www.open.edu/openlearn/science-maths-technology/mathematics-and-
statistics/mathematics/surfaces/content-section-0
2. http://www.math.uchicago.edu/~may/VIGRE/VIGRE2008/REUPapers/Huang.pdf
3. http://www.cis.upenn.edu/~jean/surfclass-n.pdf