You are on page 1of 75

LECTURE NOTES ON GENERAL TOPOLOGY

BIT, SPRING 2022

DAVID G.L. WANG

Contents

1. Introduction 3

1.1. Who cares topology? 3

1.2. Geometry v.s. topology 10

1.3. The origin of topology 11

1.4. Topological equivalence 14

2. Topological Spaces 16

2.1. Topological structures 16

2.2. Subspace topology 22

2.3. Point position with respect to a set 24

2.4. Bases of a topology 30

2.5. Hierarchy of topologies 33

2.6. Metrics & the metric topology 33

3. Continuous Maps & Homeomorphisms 43

3.1. Continuous maps 43

3.2. Covers 49

3.3. Homeomorphisms 52

Date: March 17, 2022.


1
4. Connectedness 62

4.1. Connected spaces 62

Appendix A. Some elementary inequalities 72

Appendix B. Subadditivity and superadditivity 74


3

1. Introduction

1.1. Who cares topology? We start from 4 examples.

Question 1.1. We have a shell of the doughnut, with a hole on it. The inside is red and
the outside is blue. Can you move the shape around stretching and bending without
tearing, so that it is with the red outside and blue inside?

Figure 1. Screenshot from a video of the PBS Digital Studios.


4 D.G.L. WANG

Figure 2. Answer to Question 1.1: Yes. Screenshot from a video of the


PBS Digital Studios.
5

Question 1.2. You have a picture frame with a loop of string fixed on the back, and
two nails. It looks either one of the nails is sufficient to hang the frame, because
removing any one of the nails does not cause the frame fall down.

Can you create a way of hanging the frame by smartly winding the loop of string
onto the nails so that pulling out any one of the nails from the wall leads the painting
falling down?

Figure 3. Screenshot from a video of the PBS Digital Studios.


6 D.G.L. WANG

Figure 4. Answer to Question 1.2: Yes. Screenshot from a video of the


PBS Digital Studios.
7

The Nobel Prize in Physics 2016 was awarded with one half to David J. Thouless,
and the other half to F. Duncan M. Haldane and J. Michael Kosterlitz “for theoretical
discoveries of topological phase transitions and topological phases of matter”;
see Fig. 5.

Topology was the key to the Nobel Laureates’ discoveries, and it explains why
electronical conductivity inside thin layers changes in integer steps. — Popular Science
Background of the Nobel Prize in Physics 2016, Page 4(5)

Figure 5. Illustration for the classfication of surfaces.


8 D.G.L. WANG

Topology can be used to solve mathematical problems that is seemingly unrelated


with topology. In 1911, Toeplitz posed the Square Peg Problem, which asks whether
every continuous Jordan curve in the Euclidean plane inscribes a square. A weaker
problem that asks whether every continuous Jordan curve inscribes a rectangle can be
solved easily using topology.

The key idea is that every unordered pair of points on a simple closed curve corresponds
bijectively to a point on a Möbius strip, and that the Möbius strip is not embeddable
into the plane.

Figure 6. August Ferdinand Möbius (1790–1868) was a German mathe-


matician and theoretical astronomer. The right part is a Möbius strip.

Theorem 1.3 (Greene and Lobb, 2021). For every smooth Jordan curve γ and rectan-
gle R in the Euclidean plane, there exists a rectangle similar to R whose vertices lie
on γ.

See “J.E. Greene and A. Lobb, The rectangular peg problem, Ann. Math. (2) 194
(2021), 509–517”. The proof relies on the theorem of Shevchishin and Nemirovski that
the Klein bottle does not admit a smooth Lagrangian embedding in C2 .
9

Something people noticed

Topology, over most of its history, has NOT generally been applied outside of
mathematics (with a few interesting exceptions).

WHY?

• TOO abstract? The ancient mathematicians could not even convince of the subject.

Figure 7. Screenshot from a video of Robert MacPherson’s talk in


Institute for Advanced Study.

• It is qualitative, not quantitative? People think of science as a quantitative endeavour.


10 D.G.L. WANG

1.2. Geometry v.s. topology. Below are some views from Robert MacPherson, a
plenary addresser at the ICM in Warsaw in 1983.

• Geometry (from ancient Greek): geo=earth, metry=measurement.


Topology (from Greek): τ óπoσ=place/position, λóγoσ=study/discourse.
• Topology is “geometry” without measurement.
It is qualitative (as opposed to quantitative) “geometry”.
• Geometry: The point M is the midpoint of the straight line segment L connecting
A to B.
Topology: The point M lies on the curve L connecting A to B.
• Geometry calls its objects configurations (circles, triangles, etc.)
Topology calls its objects spaces.
11

1.3. The origin of topology. Here are three stories about the origin of topology.

1.3.1. The seven bridges of Königsberg. The problem was to devise a walk through the
city that would cross each of those bridges once and only once; see Fig. 8.

Figure 8. Map of Königsberg in Euler’s time showing the actual layout


of the seven bridges, highlighting the river Pregel and the bridges. Stolen
from Wiki.

The negative resolution by Leonhard Paul Euler (1707–1783) in 1736 laid the founda-
tions of graph theory and prefigured the idea of topology. The difficulty Euler faced
was the development of a suitable technique of analysis, and of subsequent tests that
established this assertion with mathematical rigor.

Euler was a Swiss mathematician, physicist, astronomer, logician and engineer who
made important and influential discoveries in many branches of mathematics like
infinitesimal calculus and graph theory, while also making pioneering contributions to
several branches such as topology and analytic number theory.

1.3.2. The four colour theorem. The four colour theorem states that given any separation
of a plane into contiguous regions, producing a figure called a map, no more than four
colours are required to colour the regions of the map so that no two adjacent regions
have the same color.

The four color theorem was proved in 1976 by Kenneth Appel and Wolfgang Haken.
It was the first major theorem to be proved using a computer.

1.3.3. Euler characteristic. χ = v − e + f ; see Fig. 9.


12 D.G.L. WANG

Figure 9. Leonhard Euler (1707–1783) was a Swiss mathematician,


physicist, astronomer, geographer, logician and engineer who founded
the studies of graph theory and topology and made pioneering and
influential discoveries in many other branches of mathematics such as
analytic number theory, complex analysis, and infinitesimal calculus. He
introduced much of modern mathematical terminology and notation,
including the notion of a mathematical function. He is also known for his
work in mechanics, fluid dynamics, optics, astronomy and music theory.
The right part is a stamp of the former German Democratic Republic
honouring Euler on the 200th anniversary of his death. Across the centre
it shows his polyhedral formula.

Theorem 1.4 (Euler’s polyhedral formula). The number of vertices v, faces f , and
edges e in a convex 3-dimensional polyhedron satisfy v − e + f = 2.

David Eppstein collected 20 proofs of Theorem 1.4.


13

In 3-dimensional space, a Platonic solid is a regular, convex polyhedron. The Euler


characteristic of every Platonic solid is 2. In fact, there are only 5 (why?) Platonic solids:
the tetrahedron, the cube, the octahedron, the dodecahedron, and the icosahedron; see
Fig. 10.

Figure 10. The Plotonic solids. Stolen from Wiki.


14 D.G.L. WANG

1.4. Topological equivalence.



: thicken, stretch, bend, twist, . . .;
× : identify, tear, . . ..

See Fig. 11.

Figure 11. Cups with 0,1,2,3,4,8 handles.

Theorem 1.5 (Classification theorem). Any closed surface is homeomorphic to S 2 with


either a finite number of handles added, or a finite number of Möbius strips added. No
two of these surfaces are homeomorphic.

Definition 1.6. The S 2 with n handles added is called an orientable surface of


genus n. Non-orientable surfaces can be defined analogously.

Historical notes. The classification of surfaces was initiated and carried through in
the orientable case by Möbius in a paper which he submitted for consideration for the
Grand Prix de Mathématiques of the Paris Academy of Sciences. He was 71 at the
time. The jury did not consider any of the manuscripts received as being worthy of the
prize, and Möbius’ work finally appeared as just another mathematical paper.
15

We are now only a short step in mathematical intuition, away from one of the most
basic and central results of topology. In terms of careful proof we have an extremely
long walk ahead.

Theorem 1.7. Topological equivalent polyhedra have the same Euler characteristic.

• The starting point for modern topology.


• Search for unchanged properties of spaces under topological equivalence.
• χ = 2 belongs to S 2 , rather than to particular polyhedra → define χ for S 2 .
• Theorem 1.7: different calculations, same answer.
16 D.G.L. WANG

2. Topological Spaces

We start with a modern version of the concept topology.

2.1. Topological structures.

Definition 2.1. Let X be a set and Ω ⊆ 2X .

