5 Lecture Notes in General Topology
5 Lecture Notes in General Topology
Table of Contents
Table of Contents i
Lecture 2 Basis 3
2.1 Basis for a topology . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2 Subbasis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
Lecture 4 Continuity 8
4.1 Continuous functions . . . . . . . . . . . . . . . . . . . . . . . . . 8
4.2 Homeomorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
4.3 Constructing continuous functions . . . . . . . . . . . . . . . . . 10
i
ii TABLE OF CONTENTS
Lecture 10 Compactnesses 31
10.1 Proof continued . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
10.2 Uniform continuity . . . . . . . . . . . . . . . . . . . . . . . . . . 32
10.3 Local compactness . . . . . . . . . . . . . . . . . . . . . . . . . . 33
10.4 One-point compactification . . . . . . . . . . . . . . . . . . . . . 33
Lecture 15 Analysis 47
15.1 Stone–Čech compactification . . . . . . . . . . . . . . . . . . . . 47
15.2 Selected topological results in analysis . . . . . . . . . . . . . . . 49
15.3 Baire category theorem . . . . . . . . . . . . . . . . . . . . . . . 50
Lecture 24 Retraction 74
24.1 Fundamental group of projective space . . . . . . . . . . . . . . . 74
24.2 Brouwer fixed-point theorem . . . . . . . . . . . . . . . . . . . . 74
Index 84
iv TABLE OF CONTENTS
LAYING THE GROUNDWORK 1
Example 1.1.4. The set T composed of all subsets of X is called the discrete
topology on X.
This is useless for a different reason: everything is an open set, so every set
in it looks geometrically the same. N
Example 1.1.5. Let X = a, b, c . There are many different topologies for
X. We have as before T1 = ∅, X , and
T2 = ∅, X, a , b , b , a, b , a, c , b, c ,
a topology is satisfied—trying the last two axioms we find that they, too, are
satisfied.
This is not immediately obvious, because:
Remark 1.1.6. Not every of subsets of X is a topology
collection for X. For in-
stance, T4 = ∅, X, a , b is not a topology, since a ∪ b = a, b 6∈
T4 , so the second axiom isviolated.
T
Similarly, 5 = ∅, X, a, b , b, c is not a topology either, since a, b ∩
b, c = b 6∈ T5 .
N
Definition 1.1.7 (Closed set). A subset S of X is defined to be closed if X \ S
is open (which of course implies there’s an underlying topology we’re working
with).
By De Morgan’s law, we therefore have the following properties of closed
sets:
(i) X \ ∅ = X and X \ X = ∅ are closed,
(ii) any intersection of closed sets is closed, and
(iii) any finite union of closed sets is closed.
The following exercises review some basic set theory which will soon come
in handy.
Exercise 1.1. State De Morgan’s laws and verify them.
Exercise 1.2. (a) What is an equivalence relation on a set?
(b) Let “∼” be an equivalence relation on a set A. Let C1 and C2 be two
equivalence classes. Show that either C1 = C2 or C1 ∩ C2 = ∅.
Exercise 1.3. Let X be a set and let T be a family of subsets U of X such that
X \ U is finite, together with the empty set ∅. Show that T is a topology. (T
is called the cofinite topology of X.)
Exercise 1.4. (a) Show that Q is countable.
(b) Is the polynomial ring Q[x] countable? Explain why?
∞
Q
Exercise 1.5. (a) Let A = 0, 1 . Define AN := Ai , where Ai = A for
i=1
i = 1, 2, . . . . Show that AN is uncountable.
(b) Show that R is uncountable.
Exercise 1.6. Let X be a topological space. Show that if U is open in X and A
is closed in X, then U \ A is open and A \ U is closed in X.
BASIS 3
Definition 1.1.8 ((Strictly) finer topology). Suppose T and U are two topolo-
gies for X. If T ⊃ U (respectively T ) U ), then we say that T is finer
(respectively strictly finer) than U .
In other words, an open set in (X, U ) is also open in (X, T ), but not
necessarily the other way around.
Example 1.1.9. Consider in R the sets I1 = (0, 1), I2 = ( 21 , 2), and I3 = (2, 3)
are open. But they are not the only open sets: we are allowed to take arbitrary
unions and finite intersections. So, for instance, I1 ∪ I3 = (0, 1) ∪ (2, 3) is also
open. N
However the open sets themselves are still the building blocks of open sets;
the open sets themselves can be complicated, like the last example, but we can
generate all of them by only looking at open intervals.
Lecture 2 Basis
2.1 Basis for a topology
In R2 ,
the basic building block of an open set is an open ball, B(a, r) = x |x −
1| < r . But this is not the only open set. The key, however, is that for any
point in an arbitrary open set, we can find an open ball around the point that
is contained in the set.
Proof. For the forward direction, suppose B is a basis. Then for any open
neighbourhood U of x, [
U= Vα ,
α
Proof. In the forward direction, (i) follows from X being open. For (ii), since
U ∩ V is open, we must have some W ∈ B between x and U ∩ V by the previous
theorem.
For the opposite direction, Let T be the collection of all subsets of X that
are unions of sets in B.
By convention, we also take ∅ ∈ T , as the empty union.
If T is a topology, then by definition B is a basis for it. Hence we claim
that T is a topology for X.
First, if x ∈ X, by (i), x ∈ Ux ∈ B. Hence
[
X= Ux ∈ T ,
x∈X
Exercise 2.1. Let B and B 0 be bases for the topologies T and T 0 respectively,
on X. Then the following are equivalent:
2.2 Subbasis
Exercise 2.2. Let X be a set and let C be a collection of subsets of X.
(a) There exists a unique smallest topology T on X such that C ⊂ T .
(b) Let B be the collection of subsets of X consisting of X, ∅, and all finite
intersections of sets in C . Show that B is a basis for T generated by
C.
Definition 2.2.1 (Generated topology, subbasis). The topology T is called
the topology generated by C and C is called a subbasis for T .
B = U1 × U2 × · · · × Un Ui is open in Xi for i = 1, 2, . . . .
8 CONTINUITY
We should check that this satisfies the criterion for being a basis in Theo-
rem 2.1.3. The first one, (x1 , x2 , . . . , xn ) ∈ X1 × X2 × · · · × Xn ∈ B, is trivial.
Secondly, if
x ∈ V1 × V2 × · · · × Vn ⊂ U1 × U2 × · · · × Un ⊂ W
is a subbasis for the topology of X × Y . So the projection maps pull back open
sets to open sets in the product topology.
In other words, the product topology makes the projection map continuous.
In a sense, this is a better description of the product topology, because it works
for infinite products too—more on that in future.
Lecture 4 Continuity
4.1 Continuous functions
Recall from analysis:
As a local condition, a function f : R → R is continuous at x0 if for any
ε > 0 there exists δ > 0 such that |x − x0 | < δ implies |f (x) − f (x0 )| < ε.
As a global condition, f : R → R is continuous on R if f is continuous at
all points in R. An important result from analysis is that this is equivalent to
f −1 (V ) is open in R for any open V ⊂ R.
This second option is the correct/productive way to generalise continuity to
arbitrary topological spaces, since all we have is a notion of openness:
Date: September 3rd, 2020.
4.1 Continuous functions 9
f −1 (C) = f −1 (Y \ U ) = f −1 (Y ) \ f −1 (U ) = X \ f −1 (U ),
so f −1 (U ) is open.
Proposition 4.1.4. Let X1 , X2 , X3 be topological spaces. Let f : X1 → X2 and
g : X2 → X3 be continuous. Then g ◦ f : X1 → X3 is continuous.
Proof. This is obvious: pull an open set U ⊂ X3 back to X1 via X2 .
10 CONTINUITY
Exercise 4.1. Let X be a topological space and let Eα α∈Λ be a collection of
S
subsets such that X = Eα . Let f : X → Y . Suppose f E is continuous for
α
α∈Λ
all α ∈ Λ.
(a) Show that if Λ is finite and Eα is closed for all α ∈ Λ, then f is continuous.
(b) Give an example where Λ is countable and Eα is closed for all α ∈ Λ, but
f is not continuous.