• A (topological) space is a pair (X, Ω), where the collection Ω, called a topology or
topological structure on X, satisfies the axioms
(i) ∅ ∈ Ω and X ∈ Ω;
(ii) the union of any members of Ω lies in Ω;
(iii) the intersection of any two members of Ω lies in Ω.
• A point in X: p ∈ X.
• An open set in (X, Ω): a member O ∈ Ω.
A closed set in (X, Ω): a subset A ⊆ X s.t. Ac ∈ Ω.
A clopen set in (X, Ω): an open set O ∈ Ω which is also closed.

Remark 2.2. Being closed is not the negation of being open! A set might be
• open but not closed, or
• closed but not open, or
• clopen, or
• neither closed nor open.

Remark 2.3. Why do we use the letter O and the letter Ω?


• Open in English
• Ouvert in French
• Otkrytyj in Russian
• Offen in German
• Öppen in Swedish
• Otvoren in Croatian
• Otevřeno in Czech
• Open in Dutch
17

Here are some topological spaces that we will meet frequently in this note.

Space 1. An indiscrete or trivial space: (X, {∅, X}).

Space 2. A discrete space: (X, 2X ).

Proposition 2.4. A space is discrete if and only if every singleton is open.

Proof. The forward direction is clear since every singleton is a particular subset. For
the backward direction, since every subset in a space is the union of the singletons
consisting of its members, it is open. □

Space 3. A particular point space (X, Ω):


Ω = {∅, X} ∪ {S ⊆ X : p ∈ S},
where p is a particular point in X.

Space 4. An excluded point space (X, Ω):


Ω = {∅, X} ∪ {S ⊆ X : p ̸∈ S},
where p is a particular point in X.

Space 5. The real line (R, ΩR ):


ΩR = {unions of open intervals}
is the canonical or standard topology on R.

Space 6. The cofinite space (R, ΩT1 ):


ΩT1 = {∅, X} ∪ {complements of finite subsets of R}
is the cofinite topology or finite-complement topology or T1 -topology.

Space 7. The arrow (X, Ω):


X = {x ∈ R : x ≥ 0} and Ω = {∅, X} ∪ {(a, ∞) : a ≥ 0}.
18 D.G.L. WANG

Space 8. The Sierpiński space (X, Ω):


X = {a, b} and Ω = { ∅, {a}, {a, b}}.

Figure 12. Waclaw Franciszek Sierpiński (1882–1969) was a Polish


mathematician. He was known for outstanding contributions to set
theory, number theory, theory of functions and topology.

Question 2.5. Are the following pairs (X, Ω) topological spaces?

(1) X = R2 and Ω = {∅, X} ∪ {open disks centred at the origin};


Answer. Yes. □

(2) X = R and Ω = {∅, X} ∪ {infinite subsets of R}.


Answer. No. □
19

Example 2.6. The set {0} ∪ {1/n : n ∈ Z+ } is closed in the real line. □

“Think geometrically, prove algebraically.” — John Tate

Figure 13. John Tate (1925–) is an American mathematician, distin-


guished for many fundamental contributions in algebraic number theory,
arithmetic geometry and related areas in algebraic geometry. He is pro-
fessor emeritus at Harvard Univ. He was awarded the Abel Prize in 2010.

Homework 2.1 (Deadline: 11.59 pm, Mar 3rd). Does there exist a topology which is
both a particular point topology and an excluded point topology?

Homework 2.2 (Deadline: 11.59 pm, Mar 3rd). Find a topological space (X, Ω) with
a set A ⊂ X satisfying the following 2 properties:

a) A is the intersection of an infinite number of open sets, but not open;


b) A is the union of an infinite number of closed sets, but not closed.
20 D.G.L. WANG

Example 2.7. The Cantor ternary set K is the number set created by iteratively
deleting the open middle third from a set of line segments, i.e.,
( )
X ak 
K= k
: a k ∈ {0, 2} = 0.a 1 a 2 · · · : a i ∈ {0, 2, 3, 5, 6, 8, 9} ⊂ [0, 1].
k≥1
3

See Fig. 14. The set K was discovered by Henry John Stephen Smith in 1874, and

Figure 14. The first figure is Henry John Stephen Smith (1826–1883), a
mathematician remembered for his work in elementary divisors, quadratic
forms, and Smith-Minkowski-Siegel mass formula in number theory. The
second figure is Georg Cantor (1845–1918), a German mathematician
who invented set theory. The third figure illustrates the Cantor ternary
set.

introduced by Georg Cantor in 1883. It has a number of remarkable and deep properties.
For instance, it is closed in R.

Definition 2.8. Given (X, Ω). Let p ∈ X. A neighbourhood of p is a subset U ⊆ X


s.t.
∃ O ∈ Ω s.t. p ∈ O ⊆ U.

Remark 2.9. In literature the letter U is used to indicate a neighbourhood since it is


the first letter of the German word “umgebung” which means neighbourhood.

Remark 2.10. We are following the Nicolas Bourbaki group and define the term “neigh-
bourhood” in the above sense. There is another custom that a neighbourhood of a
point p is an open set containing p. Nicolas Bourbaki is the collective pseudonym
under which a group of (mainly French) 20th-century mathematicians, with the aim
of reformulating mathematics on an extremely abstract and formal but self-contained
basis, wrote a series of books beginning in 1935. With the goal of grounding all of
mathematics on set theory, the group strove for rigour and generality. Their work
21

led to the discovery of several concepts and terminologies still used, and influenced
modern branches of mathematics. While there is no one person named Nicolas Bourbaki,
the Bourbaki group, officially known as the Association des collaborateurs de Nicolas
Bourbaki (Association of Collaborators of Nicolas Bourbaki), has an office at the École
Normale Supérieure in Paris.

Question 2.11. A topology can be defined by assigning open sets; see Definition 2.1.
Can it be defined by assigning closed sets?

Answer. Yes. Here is a list of axioms for assigning closed sets:

(i)’ ∅ and X are closed;


(ii)’ the union of any finite number of closed sets is closed;
(iii)’ the intersection of any collection of closed sets is closed.

Remark 2.12. Given (X, Ω). Since the union of all members in Ω is X, the topology Ω
itself carries enough information to clarify a topological space. However, the topological
space (X, Ω) is often denoted simply by X, because different topological structures in
the same set X are often considered simultaneously rather seldom. Moreover, to exclaim
a set is in general easier than clarifying a topology. As will be seen in Section 2.2,
subspace topology helps the clarification by calling the knowledge of common topologies.
22 D.G.L. WANG

2.2. Subspace topology.

Definition 2.13. Given (X, Ω) and A ⊆ X. The subspace topology (or induced
topology) of X induced by A:

ΩA = {O ∩ A : O ∈ Ω}.

The topological subspace induced by A: (A, ΩA ).

Question 2.14. Answer the following questions.

1) Describe the topological structure of ΩR induced by Z+ .

Answer. The discrete topology. □

2) Describe the topological structure of the arrow induced by Z+ .

Answer. All sets of the form {n ∈ N : n ≥ a} where a ∈ N. □

3) Describe the topological structure of ΩT1 induced by the two-element set {1, 2}.

Answer. The discrete topology. □

4) Describe the topological structure of the arrow induced by {1, 2}.

Answer. {∅, {2}, {1, 2}}. □

Theorem 2.15. Let (X, Ω) be a topological space and A ⊆ X. The subspace topology
of (X, Ω) induced by A can be defined alternatively in terms of closed sets as

S is closed in A ⇐⇒ S = F ∩ A, where F is a closed set in X. □

Theorem 2.16. Let X be a topological space and let A ⊆ M ⊆ X.

1) If A is open in X, then A is open in M .


If A is closed in X, then A is closed in M .
2) If A is open in M , and if M is open in X, then A is open in X.
If A is closed in M , and if M is closed in X, then A is closed in X.

Proof. The openness of A in M follows from the formula A = A ∩ M . Conversely, if


A is open in M , then A = O ∩ M , where O is open in X. As a consequence, this
intersection A = O ∩ M is open in X as long as M is also open in X. The “closed”
version can be shown along the same line with aid of Theorem 2.15. □
23

Remark 2.17. The condition that M is open/closed in X in Theorem 2.16 is necessary.



For instance, the set {x ∈ Q : x > 2} is clopen in Q, but neither closed nor open in R.

Theorem 2.16 is about the openness of A in the set inclusion structure A ⊆ M ⊆ X.


Theorem 2.18 concerns the topology of A in that structure.

Theorem 2.18. Let A ⊆ M ⊆ X. Then the topology of X induced by A coincides with


the topology of (M, ΩM ) induced by A, where ΩM is the subspace topology of X induced
by M . In other words, it is safe to say “the subspace topology of A”.

Proof. {O ∩ A : O ∈ ΩM } = {(O′ ∩ M ) ∩ A : O′ ∈ ΩX } = {O′ ∩ A : O′ ∈ ΩX }. □


24 D.G.L. WANG

2.3. Point position with respect to a set.

Definition 2.19. Given (X, Ω) and A ⊆ X.

• A limit point of A: a point p ∈ X s.t.

U ∩ (A \ {p}) ̸= ∅, ∀ neighbourhood U of p.

An isolated point of A: a point p ∈ A s.t.

∃ a neighbourhood U of p s.t. U ∩ (A \ {p}) = ∅.

• A is perfect, if it is closed and has no isolated points.


• The closure of A: the union of A and its limit points, denoted A, i.e.,

A = {x ∈ X : U ∩ A ̸= ∅, ∀ neighbourhood U of x}.

It is alternatively written as ClA when considered as a set operator.


An adherent point of A: a point in A.
• An interior point of A: a point having a neighbourhood in A.
An exterior point of A: a point having a neighbourhood in Ac .
A boundary point of A: a point s.t. each neighbourhood meets both A and Ac .
• The interior of A: A◦ = {interior points} = ∪{O ∈ Ω : O ⊆ A};
The exterior of A w.r.t. X: (Ac )◦ = {exterior points};
The boundary of A: ∂A = {boundary points} = A \ A◦ .

Remark 2.20. The symbols Cl(A), Int(A), and Ext(A) are used to denote the closure A,
the interior A◦ , and the exterior (Ac )◦ resp., when one emphasizes that they are set
operators. The symbol Bd(A) is sometimes recognized as the boundary of A, but used
uncommonly since the symbol ∂ well plays the role of a set operator.

Example 2.21. ∀ A ⊆ X, we have

A◦ ⊆ A ⊆ A, ExtA ⊆ Ac , and Ext∅ = X.

In R, we have IntQ = ∅ and ExtQ = ∅. For the Cantor set in Example 2.7, we have

K = K, K ◦ = ∅, ExtK = I \ K, and ∂K = K.

Notation 2.22. The unit interval: the interval [0, 1] in R, denoted by I.


25

Puzzle 2.23. The Smith-Volterra-Cantor set K ′ is a set of points on the real line
R; see Fig. 15. It can be obtained by removing certain intervals from I as follows. After
removing the middle 1/4 from I, remove the subintervals of length 1/4n from the
middle of each of the remaining intervals. For instance, at the first and second step the
remaining intervals are
           
3 5 5 7 3 5 25 27
0, ∪ ,1 and 0, ∪ , ∪ , ∪ ,1 .
8 8 32 32 8 8 32 32
Show that the set K ′ is nowhere dense.

Figure 15. The Smith-Volterra-Cantor set.

Theorem 2.24 (Characterization of the closure and interior). The closure of a set is
the smallest closed set containing that set, and the interior of a set is the largest open
set contained in that set.

Proof. The closure of a set is the intersection of all closed sets containing it, and the
interior of a set is the union of all open sets contained in it. □

Proposition 2.25. Given a topological space X.

1) The interior and exterior of a set are open, while the boundary and closure are closed.
2) The whole set X can be decomposed w.r.t. a subset A:
X = (∂A) ⊔ (IntA) ⊔ (ExtA) = (ClA) ⊔ (ExtA),

3) ∂A = A ∩ Ac .
4) ∂(A◦ ) ⊆ ∂A.

Definition 2.26. Let f be an operator.

• f is an involution: f ◦ f = id. is the identity map.


• f is idempotent: f ◦ f = f .

In the operator theory, we denote the set operator of closure, interior, complement
resp. by k, i, and c. For instance, the set operator c is an involution.
26 D.G.L. WANG

Theorem 2.27. Given (X, Ω) and A ⊆ X.

1) Both the set operations interior and closure preserve the inclusion, i.e.,

A⊆X =⇒ A◦ ⊆ X ◦ and A ⊆ X.

2) ci = kc.
3) The operators k, i, and ki = kckc are idempotent.
4) The distributivity of k with ∪, and the distributivity of i with ∩:

A∪B =A∪B and (A ∩ B)◦ = A◦ ∩ B ◦ .

The operators k does not work well with ∩, neither does i with ∪:

A∩B ⊆A∩B and (A ∪ B)◦ ⊇ A◦ ∪ B ◦ .

Proof. We shall show that ki is idempotent.

On the one hand, since the open set IntA is contained in the set Y = Cl ◦ IntA, we
find IntA ⊆ IntY by Theorem 2.24. Then by using the closure relation in (1) we obtain

Y = Cl ◦ IntA ⊆ Cl ◦ IntY.

On the other hand, since the set IntY is contained in the closed set Y , applying
Theorem 2.24 we find
Cl ◦ IntY ⊆ Y.

Combining the two relations above, we obtain Y = Cl ◦ IntY , as desired. □

Remark 2.28. It is clear that the operation of taking exterior does not preserve the
inclusion, and that it is not idempotent.

From Theorem 2.27, one may see that the set operations of taking closure and taking
interior share some properties. However, there are some properties that they do not
share.

Theorem 2.29. Let X be a topological space and A ⊆ M ⊆ X. Then

ClM (A) = ClX (A) ∩ M,

but IntM (A) ̸= IntX (A) ∩ M in general.


27

Proof. By Theorem 2.15, the formula for closure holds since


\ \ \
ClM (A) = F = (C ∩ M ) = M ∩ C = M ∩ ClX (A),
A⊆F A⊆C A⊆C
F is closed in M C is closed in X C is closed in X

The other formula holds false for instance X = R2 , M = A = R. □

Puzzle 2.30 (Kuratowski’s closure-complement problem). How many pairwise distinct


sets can one obtain from of a given subset of a topological space by using the set
operators k and c?

Figure 16. Kazimierz Kuratowski (1896–1980) was a Polish mathemati-


cian and logician. He was one of the leading representatives of the Warsaw
School of Mathematics.
Theorem 2.31 (Kuratowski). Let X be a topological space and A ⊂ X. Then at most
14 distinct subsets of X can be formed from E by taking closures and complements.

An example in the real line: (0, 1) ∪ (1, 2) ∪ {3} ∪ ([4, 5] ∩ Q).

Puzzle 2.32. Recall that we can define a topology either in terms of open sets or in
terms of closed sets. Can we define a topology with the aid of the closure operation, or
the interior operation?

Answer. Let X be a set. Let Cl∗ be a transformation on the power set 2X s.t.

(i) Cl∗ ∅ = ∅;
(ii) A ⊆ Cl∗ A;
(iii) distributive with the union operation: Cl∗ (A ∪ B) = Cl∗ A ∪ Cl∗ B;
(iv) idempotent: Cl∗ Cl∗ A = Cl∗ A.
28 D.G.L. WANG

Then the set Ω = {O ⊆ X : Cl∗ (Oc ) = Oc } is a topology on X. Moreover, the set Cl∗ A
is the closure of a set A in the topological space (X, Ω). □
Definition 2.33. Given (X, Ω) and A, B ⊆ X.

• A is dense in B: B ⊆ A.
• A is everywhere dense: A = X, i.e., ExtA = ∅.
• A is nowhere dense: ExtA is everywhere dense, i.e., ExtExtA = ∅.
Remark 2.34. Concerning topics on subset density, an often helpful fact is Theorem 2.24;
see the proofs of Theorem 2.36, Corollary 2.37, Corollary 2.38.
Example 2.35. The whole set is everywhere dense. Continuing Example 2.21, the
set Q is everywhere dense in R, and the Cantor set is nowhere dense in I.
Theorem 2.36 (Characterization for everywhere density). A set is everywhere dense
⇐⇒ it meets every nonempty open set ⇐⇒ it meets every neighbourhood.

Proof. The second equivalent is clear. We show the first. Let (X, Ω) be a topological
space with A ⊆ X.

⇒. Let O ∈ Ω s.t. A ∩ O = ∅. Then A is a subset of the closed set Oc . It follows that


X = A ⊆ Oc . Hence O = ∅.
⇐. Let F ⊂ X be a closed set containing A. From premise, we have A ∩ F c ̸= ∅, that is,
A ̸⊆ F . Therefore, the smallest closed set containing A is X. By Theorem 2.24, we
have A = X. Hence A is everywhere dense.

This completes the proof. □


Corollary 2.37. If A is everywhere dense and O is open, then O ⊆ A ∩ O.

Proof. If not, then ∃ x ∈ O with a neighbourhood U s.t. U ∩ (A ∩ O) = ∅. W.l.o.g.,


we can suppose that U is open. Then the everywhere dense set A does not meet the
open set U ∩ O. By Theorem 2.36, we infer that U ∩ O = ∅, contradicting the fact that
x ∈ U ∩ O. □
Corollary 2.38. Let X be a topological space.

(1) X is indiscrete ⇐⇒ only the empty set ∅ is not everywhere dense.

Proof. Let A ⊆ X s.t. A ̸∈ {∅, X}.