(c) Suppose Eα α∈Λ is locally finite, that is each x ∈ X has a neighbour-
hood that intersects only finitely many Eα , and Eα is closed for all α ∈ Λ.
Show that f is continuous.
4.2 Homeomorphism
Definition 4.2.1 (Homeomorphism). Let X and Y be topological spaces. Let
f : X → Y be one-to-one and onto. The map f is called a homeomorphism
if both f and f −1 are continuous.
In other words, f is a homeomorphism if is has the property that U is open
in X if and only if f (U ) is open in Y .
S
(ii) If X = Vα , Vα open and f Vα is continuous, then f : X → Y is contin-
α
uous.
(f Z )−1 (U ) = f −1 (U ) ∩ Z
π1 : X → Y → X, π1 (x, y) = x
π2 : X → Y → Y, π2 (x, y) = y
f −1 (U × V ) = f1−1 (Y ) ∩ f2−1 (V ).
Example 5.1.2. Assume Xα are compact for all α ∈ I. The product space
Y
Xα
α∈I
Example 5.1.3. Assume Xα are connected for all α ∈ I. The product might
be totally disconnected, meaning the connected component contains only one
point. N
with Uα open in Xα for all α ∈ I, like with the box topology, only this time
Uα = Xα for all but finitely many α.
Hence the topology on X generated by B, called the product topology,
has much fewer open sets than the box topology (if I is infinite), i.e.:
Remark 5.2.1. If the index set I is finite, then the box topology and the product
topology coincide.
Date: September 10th, 2020.
5.3 Metric topology 13
Remark 5.2.2. Since this is much more useful than the box topology, we shall
assume that Y
X= Xα
α
πα : X → Xα
kx + yk ≤ kxk + kyk
for all x, y ∈ X.
Norms induce metrics:
Lemma 5.4.2. Let (X, k·k) be a normed linear space. Then d(x, y) := kx − yk
for x, y ∈ X is a metric on X.
Proof. Clearly d(x, y) ≥ 0 and d(x, y) = 0 if and only of kx − yk = 0 if and only
if x − y = 0 if and only if x = y.
The metric is symmetric since
Finally
Definition 5.4.3 (Normed topology). Let (X, k·k) be a normed linear space.
The topology induced by the metric d(x, y) = kx − yk is called the normed
topology of x induced by k·k.
Example 5.4.4. The Euclidean space Rn is a dimension n vector space over
R. There are several common norms and metrics on this:
(i) The Euclidean norm and metric: For x = (x1 , x2 , . . . , xn ), we have the
usual Euclidean norm
kxk∞ = max|xi |. N
i
Here’s a natural and interesting question: we have one and the same set
X = Rn , with three different kinds of norms and metrics. How do their open
sets, in the induced metric topologies, relate?
The answer is that the topologies are the same, which we will discover below.
Definition 5.4.5 (Equivalent norms). Let k·k1 and k·k2 be two norms on
a linear space X. We say that k·k1 and k·k2 are equivalent if there exist
constants c and k such that
For y ∈ B2 (x0 , ε), there exists εy > 0 such that B2 (y, εy ) ⊂ B2 (x0 , ε) since
the ball is open in k·k2 .
ε
We know B1 (y, ky ) ⊂ B2 (y, εy ), where notably the first ball is open in
(X, k·k1 ). Thus ε
y
[
B2 (x0 , ε) = B1 y,
k
y∈B2 (x0 ,ε)
In other words, if kxk2 < r, then kxk1 < 1. Take C = 2r . Then if kxk2 ≤ C
implies kxk1 < 1.
cx
Then for any x 6= 0 in X, consider y = kxk 2
. Then kyk2 = c. Hence
kyk1 < 1, so
cx
< 1,
kxk2
kxk2 ≤ kkxk1
for some k > 0. Hence the norms k·k1 and k·k2 are equivalent.
Let k·k be any norm on X. We want to show that k·k is equivalent to k·k∞ .
If so, then any two norms on X induce the same topology on X, which in turn
implies any two norms are equivalent.
By the triangle inequality,
Xn
X n
kxk =
ci ei
≤ |ci |kei k ≤ Kkxk∞
i=1 i=1
In other words,
kxk ≤ Kkxk∞
for all x ∈ X.
Next we need to show that there exists some c > 0 such that ckxk∞ ≤ kxk
for all x ∈ X. This requires two things from analysis which we will also discuss
later:
Definition 6.1.2 (Compact set). A set is compact if every open cover has a
finite subcover.
ε
by taking δ = K . Since k·k : X → R is continuous and B is compact with
respect to k·k∞ , the norm k·k must have a minimum on B, say c > 0.
In other words, if kxk∞ = 1, then kxk ≥ c. Then for any nonzero x ∈ X,
consider y = kxkx ∞ . Then kyk∞ = 1, so kyk ≥ c, whereby
x
≥ c,
kxk∞
or in other words
kxk ≥ ckxk∞ .
We leave it as an exercise to check the two details left out in the above
argument:
Exercise 6.1. Verify that k·k∞ is indeed a norm on X.
Exercise 6.2. Verify that B = x kxk∞ = 1 is compact in (X, k·k∞ ).
Corollary 6.1.3. All norms on Rn are equivalent. I.e., Rn has a unique norm
topology.
18 CONNECTEDNESS AND COMPACTNESS
Theorem 6.2.2. Let X be a topological space and (Y, d) be a metric space. Let
fn : X → Y be continuous, n = 1, 2, 3, . . . . Suppose fn converges uniformly
to f . Then f is continuous.
d(f (x), f (x0 )) ≤ d(f (x), fN (x)) + f (dN (x), fN (x0 )) + d(fN (x0 ), f (x0 ))
ε ε ε
< + + = ε.
4 2 4
Hence f (x) ∈ B(y0 , ε).
We know from analysis that if the convergence is not uniform, the limiting
function needn’t be continuous,
but it might be continuous on some set.
So, question: Let fn be a sequence of continuous functions. Suppose
fn (x) → f (x) pointwise for all x ∈ X. How large is the set
x ∈ X f is continuous at x ?
6.3 Connectedness
Definition 6.3.1 (Connected, disconnected). (i) A topological space X is
called disconnected if there exist open sets U and V such that X = U tV ,
i.e. U, V 6= ∅ and U ∩ V = ∅.
Remark 6.3.2. If X is disconnected, then the U and V in question are also closed
sets, since U = X \ V .
We immediately get the following characterisation:
6.3 Connectedness 19
is connected.
Proof. Let [
E= Eα
α∈Λ
and let F 6= ∅ be a subset of E that is both open and closed. We need to show
that F = E, which is equivalent to Eα ⊂ F for all α.
Let x ∈ F . Then x ∈ Eα0 for some α0 ∈ Λ since F ⊂ E.
Now F ∩Eα0 6= ∅, since they meet in at least x, and this set is both open and
closed in Eα0 . But Eα0 is by assumption connected, meaning F ∩ Eα0 = Eα0 is
the only option. In other words, Eα0 ⊂ F .
For any β ∈ Λ, F ∩ Eβ is open and closed in Eβ . But this can’t be empty,
since Eα0 ⊂ F . I.e.,
(Eα0 ∩ Eβ ) ⊂ (F ∩ Eβ )
implies F ∩ Eβ 6= ∅, so F ∩ Eβ = Eβ . Hence Eβ ⊂ F for all β ∈ Λ. Thus
E = F.
20 DIFFERENT KINDS OF CONNECTEDNESS
is connected.
Exercise 7.3. Verify that this set (known as the topologist’s sine curve) is
not path-connected.
Remark 7.2.5. This gives an example showing that the closure of a path con-
nected set is not necessarily path connected.
Exercise 7.4. Prove that the path components of a locally path-connected space
coincide with the connected components.
Exercise 7.5. Show that an open subset of Rn is connected if and only if it is
path-connected.
7.3 Compactness
Definition 7.3.1 (Compact). A topological space X is compact if every open
cover of X has a finite subcover.
In other words, for any family of open sets Uα α∈Λ with
[
X= Uα
α∈Λ
Proof. Let X be compact and E ⊂ X closed. For any open cover Uα α∈Λ
of
E, [
(X \ E) ∪ Uα
α∈Λ
f −1(Uα )
α∈Λ
Mapping back,
f (X) ⊂ Uα1 ∪ Uα2 ∪ · · · ∪ Uαn .