29

⇒. If X is indiscrete, then there is only one nonempty open set, that is, the whole
set X. Certainly A meets X. By Theorem 2.36, the set A is everywhere dense.
⇐. Since A = X ̸= A, no A is closed. Hence X is indiscrete.

(2) X is discrete ⇐⇒ only the whole set X is everywhere dense.

Proof. Let A ⊆ X s.t. A ̸∈ {∅, X}.


⇒. If X is discrete, then A is closed, and A = A ̸= X.
⇐. By Theorem 2.36, the set A does not meet some nonempty open set. Taking A to
be the complement of each singleton, we find that every singleton is open. Hence
X is discrete.

(3) A set S is everywhere dense in the arrow ⇐⇒ sup S = ∞.

Proof.⇐. If sup S = ∞, then (s, ∞) ∩ S ̸= ∅, ∀ s ≥ 0. It is clear that [ 0, ∞) ∩ S ̸= ∅.


Thus S meets every nonempty open set, and is everywhere dense by Theorem 2.36.
⇒. If sup S ̸= ∞, then ∃ M > 0 s.t. s < M , ∀ s ∈ S. It follows that S ∩ (M, ∞) = ∅.
By Theorem 2.36, we infer that S is not everywhere dense.

(4) A set S is everywhere dense in the cofinite space ⇐⇒ S is infinite. □

Theorem 2.39 (Characterization for nowhere denseness). A set A is nowhere dense in


a topological space X ⇐⇒
each open set in X contains an open set that is contained in Ac .

Proof. By Theorem 2.36, we deduce that A is nowhere dense


⇐⇒ ExtA is everywhere dense
⇐⇒ ExtA meets every neighbourhood
⇐⇒ each neighbourhood contains an exterior point of A
⇐⇒ each neighbourhood contains a neighbourhood that is contained in Ac
⇐⇒ each open set contains an open set that is contained in Ac .
This completes the proof. □
30 D.G.L. WANG

Theorem 2.40. The real line is not the union of a countable number of nowhere dense
sets.

Proof. Assume to the contrary that R = ∪n∈N Yn , where Yn is nowhere dense in R. Since
Y0 is nowhere dense, by Theorem 2.39, ∃ [a0 , b0 ] ⊆ Y0c . Since Y1 is nowhere dense, by
Theorem 2.39, ∃ [a1 , b1 ] ⊆ [a0 , b0 ] ⊆ Y1c . Continuing in this way, one obtains a nested
sequence of closed intervals [an , bn ] which is contained in Ync . From calculus, we know
that the set A = ∩n∈N [an , bn ] ̸= ∅. But
!c
\ [
A⊆ Ync = Yn = Rc = ∅,
n∈N n∈N

a contradiction. □

2.4. Bases of a topology.

Definition 2.41. Given (X, Ω).

• A base for Ω: a collection β of open sets s.t.


any open set of X is the union of some members of β.

• Ω is generated by β: β is a base for a topology Ω; denoted Ω = ⟨β⟩.

Remark 2.42. In contrast to a basis of a vector space in linear algebra, a base need
not to be maximal, e.g., any open set can be safely added to a base. Moreover,
a topological space may have disjoint bases of distinct sizes, e.g., the standard
topology ΩR of the real line has a base of all open intervals with rational ends, and
another base of all open intervals with irrational ends.

Question 2.43. Here are some questions about bases.

(1) Describe a smallest base for an indiscrete space.

Proof. The whole set. □

(2) Describe a smallest base for a discrete space. Is it unique?

Answer. The collection of all singletons. □

(3) Describe a base for the arrow. Is it unique?

Answer. {(r, ∞) : r ∈ Q+ } ∪ { [0, ∞)}. No. □


31

Here is a criterion for a collection to be a base of a given topology.

Theorem 2.44. Given (X, Ω) and β ⊆ Ω. Then β is a base for Ω

⇐⇒ for any pair (x, O) where x ∈ O ∈ Ω, ∃ B ∈ β w. x ∈ B ⊆ O.

Proof. Let O ∈ Ω. ⇒: Since β is base, ∃ Γ ⊆ β s.t. O = ∪B∈Γ B.


⇐: ∀ x ∈ O, let Bx ∈ β such that x ∈ Bx ⊆ O. Then O = ∪x∈O Bx . □

Theorem 2.45. The real line has no minimal bases. In other words, for any base β of
the real line R, ∃ B ∈ β s.t. β \ {B} is a base of R.

Proof. Assume that R has a minimal base β. Let A ∈ β. By Theorem 2.44, there
exists a pair (xA , OA ) where xA ∈ OA ⊆ ΩR s.t. the set A is the unique M ∈ β s.t.
xA ∈ M ⊆ OA .

Let IA be the maximal open interval in A that contains xA . Then IA ⊆ A. If IA ̸= A,


then IA is generated by β \ {A}. By Theorem 2.44, there exists B ∈ β \ {A} s.t.
xA ∈ B ⊆ IA , contradicting contradicting the uniqueness of A. This proves IA = A. It
follows that A is an open inteval. Let A = (u, v). We proceed according to the values
of u and v.

• If u, v ∈ R, then the interval Jn = (u + (v − u)/n, v) is generated by β \ {A} for


every integer n ≥ 1. Therefore, A = ∪n≥1 Jn is generated by β \ {A}. This proves
that β \ {A} is a base of R.
• If u ∈ R and v = ∞, then by considering the intervals Jn = (u + 1/n, v) instead, one
may show that β \ {A} is a base of R.
• If u = −∞ and v ∈ R, the same idea works.
• If A = (−∞, ∞), then by considering the intervals Jn = (−n, n) instead, one may
obtain the desired result.

This completes the proof. □

Next is a criterion for a collection to be a base of a certain topology.


32 D.G.L. WANG

Theorem 2.46. Given X and β ⊆ 2X . Then

β is a base for a certain topology on X



X is the union of all members of β,
⇐⇒
the intersection of any two members of β is the union of some sets in β.

Proof. ⇒: Since any member of β is open for the topology generated by β, the
intersection of any two members is open and hence can be generated by the base β. ⇐:
Define Ω ⊆ 2X to be the collection of unions of members of β. By using distributive
law for the unions and intersections, one may verify that Ω is a topology on X. □

In contrast to Euclid’s classical proof of the infinitude of primes recorded in the


Elements, Hillel (Harry) Furstenberg published a topological proof in 1955 (H. Fursten-
berg, On the infinitude of primes, Amer. Math. Monthly 62 (5) 353), when he was an
undergraduate student at Yeshiva Univ.; see Fig. 17.

Figure 17. Hillel (Harry) Furstenberg (1935 — ) is an American-Israeli


mathematician, an Abel Prize Laureates 2020, a member of the Israel
Academy of Sciences and Humanities and U.S. National Academy of
Sciences and a laureate of the Wolf Prize in Mathematics. Stolen from
Wiki.

We call a sequence of the form {a + nd : n ∈ Z} an arithemetic progression. Note


that in our note it is infinite to both the left and the right.

Space 9. The evenly spaced integer topology (Z, Ω): Ω is generated by all
arithmetic progressions.
33

Furstenberg’s proof of the infinitude of primes. In the evenly spaced integer topology,
each set P (a, d) = {a + nd : n ∈ Z} is clopen. Note that
[
Z \ {−1, 1} = P (0, p).
p is prime

If the number of primes is finite, then the set {−1, 1} is clopen. However, every open
set is infinite from definition, a contradiction. □

2.5. Hierarchy of topologies. In this section, we consider a set X with two topologies
Ω1 and Ω2 .

Definition 2.47. When a set X has two topologies Ω1 and Ω2 such that Ω1 ⊆ Ω2 , we
say that Ω1 is coarser than Ω2 , and that Ω2 is finer than Ω1 .

Proposition 2.48. Given a set X. Then the discrete topology is the finest topology
on X, and the indiscrete topology is the coarsest topology on X.

Example 2.49. The cofinite topology on the real line is coarser than the canonical
topology.

Proposition 2.50. Given (X, Ω1 ) and (X, Ω2 ) s.t. Ω1 ⊆ Ω2 . Let A ⊆ X.

1) Denote by Inti the interior w.r.t. Ωi . Then Int1 A ⊆ Int2 A.


2) Denote by Cli the closure w.r.t. Ωi . Then Cl1 A ⊇ Cl2 A.

2.6. Metrics & the metric topology.

Definition 2.51. Let X be a set.

• A metric on X: a function d : X × X → R s.t.


(a) triangle inequality: d(x, y) ≤ d(y, z) + d(z, x);
(b) symmetry: d(x, y) = d(y, x); and
(c) identity of indiscernibles: d(x, y) = 0 ⇐⇒ x = y.
• A metric space: the pair (X, d).

We distinguish the notation of a topological space (X, ΩX ) and the notation of a


metric space (X, d) by statement. As will be seen in Definition 2.69, a metric induces a
topological structure.
34 D.G.L. WANG

Proposition 2.52. Every metric is nonnegative.