Hence f (X) is compact.
24 TYCHONOFF’S THEOREM
Like with the Intermediate value theorem above, this gives a slick proof of
the Extreme value theorem in analysis.
Recall another theorem from analysis:
Theorem 7.3.8 (Heine–Borel theorem). Let E ⊂ Rn . Then E is compact if
and only if E is closed and bounded.
Remark 7.3.9. Let X be a metric space. Then E ⊂ X implies E is closed and
bounded (closed because metric spaces are Hausdorff).
The converse is not true in general, though it is true for any finite-dimensional
vector space.
C = Uα × Vα α∈Λ
Then
Vy := Vy,1 ∩ Vy,2 ∩ · · · ∩ Vy,n
is an open neighbourhood of y in X2 (it’s a finite intersection of open sets).
Note that for π2 : X1 × X2 → X2 defined by π2 (x, y) = y, π2−1 (Vy ) is covered
by
Uy,1 × Vy,1 , Uy,2 × Vy,2 , . . . Uy,n × Vy,n .
Since Vy y∈X2 is an open cover of X2 and X2 is compact,
This seems very natural, but it can’t be proved from basic set theory axioms.
It also has some seemingly unnatural consequences, like the Banach–Tarsky
theorem.
is partially
ordered.
Let Dα α∈Λ be a totally ordered subset of P. Set
[
D= Dα .
α∈Λ
from Dα1 ⊂ Dα2 ⊂ · · · ⊂ Dαn . Then these sets must belong to Dα for some
α ∈ Λ, which contradicts Dα having no such finite subcover.
By Zorn’s lemma, P therefore has a maximal element, say E .
Finally we claim that E ∩ B covers X. If true, since E ⊂ P says this is a
cover of X by sets in B that has no finite subcover. This is a contradiction.
Let x ∈ X. It suffices to show that there exists V ∈ E ∩ B such that x ∈ V .
Choose U ∈ E such that x ∈ U . Since B is a subbasis there exists
V1 , V2 , . . . , Vm ∈ B such that
x ∈ V1 ∩ V2 ∩ · · · ∩ Vm ⊂ U.
But V1 ∩ V2 ∩ · · · ∩ Vm ⊂ U by choice of U , so
[
X=U∪ Wi,j ,
i,j
Now take B to be the subbasis for the product topology of X of the form
−1
α (U α ), α ∈ Λ, where Uα is open in Xα . These pullbacks look like Uα ×
π
Xβ β6=α .
Let D be a family of subsets such that D ⊂ B and D has no finite subcover
of X.
We want to show that D does not cover X.
For any β ∈ Λ, consider the open subsets V ⊂ Xβ such that πβ−1 (V ) ∈ D.
The family of such sets V cannot cover Xβ . Otherwise, since Xβ is compact,
we would have a finite subcover of Xβ , say V1 , V2 , . . . , Vn , for which πβ−1 (Vi ) ∈ D.
But those pullbacks would cover X, and D has no finite subcover of X.
Hence we can choose a point
Note that this last notion, in Rn , is usually discussed in terms of the Bolzano–
Weierstrass theorem, saying that any bounded sequence has a convergent
subsequence (since a bounded sequence is contained in a compact set).
Note that in Rn , all of these notions are equivalent. We will show that in
fact in any metrisable space, all three notions are equivalent.
Recall that x ∈ X is a limit point of a subset A ⊂ X if for every open
neighbourhood U of x, we have U ∩(A\ x ) 6= ∅. In other words, x ∈ A \ x .
Proof. We need to show that if A ⊂ X is infinite, then A has a limit point. This
is equivalent to showing that if A has no limit point, then A is a finite set.
Assume A has no limit point.
We claim A is closed, i.e. X \ A is open.
For any x ∈ X \ A, x is not a limit point of A. Hence x 6∈ A \ x = A.
This means x ∈ X \ A which is open, so there exists an open neighbourhood U
of x such that U ⊂ X \ A. Therefore x ∈ U ⊂ (X \ A) ⊂ X \ A. Hence X \ A
is open, so A is closed.
Our goal is to show that A is finite. To this end, notice how for each x ∈ A,
since x is not a limit point
of A, there exists an openneighbourhood
Ux of x
such that Ux ∩ (A \ x ) = ∅. This means Ux ∩ A = x , i.e., every point in
A is an isolated point.
Now consider
C = X \ A ∪ Ux x ∈ A .
is a finite set.
Remark 9.2.4. In general, limit point compactness does not imply compactness.
Counterexample 9.2.5. Let Y = a, b be a two point set with the trivial
topology, i.e., T = ∅, Y . Let X = Z+ × Y where
Z+ = n > 0 n ∈ Z
with the discrete topology (so every single point is an open set).
Then X is limit point compact. In fact, for any nonempty A ⊂ X, A has a
limit point. If (n, a) ∈ A, then (n, b) isa limit point
of A.
But X is not compact. Take C = Un = n × T n ∈ Z+ . This is an
open cover of X, but it has no finite subcover since removing any set makes it
miss a point of Z+ . N
Theorem 9.2.6. Let X be a metrisable space. Then the following are equiva-
lent:
(i) X is compact.
Proof. That (i) implies (ii) is true in general: this is Theorem 9.2.3.
To show (ii) implies
(iii), let xn be a sequence in X. Consider the set
A = xn n ∈ Z+ .
First, if A is a finite set, then xn must repeat at least one element infinitely
many times. I.e.,there exists some x ∈ X such
that x = xn for infinitely many
n. This means xn has a subsequence xnk = x n which is a constant
k
sequence, hence convergent (to x).
Second, if A is an infinite set, then since X is limit point
compact A must
have a limit point, say x. We select a subsequence of xn that converges to
x as follows.
For any δ > 0, B(x, δ) ∩ A has infinitely many points since x is a limit
point, because otherwise there exists some sufficiently small δ 0 > 0 such that
B(x, δ 0 ) ∩ A = ∅, contradicting x being a limit point of A.
Thus we can choose n1 < n2 < n3 < . . . such that xni ∈ B(x, 1i ) ∩ A. Then
∞
xni i=1 converges to x.
(Note that this is really the same argument one uses to prove the Bolzano–
Weierstrass theorem in Rn .)
Finally, (iii) implies (i). Assume X is sequentially compact with a metric d.
First, an important lemma:
1 ε
d(y, x) ≤ d(y, xni ) + d(xni , x) < + < ε.
ni 2
Lecture 10 Compactnesses
10.1 Proof continued
Proof of Theorem 9.2.6, continued. We are left with (iii) implying (i), for which
we showed the Lebesgue number lemma.
Next, if X is sequentially compact, then given any ε > 0, there exists a finite
cover of X by ε-balls (in other words, X is totally bounded).
We show this by contradiction. Assume there exists some ε > 0 such that
X cannot be covered by finitely many ε-balls.
Choose x1 ∈ X, then x2 ∈ X \ B(x1 , ε), x3 ∈ X \ (B(x1 , ε) ∪ B(x2 , ε)), and
so on. The sets we are removing are nonempty by the assumption of ε-balls not
covering X.
Hence we get a sequence xn in X that does not have a convergent sub-
sequence, since d(xn+1 , xi ) ≥ ε for i = 1, 2, . . . , n. This contradicts X being
sequentially compact.
Third: we are not equipped to show that if X is sequentially compact, then
X is compact.
Let C be an open cover of X. By Lebesgue number lemma, there exists
some δ > 0 such that for any V ⊂ X with diam(V ) < δ, we have V ⊂ U for
some U ∈ C .
Take ε = 3δ . By the second step, we know X can be covered by finitely many
ε-balls. In other words,
Now diam(B(xi , ε)) < δ, so B(xi , ε) ⊂ Ui for some Ui ∈ C . Hence from the
covering of balls we get a finite cover
X = U1 ∪ U2 ∪ · · · ∪ Un
of X.
Remark 10.1.1. There exist sequentially compact spaces that are not compact.
There also exist compact spaces that are not sequentially compact.
For posterity we write down the definition from inside the proof:
Definition 10.1.2 (Totally bounded). A metric space X is totally bounded
if for any ε > 0 there is a finite cover of X by ε-balls.
In other words, we showed in the second step of the Theorem 9.2.6 that
Corollary 10.1.3. If X is a compact metric space, then X is totally bounded.
The converse is not true: we will show later that if X is a metric space, X
is compact if and only if X is complete and totally bounded.
Exercise 10.1. Let X be a metric space.
(a) Show that if E is a compact subset of X, then E is closed and bounded.
(b) Give an example to show the converse is false. In other words, find a metric
space in which not every closed and bounded subset if compact.
Date: September 29th, 2020.
32 COMPACTNESSES
C = f −1 (Vy ) y ∈ Y
be the graph of f .
for all x, y ∈ X. Show that there exists a unique point x ∈ X such that f (x) = x.
This is known as the Banach fixed-point theorem.
10.3 Local compactness 33
A \ Y \C
is a finite subcover of Y .
That Y is Hausdorff is easy. Let x, y ∈ Y . If x, y ∈ X, then since X is
Hausdorff, there exist open U, V ⊂ X such that x ∈ U , y ∈ V , and U ∩ V = ∅.
If x ∈ X and y = ∞, since X is locally compact, there exists an open U
such that x ∈ U and U is compact. Then y = ∞ ∈ Y \ U , which is open in Y ,
and (Y \ U ) ∩ U = ∅.
Example 10.4.3. The
one-point compactification
R ∪ ∞ is homeomorphic
to S 1 = (x, y) ∈ R2 x2 + y 2 = 1 . N
Similarly,
Example 10.4.4. The one-point compactification Rn ∪ ∞ ' S n , where
N
Example 10.4.5. The one-point compactification C ∪ ∞ is the Riemann
sphere. N
Example 10.4.6. Since (0, 1) and R are homeomorphic, so must their one-point
compactification be, i.e.
(0, 1) ∪ ∞ ' R ∪ ∞ = S 1 .
On the other hand, we can compactify (0, 1) by adding both endpoints, i.e.,
[0, 1]. But these two spaces are not homeomorphic. N
is open.
Conversely, for distinct x, y ∈ X, X \ x is open since x is closed.
Hence there exists an open U such that y ∈ U and x 6∈ U .
Exercise 11.5. Let X be a topological space and let X0 be the topological space
made up of the set X with the cofinite topology. Show that the identity map
from X to X0 is continuous if and only if X is a T1 -space.
(c) R \ Q is T -closed.
This provides an example of a Hausdorff space (T2 ) that is not regular (T2 ).
In other words,
T4 =⇒ T3 =⇒ T2 =⇒ T1 .
and r(y)
[
V = B y,
2
y∈F
r(x) r(y)
d(x, y) ≤ d(x, z) + d(z, y) < + ≤ max r(x), r(y) .
2 2
But this means x ∈ B(y, r(y)) or y ∈ B(x, r(y)), which is a contradiction, since
r(x) and r(y) were radii specifically chosen to separate x and y.
E ⊂ U 21 ⊂ U 21 ⊂ V.
Since E is closed and U 12 is open, we can insert, using the lemma, another pair
U 14 ⊂ U 14 between those, and similarly between U 12 and V we insert U 34 ⊂ U 34 .
Hence
E ⊂ U 14 ⊂ U 14 ⊂ U 21 ⊂ U 21 ⊂ U 34 ⊂ U 34 ⊂ V.
Now repeat this process ad nauseam between each closed-open pair, at each
finite step adding sets for more and more dyadic rational numbers between 0
and 1. Hence we construct open sets Ur for every dyadic rational 0 < r < 1
such that
Ur ⊂ Us
for 0 < r < s < 1,
E ⊂ Ur
for all 0 < r < 1, and
Ur ⊂ V
for all 0 < r < 1. Now the idea is to use the boundaries of the Ur as level
curves f = r for a function f : X → [0,
1]. I.e., define f (x) = 0 if x ∈ Ur for all
0 < r < 1 and f (x) = sup r x 6∈ Ur .
Thus 0 ≤ f (x) ≤ 1, f = 0 on E, and f = 1 on F .
It remains to verify that f is continuous. Let x ∈ X, and let us show that f
is continuous at x.
First, suppose f (x) = 0. For any ε > 0, take a dyadic rational r such that
0 < r < ε. Then x ∈ Ur (since otherwise f (x) ≥ r > 0). Hence Ur is an open
Date: October 6th, 2020.
12.1 Proof of Urysohn’s lemma 39
Theorem 12.1.3 (Tietze extension theorem). Let X be a normal space and let
Y ⊂ X be a closed subset. Let f be a bounded continuous real-valued function
on Y . Then there exists a bounded continuous real-valued function h on X such
that h = f on Y .
n c0 o
E0 = y ∈ Y f (y) ≤ −
3
and n c0 o
F0 = y ∈ Y f (y) ≥ .
3
Then E0 and F0 are closed (since they’re continuous pullbacks of closed sets) and
disjoint subsets of X. By Urysohn’s lemma there exists a continuous real-valued
function g0 on X such that g0 = − c30 on E0 and g0 = c30 on F0 , and
c0 c0
− ≤ g0 (x) ≤
3 3
c0
for all x ∈ X. In particular, |g0 (x)| ≤ 3 for all x ∈ X, and
2
|f (x) − g0 (x)| ≤ c0
3
for all x ∈ Y (since f is only defined on Y ). In other words we have constructed
a continuous function g0 which approximates f with a maximum error of 23 c0 .
40 URYSOHN’S LEMMA
∞
Step 2: Construct a sequence gn (x) n=0
inductively such that |gn (x)| ≤
2n
3n+1 c0 for all x ∈ X and
∞
X 2n+1
f (x) − gk (x) ≤ n+1 c0
3
k=0
for all x ∈ Y . Then rinse and repeat with f2 = f (x) − g0 (x) − g1 (x), and so on.
Step 3: Let
Xn
hn (x) = gk (x).
k=0
Then hn (x) converges to a bounded continuous function h(x) on X and
h = f on Y .
We show this by showing hn (x) is uniformly Cauchy in X, which means
hn → h for some continuous function h on X. For n ≥ m,
|hn (x) − hm (x)| = |gm+1 (x) + · · · + gn (x)| ≤ |gm+1 (x)| + · · · + |gn (x)|
2 m+1 2 n c
0
≤ + ··· +
3 3 3
2 m+1 2 2 2 c
0
≤ 1+ + + ...
3 3 3 3
2 m+1 1 c 2 m+1
0
≤ 2 = c0 → 0
3 1− 3 3 3
uniformly (for all x ∈ X) as m → ∞. Hence hn (x) is uniformly Cauchy.
Thus hn (x) → h(x) is continuous.
Moreover
∞ ∞ n
X X 2 c0
|h(x)| ≤ |gn (x)| ≤ = c0 .
n=0 n=0
3 3
Finally, for x ∈ Y ,
2 n+1
|f (x) − h(x)| ≤ |f (x) − hn (x)| + |hn (x) − h(x)| ≤ c0 + |hn (x) − h(x)|.
3
The last term converges to 0 uniformly for all x ∈ X, and the first part also
goes to 0 uniformly as n goes to 0. Hence f (x) = h(x) for all x ∈ Y .
We should always keep in our mind that metric spaces are particular exam-
ples of normal spaces.
Proof. Let B be a countable basis for X. Let E and F be two disjoint closed
sets in X. For each x ∈ E, by regularity there exist open sets U and V such
that x ∈ U , F ⊂ V , and U ∩ V = ∅.
Since x ∈ U ⊂ (X \V ), where X \V is closed, U ⊂ (X \V ). Thus U ∩F = ∅.
Now since U is open, take Ux ∈ B such that x ∈ Ux ⊂ U . Repeat
∞ this for
every x ∈ E. Since B is countable, we have an open cover Ui i=1 of E with
Ui ∩ F = ∅, with Ui ∈ B. ∞
Similarly F has an open cover Vi i=1 with Vi ∩ E = ∅ and Vi ∈ B.