Proof. Let (X, d) be a metric space and let x, y ∈ X. By the triangle inequality, we
have
d(x, y) + d(y, x) ≥ d(x, x).
Applying the symmetry property to the second summand on the left side, and using
the identity of indiscernibles on the right side, we obtain
2d(x, y) ≥ 0,
which reduces to the nonnegativity by dividing 2 on both sides. □

Example 2.53. Let d1 and d2 be metrics on X. Then ρ+ = d1 + d2 is a metric, and so


is ρmax = max{d1 , d2 }.

Proof. The symmetry and positive semi-definity for each of the functions ρ+ and ρmax
are clear. The triangle inequality is shown below.
 
ρ+ (x, z) + ρ+ (y, z) = d1 (x, z) + d2 (x, z) + d1 (y, z) + d2 (y, z)
 
= d1 (x, z) + d1 (y, z) + d2 (x, z) + d2 (y, z)
≥ d1 (x, y) + d2 (x, y) = ρ+ (x, y);
ρmax (x, z) + ρmax (y, z) = max{d1 (x, z), d2 (x, z)} + max{d1 (y, z), d2 (y, z)}
≥ max{d1 (x, z) + d1 (y, z), d2 (x, z) + d2 (y, z)}
≥ max{d1 (x, y), d2 (x, y)} = ρmax (x, y).
The identity of indiscernibles can be shown as follows.
ρ+ (x, y) = 0 ⇐⇒ d1 (x, y) = 0 and d2 (x, y) = 0 ⇐⇒ x = y.
ρmax (x, y) = 0 ⇐⇒ d1 (x, y) = 0 and d2 (x, y) = 0 ⇐⇒ x = y.
This completes the proof. □

Example 2.54. The map d∆ (A, B) = S(A) + S(B) − 2S(A ∩ B), where S(C) is the
area of a polygon C, defines the area metric on the set of all bounded plane polygons.
In fact, the number d∆ (A, B) is the area of the symmetric difference A∆B. Thus
the symmetry and being positive-definite are clear. The subadditivity holds because
A∆C ⊆ (A∆B) ∪ (B∆C) for any bounded plane polygons A, B, C.

Proposition 2.55. Let d be a metric on a set X, and f a function defined on R≥0 .


Then the function g(x, y) = f (d(x, y)) is a metric if
35

i) f (x) = 0 if and only if x = 0;


ii) f is increasing; and
iii) f is subadditive, i.e., f (x + y) ≤ f (x) + f (y) for any x, y ∈ R.

Proof. We check the 3 conditions for a metric.

(1) Triangle inequality.

f (d(x, y)) ≤ f (d(x, z) + d(y, z)) (by the increasing property of f )


≤ f (d(x, z)) + f (d(y, z)) (by the subadditivity of f ).

(2) Symmetry. f (d(x, y)) = f (d(y, x)) by the symmetry of d.


(3) Identity of indiscernibles. By the first two properties, we have

f (d(x, y)) = 0 ⇐⇒ d(x, y) = 0 ⇐⇒ x = y.

This completes the proof. □

See Appendix B for a bit more information on subadditive functions.

Proposition 2.56. For any metric d, the function min{d(x, y), 1} is a metric.

Proof. We check that the function ρ(x, y) = min{d(x, y), 1} satisfy the 3 conditions for
a metric.

(1) Triangle inequality. Since ρ(x, z) = min{d(x, z), 1} ≤ 1, by the non-negativity of the
metric d, the desired inequality holds if d(x, y) ≥ 1 or d(y, z) ≥ 1. Consequently, we
may suppose that d(x, y) < 1 and d(y, z) < 1. Then, by the triangle inequality of d,
we infer that

ρ(x, y) + ρ(y, z) = d(x, y) + d(y, z) ≥ d(x, z) ≥ ρ(x, z).

(2) Symmetry: Direct by the symmetry of d.


(3) Identity of indiscernibles:

ρ(x, y) = 0 ⇐⇒ d(x, y) = 0 ⇐⇒ x = y.

This completes the proof. □


36 D.G.L. WANG

In linear algebra, functional analysis and related areas, we use the concept of norm,
which has a close relationship with metrics.

Definition 2.57. Let V be a vector space over R.

• A seminorm on V : a function ∥∥ : V → R satisfying


(1) subadditivity: ∥u + v∥ ≤ ∥u∥ + ∥v∥, ∀ u, v ∈ V ; and
(2) absolute homogeneity: ∥λv∥ = |λ|· ∥v∥, ∀ λ ∈ R, v ∈ V .
• Seminorm is nonnegative: ∥v∥ ≥ 0, ∀ v ∈ V .
(2) ⇒ ∥0∥ = 0 and ∥−v∥ = ∥v∥. Taking u = −v in (1) ⇒ ∥v∥ ≥ 0.
• The trivial seminorm: ∥v∥ = 0, ∀ v ∈ V .
• A norm on V : a function ∥∥ satisfying (1),(2), and
(3) identity of indiscernibles: ∥v∥ = 0 ⇒ v = 0.
• A normed space: a vector space equipped with a norm.
• The metric induced by the norm: d(x, y) = ∥x − y∥.

Notation 2.58. We denote H = {x : x ≥ 1} ∪ {+∞}.

Definition 2.59. Let x = (x1 , . . . , xn ) ∈ Cn and p ∈ H.

• The p-norm of x:
n
!1/p
X p
(1) ∥x∥p = |xi | .
i=1

The subadditivity of the p-norm is called Minkowski’s inequality; see Theorem A.6.
• The grid/Manhattan/taxicab norm or snake distance: the 1-norm. See Fig. 18.

• The Euclidean norm: the 2-norm.


• The infinity/maximum norm: ∥x∥∞ = maxi |xi |.
In fact, if M = maxi |xi |, then
n
!1/p n
!1/p
X X |xi |p
∥x∥∞ = lim |xi |p = M · lim
p→∞
i=1
p→∞
i=1
Mp
= M · lim |{1 ≤ i ≤ n : |xi | = M }|1/p = M.
p→∞
37

Figure 18. The Manhattan distance. Stolen from Bing image.

Example 2.60. Let p ∈ H. Consider the metric


n
!1/p
X
(2) ρ(p) (x, y) = |xi − yi |p
i=1

induced by the p-norm.


Pn
(1) ρ(1) (x, y) = i=1 |xi − yi |.
n
p
(2) ρ(2) (x, y) =
P 2
i=1 (xi − yi ) is the Euclidean metric.

(3) The Chebyshev/chessboard distance or L∞ -metric:


ρ(∞) (x, y) = max |xi − yi |.
i

The Chebyshev distance is used in warehouse logistics, as it effectively measures the


time an overhead crane takes to move an object. In fact, the crane can move on the
x and y axes at the same time but at the same speed along each axis; see Fig. 19.

Definition 2.61 (generalization). Let x = (x1 , . . . , xn ) ∈ Cn and p ∈ H.

• The ℓp -space: the set of infinite complex-valued sequences (x1 , x2 , . . .) with conver-
gent ℓp -norm
!1/p
X p
∥(x1 , x2 , . . .)∥ = |xi | .
i

• The L -space: the set of functions f on X ⊆ R with convergent Lp -norm


p

Z !1/p
∥f ∥p,X = |f (x)|p dx .
X

Definition 2.62. Let (X, d) be a metric space. Let r ≥ 0.


38 D.G.L. WANG

Figure 19. Pafnuty Lvovich Chebyshev (1821–1894) was a Russian


mathematician. His name can be alternatively transliterated as Cheby-
chev, Chebysheff, Chebychov, Chebyshov; or Tchebychev, Tchebycheff
(French transcriptions); or Tschebyschev, Tschebyschef, Tschebyscheff
(German transcriptions). The right part is a photo for an overhead crane
in Qingdao Beihai Shipbuilding Heavy Industry in China.

• The open ball of radius r centerd at x ∈ X:


B(x, r) = {y ∈ X : d(x, y) < r}.

• The disk of radius r centerd at x ∈ X:


D(x, r) = {y ∈ X : d(x, y) ≤ r}.

• The sphere of radius r centerd at x ∈ X:


S(x, r) = {y ∈ X : d(x, y) = r}.

Example 2.63. The names in Definition 2.62 come from the Euclidean space.

• Dn : the unit disk in Rn , i.e., the disk of radius 1 centered at 0.


D1 is the interval [−1, 1], D2 is a plane disk, and D3 a ball.
• S n : the unit sphere in Rn+1 , i.e., the sphere of radius 1 centered at 0.
S 0 is the set {−1, 1}, S 1 is a circle, and S 2 a sphere.

However, they may rather surprising in other metric spaces.