Let
n
[
Un0 := Un \ Vi
i=1
and
n
[
Vn0 := Vn \ Ui ,
i=1
and
∞
[
0
F ⊂ V := Vn0 ,
n=1
and importantly U 0 ∩ V 0 = ∅.
Date: October 8th, 2020.
42 URYSOHN’S METRISATION THEOREM
Lemma 13.1.2. Let X be a regular space with a countable basis. Then there
exist a countable collection of continuous functions fn : X → [0, 1] such that for
any x0 ∈ X and any neighbourhood U of x0 there exist an index n such that
fn (x0 ) > 0 and fn = 0 on X \ U .
∞
Proof. Let Bn n=1 be a countable basis for X. For each pair Bn , Bm with
Bn ⊂ Bm we have that Bn and X \ Bm are disjoint closed sets.
By Lemma 13.1.1, X is normal, and so we can apply Theorem 12.1.1 to
construct a continuous function gn,m : X → [0, 1] with gn,m (Bn ) = 1 and
gn,m (X \ Bm) = 0.
Reindex gn,m to fn . Now for x0 ∈ X and x0 ∈ U ⊂ X open, we can
choose Bn and Bm such that x0 ∈ Bn ⊂ Bm ⊂ U (by regularity).
Then gn,m (x0 ) = 1 and gn,m (X \ Bm ) = 0. Hence g(X \ U ) = 0 since
Bm ⊂ U .
Now the idea for Urysoh’s metrisation theorem is to show that X, using this
∞
Q
previous lemma, embeds homeomorphically to the infinite product RN = R.
i=1
Let
¯ b) := min |a − b|, 1
d(a,
be the standard bounded metric on R.
Exercise 13.1. Verify that d¯ is a metric on R.
Let Y = RN . For x = (xi ) and y = (yi ) in Y , define
¯ i , yi ) o
n d(x
D(x, y) = sup .
1≤i<∞ i
Proof. Let U be an open set in the metric topology and let x = (xi ) ∈ U .
We want to show that there exists an open set V in the product topology
such that x ∈ V ⊂ U . (Thus U is open in the product topology.)
Since U is open in the metric topology, choose ε > 0 such that BD (x, ε) ⊂ U .
Take N ∈ N large enough such that N1 < ε. Let
V = (x1 − ε, x1 + ε) × · · · × (xN − ε, xN + ε) × R × R × · · · × R.
13.1 Preliminaries 43
This is open in the product topology (it is a basis element). For y = (yi ) ∈ V ,
if i ≥ N then
¯ i , yi )
d(x 1
≤ < ε.
i N
If i < N , then
¯ i , yi )
d(x ε
≤ < ε.
i i
Thus, taking supremum, D(x, y) < ε, so y ∈ BD (x, ε), meaning that V ⊂
Bd (x, ε) < U , as desired.
Q∞
Conversely, let V = Vi be a basis element for the product topology. In
i=1
other words, Vi = R for i > N for some N ∈ N and Vi is open in R for i ≤ N .
For x = (xi ) ∈ V there exists an open U in the metric topology such that
x ∈ U ⊂ V . To see this, for each i ≤ N , choose 0 < εi < 1 such that
(xi − εi , xi + εi ) ⊂ Vi . Define
nε o
i
ε = min .
1≤i≤N i
|xi − yi | < εi ,
so yi ∈ (xi − εi , xi + εi ) = Vi . Therefore U ⊂ V .
is an embedding of X into RJ .
Tychonoff’s theorem, infinite case and Proposition 5.2.3 the product is too). By
Lemma 14.1.7, X has a compactification induced by h.
both exist and are equal (since in the compactification those endpoints
are the same).
(ii) On the other hand, let Y = [0, 1]. Then a bounded continuous function
f : X → R is extendable to Y if and only if
exist. Here they don’t need to be equal, since the endpoints are not
identified. N
Proof. Let fα α∈J be the collection of all bounded continuous real-valued
functions on X. For each α ∈ J, define Iα = [inf fα (X), sup fα (X), ⊂]R (since
fα is bounded).
Q
Then define h : X → Iα by h(x) = (fα (x))α∈J .
α∈J
Since X is completely regular, fα satisfies the condition inQthe Embedding
theorem. Note also that by Tychonoff’s theorem, infinite case, Iα is compact
α∈J
Hausdorff.
Let Y be the compactification of X induced by h. Then we have an embed-
ding
Y
H: Y → Iα
α∈J
with H = h on X.
The construction of Y is independent of X since we built it using the set of
all bounded continuous real-valued functions on X.
Q a bounded continuous f : X → R, then f = fβ for some β ∈ J.
Now given
Let πβ : Iα → Iβ be the projection map. Then
α∈J
H Q πβ
πβ ◦ H : Y Iα Iβ
α∈J
is continuous. For x ∈ X,
so πβ ◦ H extends f = fβ to Y .
Lecture 15 Analysis
15.1 Stone–Čech compactification
The uniqueness of the extension of f to Y is a consequence of the following
lemma:
X ι2 =f2
Y2
Similarly ι1 = f1 extends to F1 : Y2 → Y1 .
Consider
F2 F1
F1 ◦ F2 : Y1 Y2 Y1 .
For x ∈ X,
F1 ◦ F2 (x) = F1 (ι2 (x)) = F1 (x) = ι1 (x) = x,
so F1 ◦ F2 = Id on X. Hence F1 ◦ F2 = Id on Y1 .
Similarly, F1 ◦ F2 = Id in Y2 . Hence Y1 and Y2 are homeomorphic, because
F1 and F2 are continuous inverses between the two.
Definition 15.1.4 (Stone–Čech compactification). The compactification of X
in Theorem 14.2.1 is called Stone–Čech compactification of X, denoted by
β(X).
It is characterised by the property that any continuous map f : X → C,
where C is compact Hausdorff, extends uniquely to a continuous map F : β(X) →
C. In a picture,
β(X)
F
X f
C
15.2 Selected topological results in analysis 49
¯ i , yi ) o
n d(x
D(x, y) = sup
i i
and
¯ i , yi ) = min |xi − yi |, 1 .
d(x
Note that, as we showed in Proposition 13.1.3, the above induces the product
topology on RN , so this space is complete.
Recall that a metric space (X, d) is totally bounded if for every ε > 0 there
exists a finite cover of X by ε-balls.
It is clear that a compact metric space is totally bounded; just cover with
ε-balls around every point and use compactness to pick a finite subcover. The
converse is not true:
For the converse, assume (X, d) is complete and totally bounded. We claim
X is sequentially compact (hence X is compact, since in a metric space they
are equivalent).
In other words,
we need to show that every sequence has a convergent sub-
∞
sequence. Let xn n=1 ⊂ X. This is a standard partition argument; partition
the space into parts containing infinitely many terms of the sequence, keep
subdividing to identify a convergent subsequence.
So first cover X by finitely many balls of radius 1 (doable since X is totally
bounded). Then there exists a ball, say B1 , containing
infinitely many xn (since
finitely many balls but infinite sequence). Say xn n∈J ⊂ B1 for some infinite
1
set J1 ⊂ N.
Now repeat: consider xn n∈J1 and cover X by finitely many radius 21 balls.
At least one ball, say B2 , contains infinitely many xn n∈J1 , say xn n∈J2 ⊂
B2 for some infinite set J2 ⊂ J1 . Rinse and repeat, so xn n∈J with Bk a ball
k
1
of radius 2k−1 and J1 ⊃ J2 ⊃ . . . .
∞
Take xnk k=1 with nk ∈ Jk . We shall show xnk is Cauchy. Thus
xnk → x for some x ∈ X since X is complete.
Given ε > 0, choose N ∈ N such that 2N1−2 < ε. For k, h > N , this means
nk , nj ∈ JN . Hence xnk , xnj ∈ BN , with radius 2N1−1 , so
1 1
d(xnk , xnh ) < 2 = N −2 < ε.
2N −1 2
Hence x nk is Cauchy.
is also dense in X.
Lemma 16.1.1. A space X is Baire if and only if for any countable collection
∞
of closed sets with empty interior En n=1 in X, the union
∞
[
En
n=1
has empty interior. I.e., for any nonempty open set U ⊂ X, there exists x ∈ U
such that x 6∈ En for all n.
Since int(E1 ) = ∅ and U is open, we must have U 6⊂ E1 , so there exists
y1 ∈ U \ E1 .
Note that our assumption on X implies that X is regular. So there exists
an open neighbourhood U1 of y1 such that U1 ∩ E1 = ∅, and U1 ⊂ U .
Now consider U1 and E2 . Then by the same argument there exists y2 ∈
U1 \ E2 and open neighbourhood U2 of y2such that U2 ∩ E2 =∅ and
U2 ⊂ U1 .
Repeating this we obtain a sequence yn and open sets Un such that
yn ∈ Un−1 \ En , Un is an open neighbourhood of yn , Un ∩ En = ∅, and Un ⊂
Un−1 .
Moreover, if X is a metric space, we can also choose Un such that diam Un <
1
n.
Now we claim
∞
\
Un 6= ∅.
n=1
Taking complement,
m
\
∅= Uni = Unm
i=1
52 BAIRE CATEGORY THEOREM
In other words, the set of all x for which the fn are uniformly close.
For any fixed n and m, let
An,m (ε) = x ∈ X d(fn (x), fm (x)) ≤ ε .
Note that An,m (ε) is closed (it’s the pullback of a continuous map).
Thus we have \
En (ε) = An,m (ε)
n,m≥N
E1 (ε) ⊂ E2 (ε) ⊂ . . . ,
and
∞
[
EN (ε) = X
N =1
since for any x ∈ X fixed, fn (x) converges in Y . Hence fn (x) is Cauchy,
so the tail is eventually close to each other, and so lies in some EN (ε).
Now let
∞
[
U (ε) := int(EN (ε)),
N =1
which is open in X.
We make two claims. First, U (ε) is dense in X for any ε > 0.
Second, f is continuous at each point in the set
∞
\ 1
C := U .
m=1
m
Since X is Baire, the first claim implies C is dense in X. Thus the theorem
follows by the second claim.
54 BAIRE CATEGORY THEOREM
To prove the first claim it suffices to show for any nonempty open V in X,
there exists N such that V ∩ int(EN (ε)) 6= ∅.
By Lemma 16.2.2, V is Baire. Second, V ∩ EN (ε) is closed in V . Now
∞
[
X= EN (ε),
N =1
so
∞
[
V = (EN (ε) ∩ V )
N =1
where EN (ε) ∩ V are closed in V . Since V is Baire this means EN (ε) ∩ V
has nonempty interior in V for some N ∈ N. In other words, there exists a
nonempty open W in V such that
W ⊂ (EN (ε) ∩ V ).
But W is open in V and V is open in X, so W is also open in X. Thus
∅ 6= W ⊂ (int(EN (ε)) ∩ V ), and we are done.
Now to prove the second claim. Let x ∈ C. We want to show f is continuous
at x. I.e., for any ε > 0, there exists an open neighbourhood W of x such that
d(f (x), f (y)) < ε for all y ∈ W .
Choose k ∈ N such that k1 < 3ε . Since
∞
\ 1
x∈C= U ,
m=1
m
π : X → X/∼
by π(x) = [x], the equivalence class of x. We would like how new topology to
make sure this projection map is continuous.
To this end we define the quotient topology on X/∼ by U ⊂ X/∼ is open
if and only if π −1 (U ) is open in X.
Exercise 17.1. Check that this in fact defined a topology on X/∼.
Remark 17.1.2. This is the smallest topology on X/∼ such that π is continuous.
f
X Y
π
g
X/∼
Example 17.2.2. The real numbers R with the usual topology is a topological
group, where the group operation is ordinary addition. N
Example 17.2.3. Consider the unit circle S 1 = eiθ 0 ≤ θ ≤ 2π ⊂ R2 with
the subspace topology. This is a topological group under multiplication. N
Proof. Obvious: La and Ra are both one-to-one and onto (their inverses are
La−1 and Ra−1 ).
It remains to check continuity, which in this case means it suffices to check
La and Ra are continuous, since they are their own inverses for some a.
But this is obvious:
embed multiplication
G G×G G
g (a, g) ag
La
is the composition of two continuous maps (embedding and the group multipli-
cation). Likewise for Ra .
Theorem 17.2.8. Let G be a connected topological group. Then for any open
neighbourhood V of the identity e ∈ G, the subgroup generated by V is G. In
other words, hV i = G.
Note that GLn (R) with the usual matrix multiplication is a group. In fact,
GLn (R) is a topological group, in other words the group operations are contin-
uous. First
m : Mn (R) × Mn (R) Mn (R)
(A, B) AB
Z×R R
(n, r) n + r.
G G
g gh−1
Exercise 18.1. Verify that the linear fractional transformation actually defines
a group action.
and
A(x0 , x1 ) = γ : [0, 1] → X continuous, γ(0) = x0 , γ(1) = x1
respectively.
Remark 19.1.4. If γ is a path, we denote its path homotopy equivalence class
by [γ].
Proof. We will show ' is an equivalence relation; path homotopy is very similar.
Reflexivity is obvious: f ' f by F (x, t) = f (x) for all x ∈ X and all t ∈ [0, 1].
For symmetry, suppose f ' g, say by a continuous F (x, t) such that F (x, 0) =
f (x) and F (x, 1) = g(x). Take G(x, t) = F (x, 1 − t). Then G(x, 0) = F (x, 1) =
g(x) and G(x, 1) = F (x, 0) = f (x). Hence g ' f .
Date: October 29th, 2020.
62 HOMOTOPY THEORY
Finally, transitivity. Suppose f ' g and g ' h, say the first one by F (x, t)
and the second by G(x, t). Take
(
F (x, 2t), if 0 ≤ t ≤ 21 ,
H(x, t) =
G(x, 2t − 1), if 21 ≤ t ≤ 1.
N
This argument works in slightly more generality:
Remark 19.1.6. Let γ1 , γ2 be two paths from x0 to x1 in a convex space X.
Then γ1 'p γ2 . (Since in a convex space the line segment connecting the two
at a fixed time is still in the space because of convexity.)
Exercise 19.1. Check that ∗ on the path homotopy classes is well-defined (i.e.,
doesn’t depend on the choice of representatives γ0 and γ1 ).
Proposition 19.2.1. (i) The operation ∗ is associative. In other words, let
γ0 be a path from x0 to x1 , γ1 be a path from x1 to x2 , and γ2 a path from
x2 to x3 . Then
(iii) The operation ∗ has inverses. Let γ be a path from x0 to x1 . Let γ̄(s) :=
γ(1 − s). Then
E
g
p
Y f
X
commute.
20.5 Existence of path lifting 65
Continuing this process (in finitely many steps since [0, 1] is compact), we
have a lift α of γ.
Next consider
F ([s1 , s2 ] × [0, t1 ]) ⊂ U10 .
Let V10 ⊂ E such that
pV : V10 → U10
10
Continue this process for each block [si−1 , si ] × [ti−1 , ti ] to fill out [0, 1] × [0, 1].
We get a lift G of F defined on [0, 1]×[0, 1] such that p◦G = F and G(0, 0) = e0 .
Now assume F is a path homotopy. So say F (0, t) = x0 and F (1, t) = x1 for
all t ∈ [0, 1].
Since p(G(0, t)) = F (0, t) = x0for all t, G( 0 × [0, 1]) ⊂ p−1 (x0 ). On the
Exercise 21.1. Let p : (E, e0 ) → (X, x0 ) be a covering map and let E be path-
connected. Prove the following:
ind(γ) = α(1) ∈ Z,
so the index counts how many times γ makes full revolutions (with sign). Let
γ1 and γ2 be two loops in S 1 based at 1. Since Φ is one-to-one, [γ1 ] = [γ2 ] if
and only if ind(γ1 ) = ind(γ2 ).
Now we claim Φ : π1 (S 1 , 1) → Z = p−1 (1) given by [γ] = ind(γ) is a group
isomorphism.
In other words, we want to show
π1 (Y, y1 )
Note that f can be moved to g by the homotopy F . Consider therefore the map
G : [0, 1] × [0, 1] → Y defined by
Let β1 , β2 , β3 , β4 be the paths along the edges of [0, 1] × [0, 1]: β1 to the right
along the bottom, β2 going up along the left-hand side, β4 going up along the
right-hand side, and β3 going to the right along the top.