• The unit ball in R2 equipped with ρ(1) is a closed square of side length 2.
• The unit ball in R2 equipped with ρ(∞) is a closed square of side length 2.
39

Definition 2.64. Let (X, d) be a metric space with A, B ⊆ X and x ∈ X.

• The diameter of A: diam(A) = supx,y∈A d(x, y).


• A is bounded: A is contained in a ball, i.e., ∃ r s.t. d(x, y) < r, ∀ x, y ∈ A.
• The distance between x and A: d(x, A) = inf a∈A d(x, a).
• The Hausdorff distance between A and B:
 
dH (A, B) = max sup d(a, B), sup d(A, b) .
a∈A b∈B

See Fig. 20.

Figure 20. The Hausdorff distance.

Remark 2.65. In computer graphics the Hausdorff distance is used to measure the
difference between two different representations of the same 3D object.

Theorem 2.66. Let (X, d) be a metric space with a closed set A and a point b. Then
the following equivalence holds:
d(A, b) = 0 ⇐⇒ b ∈ A.

Proof. ⇐: Clear. ⇒: Since d(A, b) = 0, each ball centered at b has a nonempty


intersection with A. Hence b ̸∈ Ac . Thus b ∈ Cl(A) = A. □

Proposition 2.67. Let (X, d) be a metric space. Let x ∈ X and A, B ⊆ X.

(1) d(x, A) = 0 ⇐⇒ x ∈ A.
(2) dH (A, B) is finite if both A and B are bounded.
(3) dH (A, B) = 0 ⇐⇒ A = B.
40 D.G.L. WANG

Theorem 2.68. For every metric space, the Hausdorff distance is a metric on the set
of its closed bounded subsets.

Proof. We should verify the three conditions for a metric. Let (X, d) be a metric space.
Let A, B, C be closed bounded subsets of X.

(1) Symmetry and positivity. Clear from definition.


(2) Being definite. By Theorem 2.66, we can infer that

dH (A, B) = 0 ⇐⇒ sup d(a, B) = sup d(A, b) = 0


a∈A b∈B

⇐⇒ d(a, B) = d(A, b) = 0 for all a ∈ A and b ∈ B


⇐⇒ a ∈ B and b∈A for all a ∈ A and b ∈ B
⇐⇒ A ⊆ B and B⊆A
⇐⇒ A = B.

(3) Subadditivity. Let r(A, B) = supa∈A d(a, B). Then



dH (A, B) = max r(A, B), r(B, A) .

To show the subadditivity of dH , it suffices to show subadditivity of the function r.


By the subadditivity of d, for all a ∈ A and all b ∈ B, we have

d(a, C) ≤ d(a, b) + d(b, C) ≤ d(a, b) + r(B, C).

It follows from the arbitrariness of b, we infer that

r(A, C) ≤ d(a, C) ≤ d(a, B) + r(B, C) ≤ r(A, B) + r(B, C).

This completes the proof. □

Homework 2.3. Find a metric space and two balls in it s.t. the smaller ball contains
the bigger ball properly. What is the minimal number of points in such a space? Show
that the largest radius in such a space is at most twice the smaller radius.

Homework 2.4. Show that the segment with endpoints a, b ∈ Rn can be described as

{x ∈ Rn : ρ(a, x) + ρ(b, x) = ρ(a, b)}

under the Euclidean metric. How does the above set look if ρ = ρ(1) or ρ = ρ(∞) ?

Definition 2.69. Let X be a set.


41

• Given a metric d on X. The metric topology of (X, d), or the topology of X


generated by d:
ΩX,d = {O ⊆ X : ∀ x ∈ O, ∃ ϵ, B(x, ϵ) ⊆ O}.
In other words, the collection of open balls in (X, d) is a base of Ωd .
• Given a topology Ω on X. X is metrizable: Ω is generated by a certain metric.
Homework 2.5. Let X be a set. Define the function d0 : X × X → R≥0 by

0, if x = y,
(3) d0 (x, y) =
1, if x ̸= y.

(1) Show that d0 is a metric.


(2) Let x ∈ X. Find D(x, 1), D(x, 1/2), and S(x, 1/2) in this metric space.
(3) What is the metric topology of (X, d0 )?
Theorem 2.70. We have the following.

(1) Any metrizable indiscrete space is a singleton.

Proof. Let (X, ΩX ) be an indiscrete space. Then ΩX = {∅, X} from definition.


Suppose that |X| ≥ 2. Then X contains two distinct points, say, x and y. If X is
metrizable, say, ΩX is generated by some metric d. Then d(x, y) > 0 from definition,
and the set S = B(x, d(x, y)/2) is open. But x ∈ S and y ̸∈ S, we infer that S ̸∈ ΩX ,
a contradiction! This completes the proof. □

(2) A finite space is metrizable if and only if it is discrete.

Proof. Let (X, ΩX ) be a finite space. If X is discrete, then ΩX = 2X from definition.


Under the metric d0 defined by Eq. (3), every singleton x ∈ X forms an open set
since {x} = B(x, 1/2) ∈ ΩX . Hence every subset of X is open under the topology
generated by d0 . Conversely, suppose that ΩX is generated by a metric d. Let
mx = min{d(x, y) : y ∈ X \ {x}}.
Since X is finite, we have mx > 0. Then every singleton {x} = B(x, mx /2) is open.
Thus ΩX is the discrete topology. This completes the proof. □

Theorem 2.71. Two metrics d1 and d2 on X are equivalent if they generate the
same topology, denoted d1 ∼ d2 . If ∃ c, C > 0 s.t. c · d1 (x, y) ≤ d2 (x, y) ≤ C · d1 (x, y),
∀ x, y ∈ X, then d1 ∼ d2 .
42 D.G.L. WANG

Table 1. Openness & closeness of balls and spheres.

open ball disk sphere


Open YES MAYBE MAYBE
Closed MAYBE YES YES

Proof. Write ΩX,di = Ωi for i = 1, 2. The desired equivalence says Ω1 = Ω2 .

First, we show that Ω2 ⊆ Ω1 . Since any open set O2 ∈ Ω2 is the union of some open
balls w.r.t. the metric d2 , it suffices to recast a typical ball of the form
B(x, r) = {z ∈ X : d2 (x, z) < r}
as a member in Ω1 , i.e., the union of open balls w.r.t. the metric d1 . Define
[  
U (x, r) = y ∈ X : d1 (y, z) < L
z∈B(x,r)

where
r − d2 (x, z)
L= .
2C
Then U (x, r) ∈ Ω1 . We shall show that B(x, r) = U (x, r). On one hand, ∀ z ∈ B(x, r),
since d1 (z, z) = 0 < L, we have z ∈ U (x, r). On the other hand, ∀ y ∈ U (x, r),
∃ z ∈ B(x, r) s.t. d1 (y, z) < L. Then
d2 (x, y) ≤ d2 (x, z) + d2 (y, z) ≤ d2 (x, z) + C· d1 (y, z) ≤ d2 (x, z) + C· L < r.
Thus y ∈ B(x, r). This proves B(x, r) = U (x, r) and Ω2 ⊆ Ω1 .

From premise, we have d1 (x, y) ≤ c−1 · d2 (x, y). Along the same lines, one may show
that Ω1 ⊆ Ω2 , which implies Ω1 = Ω2 immediately. □

Homework 2.6. Let H = {x ∈ R : x ≥ 1} ∪ {+∞}. The metrics ρ(p) defined in Eq. (2)
for p ∈ H are equivalent.

Hint. By Theorem 2.71. □

Homework 2.7. Let d be a metric on a set X. Then the function d/(1 + d) is a metric,
and it is equivalent to d.
43

3. Continuous Maps & Homeomorphisms

3.1. Continuous maps.

Definition 3.1. Given (X, ΩX ) and (Y, ΩY ). Let f : X → Y be a map.

• f is open: open 7→ open, i.e., f (ΩX ) ⊆ ΩY .


• f is closed: closed 7→ closed.
• f is continuous: the preimage of any open set is open.
• f is continuous at a point x ∈ X: the set f −1 (V ) is a neighbourhood of x for
every neighbourhood V of f (x).

Example 3.2. Here are some continuous maps.

(1) The identity map, i.e., the map id : X → X defined by id(x) = x.


(2) Any constant map, i.e., any map f : X → Y such that f (x) = y0 for all x ∈ X and
for some fixed y0 ∈ Y .
(3) Any inclusion, i.e., any map f : X → Y where X ⊆ Y and ΩX is the subspace
topology of ΩY , defined by f (x) = x.

Example 3.3. Let X = [0, 2] with subspace topology of ΩR .

(1) Let Y = X be the same space. Then the map f : X → Y defined by



x, if x ∈ [0, 1),
f (x) =
3 − x, if x ∈ [1, 2].

is not continuous.
(2) Let Y be the arrow. Then the map g : X → Y defined by

x, if x ∈ [0, 1],
g(x) =
x + 1, if x ∈ (1, 2].

is continuous!

Proof. The topology of the space X is X ∩ ΩR . We denote it by ΩX .

(1) The point x = 1 is a jump discontinuity.


44 D.G.L. WANG

(2) We check the continuity of g by definition. It is clear that g −1 ([0, ∞)) = X is


open. For a typical open set (t, ∞) of the arrow with t ≥ 0,
g −1 ((t, ∞)) = (s, 2] = (s, ∞) ∩ X ∈ ΩX ,
where 

 t − 1, if t ≥ 2,

s = 1, if 1 ≤ t ≤ 2,


t, if 0 ≤ t ≤ 1.

Question 3.4. Which maps X → Y are continuous if

1) X is discrete?
Answer. All maps. □

2) Y is indiscrete?
Answer. All maps. □

Theorem 3.5 (Characterizations of map continuity). Given topological spaces (X, ΩX )


and (Y, ΩY ). Let f : X → Y be a map. Then the followings are equivalent.

(1) The preimage of any open set under f is open.


(2) The preimage of any closed set under f is closed.
(3) The preimage of any member of a base under f is open.
(4) The preimage of any neighbourhood under f is a neighbourhood, i.e., f is continuous
at each point.
(5) f (A) ⊆ f (A), ∀ A ⊆ X.
(6) f −1 (B) ⊆ f −1 (B), ∀ B ⊆ Y .

Proof. Straightforward by definition: each of (2) to (4) is equivalent to (1).

(4) ⇐⇒ (5):

⇒. Let x ∈ A. Let V be a neighbourhood of f (x). By (4), the set f −1 (V ) is a


neighbourhood in X. Since f −1 (V ) is a neighbourhood of x, we have f −1 (V ) ∩ A ̸= ∅,
and
V ∩ f (A) = f f −1 (V ) ∩ A ̸= ∅.

45

⇐. Let x ∈ X. Let V be a neighbourhood of f (x). Assume that (4) is false. Then


f −1 (V ) is not a neighbourhood of x. Define A = X \ f −1 (V ). Then x ∈ A. It
follows that f (x) ∈ f (A) ⊆ f (A). Thus f (A) meets every neighbourhood of f (x).
In particular, we have f (A) ∩ V ̸= ∅, contradicting the definition of A.

(6) ⇐⇒ (2):

⇒. Let B be a closed subset of Y . Then B = B, and f −1 (B) ⊆ f −1 (B) = f −1 (B). It


follows that f −1 (B) = f −1 (B) is closed.
⇐. Let B ⊆ Y . Since B ⊆ B, we have f −1 (B) ⊆ f −1 (B). Since B is closed, we infer
that f −1 (B) is closed. Since f −1 (B) is the smallest closed set containing f −1 (B), we
derive that f −1 (B) ⊆ f −1 (B).

This completes the proof. □

Theorem 3.6. Continuous surjections preserve the property of being everywhere dense.

Proof. Let f : X → Y be a surjective continuous map between topological spaces. Let


A be an everywhere dense set of X. Since f is surjective, we have
f (A) = f (X) = Y.
Since f is continuous, by Theorem 3.5 (5), we have f (A) ⊆ f (A). Therefore, we obtain
f (A) = Y , i.e., f (A) is everywhere dense. □

Theorem 3.7. Let f1 , . . . , fl , g1 , . . . , gm , h1 , . . . , hn : X → R be continuous functions.


Let
  
f1 (x) ≤ 0, . . . , fl (x) ≤ 0,

 
 

 
A = x ∈ X : g1 (x) = 0, . . . , gm (x) = 0, and

 
 

h1 (x) ≥ 0, . . . , hn (x) ≥ 0
  
  
 f (x) < 0, . . . , f (x) < 0, 
1 l
B = x ∈ X:
 h1 (x) > 0, . . . , hn (x) > 0. 

Then A is closed and B is open.

Proof. The set A is the intersection of the preimages of the closed sets (−∞, 0 ] or {0}
or [ 0, ∞) under continuous functions, and hence closed. The set B is the intersection
of the preimages of the open sets (−∞, 0) or (0, ∞) under continuous functions, and
hence closed. □
46 D.G.L. WANG

Corollary 3.8. The kernel {x ∈ X : f (x) = 0} of any continuous f : X → R is


closed. □

Corollary 3.9. The set {x ∈ R : f (x) = x} of fixed points of any continuous f : R → R


is closed. □

Lemma 3.10. Let f : X → R be a continuous function. Then ∀ p ∈ X, ∃ a neighbour-


hood U of p s.t. f is bounded in U . If 0 ̸∈ f (X), then ∃ a neighbourhood V of p in
which, for any x ∈ V ,

(i) the absolute value |f (x)| has a positive lower bound in V , and
(ii) f (x)f (p) > 0.

Proof. Let p ∈ X and ϵ > 0. Since f is continuous, ∀ ϵ > 0, ∃ a neighbourhood U of p


s.t. |f (x) − f (p)| < ϵ. Thus |f (x)| ≤ |f (p)| + ϵ, i.e., f is bounded in U .

If 0 ̸∈ f (X), then f (p) ̸= 0, one may require the above ϵ < |f (p)/2|. Then
|f (x)| ≥ |f (p)| − ϵ > ϵ. Moreover, the inequality |f (x) − f (p)| < ϵ implies that
f (x)f (p) > 0. □

Theorem 3.11. Let f, g : X → R be continuous functions. Then the functions X → R


defined below are continuous:

x 7→ f (x) ± g(x),
x 7→ f (x)g(x),
x 7→ f (x)/g(x), if 0 ̸∈ g(X),
x 7→ |f (x)|,
x 7→ min{f (x), g(x)},
x 7→ max{f (x), g(x)}.

Proof. We show them individually. Let p ∈ X and ϵ > 0.

(1) Let h(x) = f (x) + g(x) and w ∈ {f, g}. Since w is continuous, ∃ a neighbourhood
Uw of p s.t. |w(x) − w(p)| < ϵ/2, ∀ x ∈ Uw . Hence

|h(x) − h(p)| ≤ |f (x) − f (p)| + |g(x) − g(p)| < ϵ, ∀ x ∈ U,

where U = Uf ∩ Ug is a neighbourhood of p. In other words, ∀ neighborhood

J = (h(p) − ϵ, h(p) + ϵ) ∈ ΩR ,
47

its preimage h−1 (J) contains U , which contains an open set. Thus h−1 (J) is a
neighbourhood in X. This proves the continuity of h(x).
The continuity of f − g can be obtained by considering the continuous functions f
and −g.
(2) Let h(x) = f (x)g(x). Since f is continuous, by Lemma 3.10, ∃ a neighbourhood Uf
of p and a bound b > 0 s.t. for any x ∈ Uf ,

|f (x)| < b and |g(p)| · |f (x) − f (p)| < ϵ/2.

Since g is continuous, ∃ a neighbourhood Ug of p s.t. for any x ∈ Ug ,

|g(x) − g(p)| < ϵ/(2b).

Therefore, one may infer that for any x ∈ Uf ∩ Ug ,

|h(x) − h(p)| = |f (x)g(x) − f (p)g(p)|


≤ |f (x)| · |g(x) − g(p)| + |g(p)| · |f (x) − f (p)| < ϵ,

This proves the continuity of h(x).


(3) By using the continuity of the product of continuous functions, it suffices to show
that the function h(x) = 1/g(x) is continuous. In fact, by Lemma 3.10, ∃ a bound
b > 0 and a neighbourhood U1 of p s.t. |g(x)g(p)| > b. By the continuity of g, ∃ a
neighbourhood U2 of p s.t. |g(x) − g(p)| < ϵb. Therefore, we can deduce that

1 1 g(x) − g(p)
g(x) − g(p) = g(x)g(p) < ϵ, ∀ x ∈ U1 ∩ U2 .

(4) Let h(x) = |f (x)|. By Lemma 3.10, ∃ a neighbourhood U1 of p s.t. f (x)f (p) > 0.
Since f is continuous, ∃ a neighbourhood U2 of p s.t. |f (x) − f (p)| < ϵ. Therefore,
we can infer that

|h(x) − h(p)| = |f (x) − f (p)| < ϵ, ∀ x ∈ U1 ∩ U2 .

(5) Direct by using the continuity of the sum, difference, absolute values of two continuous
functions, and the identity
f (x) + g(x) |f (x) − g(x)|
min{f (x), g(x)} = − .
2 2
(6) The continuity of the function max{f, g} = − min{−f, −g} can be obtained by the
continuity of the minimum of continuous functions.

This completes the proof. □


48 D.G.L. WANG

Theorem 3.12. A function

f : X → Rn

x 7→ f1 (x), . . . , fn (x)

is continuous ⇐⇒ so are all the functions f1 , f2 , . . . , fn .