So G ◦ β1 = F (γ(s), 0) = f ◦ γ, G ◦ β2 = F (γ(0), t) = F (x0 , t) = α. We have
G ◦ β3 = F (γ(s), 1) = g ◦ γ, and finally G ◦ β4 = F (γ(1), t) = F (x0 , t) = α.
Note that [0, 1] × [0, 1] is convex. Hence any path in the square is homotopic
to any other path with the same start and end points. Hence β3 ' β̄2 ∗ β1 ∗ β4 .
Compose with G, so
G ◦ β3 ' G ◦ (β̄2 ∗ β1 ∗ β4 ),
so in other words
[g ◦ γ] = [ᾱ ∗ (f ◦ γ) ∗ α].
Definition 22.2.2 (Homotopy type). Let X and Y be topological spaces.
(i) A map f : X → Y is a homotopy equivalence if there exists a map
g : Y → X such that f ◦ g ' IdY and g ◦ f ' IdX .
The map g is called a homotopy inverse of f .
(ii) The spaces X and Y are homotopy equivalent (or have the same ho-
motopy type) if there is a homotopy equivalence between X and Y .
Theorem 22.2.3. Let f : (X, x0 ) → (Y, y0 ) be a homotopy equivalence. Then
the induced map f∗ : π1 (X, x0 ) → π1 (Y, y0 ) is an isomorphism.
70 HOMOTOPY TYPE
g◦f 'IdX
Then
(g ◦ f )∗ = g∗ ◦ f∗ : π1 (X, x0 ) → π1 (Y, y0 ) → π1 (X, x1 ),
Remark 22.2.4. If X and Y are homeomorphic, then X and Y have the same
homotopy type.
The converse is not true: for instance a single point x0 is homotopy
equivalent to R, but they are certainly not homeomorphic.
X is contractible if X is homotopy
Definition 22.2.5 (Contractible). A space
equivalent to a single-point space Y = y0 .
F (x, 0) = x0 , x ∈ X,
F (x, 1) = x, x ∈ X,
F (x0 , t) = x0 , 0 ≤ t ≤ 1.
to show ι is an homotopy
Proof. It suffices
x
equivalence. Define the continuous
map r : R2 \ 0 → S 1 by r(x) = kxk . Then r ◦ ι = IdS 1 quite obviously. We
need to also show
r ι
ι ◦ r : R2 \ 0 S1 R2 \ 0
is homotopic to Id . Note
R2 \ 0
x x
ι ◦ r(x) = ι
= .
kxk kxk
Define F : R2 \ 0 × [0, 1] → R2 \ 0 by
x
F (x, t) = t + (1 − t)x.
kxk
x
Then F (x, 0) = x = Id 2 (x) and F (x, 1) = kxk = ι ◦ r(x), so ι ◦ r '
R \ 0
Id .
R2 \ 0
where
2
t(y) = .
1 + kyk2
Remark 23.2.3. Similarly, S n \ q is homeomorphic to S n \ p by the re-
flection map (x1 , . . . , xn , xn+1 ) 7→ (x1 , . . . , xn , −xn+1 ). Hence S n \ q is also
homeomorphic Rn .
Step 2: Let U = S n \ p and V = S n \ q . Then
Sn = U ∪ V
and
U ∩ V = S n \ p, q .
We won’t explore higher homotopic theory here, so we omit the proof. But
it does tell us all S n are not homeomorphic.
It does raise a question though: what is πn (S m ) for n > m? In general, we
don’t know, there are only some partial results (mostly by J. P. Serre, who in
their PhD thesis studied this problem and won the Fields medal).
so Φ is onto.
Second, one-to-one. Suppose [γ] ∈ ker Φ. Then Φ([γ]) = ([p ◦ γ], [q ◦ γ]) = 0,
so p◦γ ' x0 and q◦γ ' y0 . Let G and H be the corresponding path homotopies.
Define F : [0, 1] × [0, 1] → X × Y by
Lecture 24 Retraction
24.1 Fundamental group of projective space
Definition 24.1.1 (Real projective plane). The real projective plane RP 2 =
S 2 /∼ where x ∼ −x in S 2 . That is, antipodal points on the sphere are identified.
This is sort of hard to imagine, because the real projective plane RP 2 can
not be embedded into R3 .
Remark 24.1.2. Since p : S 2 → S 2 /∼ and S 2 is compact, RP 2 is compact.
The quotient map p is a covering map: for [x] ∈ RP 2 , take a small neigh-
bourhood U , small enough that the pullback around x and −x don’t meet.
Then
h∗ = k∗ ◦ j∗ : π1 (S 1 ) → π1 (B 2 ) → π1 (X),
76 BROUWER FIXED-POINT THEOREM
j r
S1 R2 \ 0 S1.
Id
F (x, t) = tx + (1 − t)w(x),
so the sort of convex, straight-line homotopy between w(x) and j(x) = x (think-
ing of j(x) as the outward unit normal of S 1 ) since if t = 0, F (x, 0) = w(x). If
t = 1, F (x, 1) = x = j(x). To make sure this is actually a homotopy, we must
make sure F (x, t) 6= 0 for all x and t.
Suppose F (x, t) = 0 for some 0 < t < 1. Then tx + (1 − t)w(x) = 0, or
t
w(x) = − x.
1−t
But x is radius of the unit circle, so this w(x) points directly inward at x, which
is a contradiction.
For w(x) pointing directly outward at some point x ∈ S 1 , we can consider
the vector field (x, −v(x)), where the − switches directly inward for directly
outward.
Proof. Suppose f (x) 6= x for all x ∈ B 2 . Let v(x) := f (x) − x. Then (x, v(x))
is a nonvanishing continuous vector field on B 2 .
By Theorem 24.2.7 there exists an x ∈ S 1 such that v(x) = f (x) − x = kx
points directly outward, so k > 0.
Hence f (x) = (k + 1)x ∈ B 2 , and f : B 2 → B 2 , but
which is a contradiction.
S1 k
S1.
k ◦ q = q ◦ h,
since (h ◦ f˜) is a path from s0 to −s0 too, so mapped through q it’s a loop
covering the whole circle. Thus k∗ is nontrivial.
Step 3: We claim h∗ is nontrivial. Hence by Lemma 24.2.5, h 6' 0.
We have k∗ : π1 (S 1 ) = Z → π1 (S 1 ) = Z. The kernel ker(k∗ ) is a subgroup
of (Z, +). If the kernel is nonzero, ker(k∗ ) = mZ since Z is cyclic. Thus
Z/mZ ∼ = Im(k∗ ). But Im(k∗) is a subgroup of Z, and Z/mZ is finite, but the
only finite subgroup of Z is 0 , contradicting ker(k∗ ) 6= 0.
Hence ker(k∗ ) = 0, so k∗ is one-to-one.
Note q∗ is also one-to-one; q∗ : π1 (S 1 ) → π1 (S 1 ) is the map m 7→ 2m.
Thus
h∗
π1 (S 1 ) π1 (S 1 )
q∗ q∗
h∗
π1 (S 1 ) π1 (S 1 )
f (x) − f (−x)
g(x) = ,
kf (x) − f (−x)k
which is defined since the denominator is never zero. Then g is continuous, and
g(−x) = −g(x) for all x ∈ S 2 .
Hence g : S 2 → S 1 is continuous and antipode-preserving, contradiction The-
orem 26.1.1.
Date: December 1st, 2020.
80 THE BORSUK–ULAM THEOREM
26.2 Applications
A question: Given a bounded region A ⊂ R2 , can you find a line L that bisects
A?
The answer is yes: consider horizontal slices of the set at y = c, and let f (c)
denote the area of the part of A below y = c.
Then f (c) is a continuous function that varies from 0 to the area of A, so by
the intermediate value theorem, it must attain half the area of A for some c.
Now a trickier question: Given two bounded regions A and B in R2 , can you
find a line L that bisects both A and B simultaneously?
The answer is still yes, but it is quite a bit less trivial.
Remark 27.1.4. There are two difficulties in generalising this for n ≥ 3. First,
there is no retraction of B n to S n−1 (for which, as should now appear natural,
we need higher homotopy or homology).
Second, we need the more general form of the Jordan separation theorem,
saying that S n−1 separates S n .
Given these two, the proof of Invariance of domain generalises nicely.
84 INDEX
Index
A norm . . . . . . . . . . . . . . . . . . . . . . 15
antipode . . . . . . . . . . . . . . . . . . . . . . . 78 even cover . . . . . . . . . . . . . . . . . . . . . 64
antipode-preserving . . . . . . . . . . . . 78
F
B fibre . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
Baire space . . . . . . . . . . . . . . . . . . . . 50 first-countable . . . . . . . . . . . . . . . . . 34
ball . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 fixed-point . . . . . . . . . . . . . . . . . . . . . 77
Banach fixed-point theorem . . . . 32 fundamental group . . . . . . . . . . . . . 63
basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
Bolzano–Weierstrass theorem . . 28, G
30 generated topology . . . . . . . . . . . . . . 5
bounded set . . . . . . . . . . . . . . . . . . . . 14
box topology . . . . . . . . . . . . . . . . . . . 12 H
Hausdorff space . . . . . . . . . . . . . . . . . 7
C homeomorphic spaces . . . . . . . . . . 10
Cauchy sequence . . . . . . . . . . . . . . . 49 homeomorphism . . . . . . . . . . . . . . . 10
closed set . . . . . . . . . . . . . . . . . . . . . . . 2 homogeneous space . . . . . . . . . . . . 59
closure . . . . . . . . . . . . . . . . . . . . . . . . . . 6 homotopic . . . . . . . . . . . . . . . . . . . . . 61
cofinite topology . . . . . . . . . . . . . . . . 2 homotopy equivalence . . . . . . . . . . 69
compact . . . . . . . . . . . . . . . . 17, 22, 28 homotopy equivalent . . . . . . . . . . . 69
compactification . . . . . . . . . . . . . . . 44 homotopy group . . . . . . . . . . . . . . . 73
induced . . . . . . . . . . . . . . . . . . . 45 homotopy inverse . . . . . . . . . . . . . . 69
complete space . . . . . . . . . . . . . . . . . 49 homotopy type . . . . . . . . . . . . . . . . . 69
completely regular . . . . . . . . . . . . . 44
connected I
component . . . . . . . . . . . . . . . . 20 indiscrete topology . . . . . . . . . . . . . . 1
space . . . . . . . . . . . . . . . . . . . . . . 18 interior . . . . . . . . . . . . . . . . . . . . . . . . . .6
subspace . . . . . . . . . . . . . . . . . . 19 isolated point . . . . . . . . . . . . . . . . . . . 6
continuity . . . . . . . . . . . . . . . . . . . . . . . 1
continuous function . . . . . . . . . . . . . 9 L
at a point . . . . . . . . . . . . . . . . . . 9 Lebesgue number lemma . . . . . . . 30
contractible . . . . . . . . . . . . . . . . . . . . 70 lift . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
convergence . . . . . . . . . . . . . . . . . . . . . 6 limit point . . . . . . . . . . . . . . . . . . .6, 28
convex space . . . . . . . . . . . . . . . . . . . 62 limit point compact . . . . . . . . . . . . 28
coordinate functions . . . . . . . . . . . 11 locally compact . . . . . . . . . . . . . . . . 33
covering map . . . . . . . . . . . . . . . . . . 64 at a point . . . . . . . . . . . . . . . . . 33
covering space . . . . . . . . . . . . . . . . . 64 locally finite . . . . . . . . . . . . . . . . . . . 10
locally path-connected . . . . . . . . . 22
D loop . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
dense . . . . . . . . . . . . . . . . . . . . . . . . . . 35 lower limit topology . . . . . . . . . . . . . 4
diameter . . . . . . . . . . . . . . . . . . . . . . . 30
disconnected space . . . . . . . . . . . . . 18 M
discrete topology . . . . . . . . . . . . . . . . 1 maximal element . . . . . . . . . . . . . . . 26
dyadic rational number . . . . . . . . .38 metric . . . . . . . . . . . . . . . . . . . . . . . . . 13
metric space . . . . . . . . . . . . . . . . . . . 14
E metric topology . . . . . . . . . . . . . . . . 13
equivalent metrisability . . . . . . . . . . . . . . . . . . . 10
compactification . . . . . . . . . . . 45 metrisable . . . . . . . . . . . . . . . . . . . . . 14
INDEX 85
N T
neighbourhood . . . . . . . . . . . . . . . . . . 1 T2 -space . . . . . . . . . . . . . . . . . . . . . . . 35
norm . . . . . . . . . . . . . . . . . . . . . . . . . . .14 T2 -space . . . . . . . . . . . . . . . . . . . . . 7, 36
normal space . . . . . . . . . . . . . . . . . . . 36 T3 -space . . . . . . . . . . . . . . . . . . . . . . . 36
normed topology . . . . . . . . . . . . . . . 15 T3.5 -space . . . . . . . . . . . . . . . . . . . . . . 44
nullhomotopic . . . . . . . . . . . . . . . . . . 75 T4 -space . . . . . . . . . . . . . . . . . . . . . . . 36
topological group . . . . . . . . . . . . . . 57
O topological property . . . . . . . . 10, 14
one-point compactification . . . . . 33 topological space . . . . . . . . . . . . . . . . 1
open map . . . . . . . . . . . . . . . . . . . . . . 24 topologist’s sine curve . . . . . . . . . . 22
open set . . . . . . . . . . . . . . . . . . . . . . . . .1 topology . . . . . . . . . . . . . . . . . . . . . . . . 1
orbit . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 finer . . . . . . . . . . . . . . . . . . . . . . . . 3
orbit space . . . . . . . . . . . . . . . . . . . . . 59 strictly finer . . . . . . . . . . . . . . . . 3
totally bounded . . . . . . . . . . . . 31, 49
P totally disconnected . . . . . . . . 12, 19
partially ordered set . . . . . . . . . . . .25 totally ordered set . . . . . . . . . . . . . 26
path . . . . . . . . . . . . . . . . . . . . . . . 21, 61 transitive action . . . . . . . . . . . . . . . .59
path homotopic . . . . . . . . . . . . . . . . 61 triangle inequality . . . . . . . . . . 13, 14
path-connected . . . . . . . . . . . . . . . . .21 trivial topology . . . . . . . . . . . . . . . . . 1
product topology . . . . . . . . . . . . 7, 12 tubular neighbourhood theorem 81
Q U
quotient space . . . . . . . . . . . . . . . . . 55 uniform convergence . . . . . . . . . . . 18
quotient topology . . . . . . . . . . . . . . 55 uniformly continuous . . . . . . . . . . . 32
universal covering space . . . . . . . . 67
R upper bound . . . . . . . . . . . . . . . . . . . 26
real projective plane . . . . . . . . . . . 74
real projective space . . . . . . . . . . . 74 V
regular space . . . . . . . . . . . . . . . . . . . 36 vector field
relative topology continuous . . . . . . . . . . . . . . . . .76
see subspace topology . . . . . . .5 nonvanishing . . . . . . . . . . . . . . 76
retract . . . . . . . . . . . . . . . . . . . . . . . . . 74
retraction . . . . . . . . . . . . . . . . . . . . . . 74
Riemann sphere . . . . . . . . . . . . . . . . 34
S
second-countable . . . . . . . . . . . . . . . 34
separable . . . . . . . . . . . . . . . . . . . . . . 35
sequentially compact . . . . . . . . . . . 28
simply connected . . . . . . . . . . . . . . .63
semilocally . . . . . . . . . . . . . . . . 67
special orthogonal group . . . . . . . 60
stabiliser . . . . . . . . . . . . . . . . . . . . . . . 60
standard bounded metric . . . . . . . 42
standard topology . . . . . . . . . . . . . . . 4
Stone–Čech compactification 46, 48
subbasis . . . . . . . . . . . . . . . . . . . . . . . . . 5
subspace . . . . . . . . . . . . . . . . . . . . . . . . 5
subspace topology . . . . . . . . . . . . . . . 5
sup-norm . . . . . . . . . . . . . . . . . . . . . . 16