Hint. The topology on Rn is generated by the infinity metric

ρ(∞) (x, y) = max{|x1 − y1 |, . . . , |xn − yn |}.

Observe that ρ(∞) (f (x), f (p)) < ϵ ⇐⇒ |fi (x) − fi (p)| < ϵ for each i. □

Homework 3.1. If a function f : R → R is continuous, then so is the graph map

Γf : R → R2

x 7→ x, f (x) .

The definition of continuity usually studied in Calculus, when applicable, is equivalent


to the definition stated in terms of topological structures.

Theorem 3.13. Let f : (X, dX ) → (Y, dY ) be a map between metric spaces. Let a ∈ X.
The followings are equivalent.

(1) f is continuous at a point a.


(2) Each ball centered at f (a) contains the image of a ball centered at a.
(3) ∀ ϵ > 0, ∃ δ > 0, ∀ x ∈ X, the inequality dX (x, a) < δ implies

dY (f (x), f (a)) < ϵ.

Theorem 3.14. For any subset A of a metric space (X, d), the function

dA : X → R
x 7→ d(x, A)

is continuous.

Hint. |dA (x) − dA (y)| = |inf a∈A d(x, a) − inf a∈A d(y, a)| ≤ d(x, y). □
49

3.2. Covers. In order to show the continuity of a map f : X → Y , we would like to


find a suitable cover of X and show that (i) the continuity of f restricted on each part
of the cover, and (ii) the cover is suitable enough so that (i) is sufficient to guarantee
the continuity of f : X → Y . As will be seen, some covers, called fundamental covers,
satisfies Condition (ii), which is the topic that we will focus on in this section.

Definition 3.15. Let X be a set.

• A cover or covering of X: a collection Γ of sets s.t. X ⊆ ∪A∈Γ A.


In this case, we say that Γ covers X.
• A subcover or subcovering of a cover Γ: a subfamily of Γ which covers X.
• A refinement of a cover Γ: a cover of X whose every member is contained in a
member of Γ.
• A partition of X: a cover Γ of X s.t. X = ⊔A∈Γ A.
To partition X: to construct a partition of X.
• Suppose that X = ∪A∈Γ A is a topological space. We say that the cover Γ is
– fundamental: the following equivalence holds

a set O ⊆ X is open ⇐⇒ O ∩ A is open in A, ∀ A ∈ Γ.

– open: every member in Γ is open.


– closed: every member in Γ is closed.
– locally finite: every point in X has a neighbourhood meeting only a finite number
of members in Γ.

Theorem 3.16. Let f : X → Y be a map between topological spaces. Let Γ be a


fundamental cover of X s.t. f |A is continuous ∀ A ∈ Γ, then f is continuous.

Proof. Let O be an open set in Y . Since f |A is continuous, we have

f −1 (O) ∩ A = (f |A )−1 (O)

is open in A, ∀ A ∈ Γ. Since Γ is a fundamental cover, we find f −1 (O) is open in X.


Hence f is continuous. □

Remark 3.17. Theorem 3.16 provides a way of proving the continuity of a map in the
spirit of divide and conquer.
50 D.G.L. WANG

Theorem 3.18 (Characterizations of fundamental covers). Let X be a set with a cover


Γ. Then

Γ is fundamental
⇐⇒ a set O ⊆ X is open, provided O ∩ A is open in A, ∀ A ∈ Γ
⇐⇒ a set F ⊆ X is closed, provided F ∩ A is closed in A, ∀ A ∈ Γ.

Proof. For the first equivalence, it suffices to check from right to left. In fact, when O
is open in X, by definition of the subspace topology, the intersection O ∩ A is certainly
an open set of A.

Let us show the second equivalence. ⇒: ∀ A ∈ Γ, the set (X \ F ) ∩ A = A \ (F ∩ A)


is open. Thus X \ F is open, i.e., F is closed. The converse statement can be shown
similarly. □

Theorem 3.19 (Criteria for the fundamentality). A cover Γ is fundamental

1) if it has a fundamental refinement;

Proof. Let O ⊆ X. Suppose that O ∩ A is open in A, ∀ A ∈ Γ. Let Γ′ be a


fundamental refinement of Γ. Let B ∈ Γ′ . Then ∃ AB ∈ Γ s.t. B ⊆ AB . By
Theorem 2.18, we can regard the topology of AB ∩ B as the one induced from AB .
It follows that O ∩ B = O ∩ AB ∩ B is open in AB ∩ B = B. By the arbitrariness of
B, and since Γ′ is fundamental, we infer that O is open in X. This proves that Γ is
fundamental. □

2) if it is open or finite closed;

Proof. Let O ⊆ X s.t. O ∩ A is open in A, ∀ A ∈ Γ. By Theorem 2.16, O ∩ A is open


in X, and so is the set ∪A∈Γ (O ∩ A) = O. By Theorem 3.18, Γ is fundamental. In
the same vein, every finite closed cover is fundamental. □

3) if it is locally finite and closed;

Proof. Each point p ∈ X has an open neighbourhood Op meeting only a finite number
of members of Γ. Then the cover

Σ = {A ∩ Op : A ∈ Γ}

of Op is finite and closed, and thus fundamental. Let O ⊆ X be a set s.t. O ∩ A is


open in A, ∀ A ∈ Γ. We shall show that O is open in X.
51

In fact, let p ∈ X and A ∈ Γ. Consider the topology of A ∩ Op as the one induced


from the topology of A, we infer that
(O ∩ Op ) ∩ (A ∩ Op ) = (O ∩ A) ∩ Op
is open in A ∩ Op . By the arbitrariness of A, and since the cover Σ is fundamental,
we infer that the set O ∩ Op is open in Op , and thus open in X by Theorem 2.16.
Hence [ [
O =O∩X =O∩ Op = (O ∩ Op )
p∈X p∈X
is open in X. This completes the proof. □

4) if ∃ a fundamental cover Σ s.t. the cover ΣS = {A ∩ S : A ∈ Γ} of S is fundamental


∀ S ∈ Σ;

Proof. Let O ⊆ X s.t. O ∩ A is open in A, ∀ A ∈ Γ. Let S ∈ Σ. Then O ∩ A ∩ S is


open in A ∩ S. Since the cover ΣS is fundamental, we infer that O ∩ S is open in S.
Since the cover Σ is fundamental, we derive that O is open in X. This proves that Γ
is fundamental. □

5) if every point has a neighbourhood U s.t. the cover ΓU = {A ∩ U : A ∈ Γ} of U is


fundamental. In other words, the cover fundamentality is a local property.

Proof. ∀p ∈ X, let Up be a neighbourhood of p s.t. the cover ΓUp is fundamental. We


claim that the cover Σ = {Up : p ∈ X} of X is fundamental. In fact, suppose that
O ⊆ X s.t. O ∩ Up is open in Up , ∀ p ∈ X. Since Up is a neighbourhood of p, ∃ open
set Op s.t. p ∈ Op ⊆ Up . Therefore, O ∩ Op = O ∩ Up ∩ Op is open in Op = Up ∩ Op ,
and thus open in X. Hence the set O = ∪p∈X (O ∩ Op ) is open in X. This proves
the claim. Now, by using Item 4), we obtain that Γ is fundamental. □

Remark 3.20. The above characterizations will be used when we meet path products.
52 D.G.L. WANG

3.3. Homeomorphisms.

Definition 3.21. Given topological spaces X and Y .

• An invertible map f : X → Y is a homeomorphism: both f and f −1 are continuous.


• X is homeomorphic to Y : ∃ a homeomorphism between X and Y .

In this note the homeomorphism relation is denoted by ∼


=.

Question 3.22. Consider the map f : [0, 1) → S 1 defined by

f (x) = (cos 2πx, sin 2πx).

Is it an invertible map? A ontinuous map? A homeomorphism?

Answer. Yes. Yes. No. □

Homework 3.2. What is the difference between the terms “homomorphism” and
“homeomorphism” in mathematics?

Proposition 3.23. We have the following.

(1) The identity map on a space is a homeomorphism.


(2) The inverse of a homeomorphism is a homeomorphism.
(3) The composition of homeomorphisms is a homeomorphism.
(4) Being homeomorphic is an equivalent relation.

Proof. The last statement is a corollary of the first three, which are direct from definition.

Theorem 3.24. Let h : (X, ΩX ) → (Y, ΩY ) be a homeomorphism. Let x ∈ A ⊆ X.

(1) A is open ⇐⇒ h(A) is open.


(2) A is closed ⇐⇒ h(A) is closed.
(3) hk = kh.
(4) hi = ih.
(5) A is a neighbourhood of x ⇐⇒ h(A) is a neighbourhood of the point h(x).
53

Proof. Since h is invertible, the inverse map g = h−1 is continuous and satisfies
A = gh(A) and h(A) = g −1 (A).
(1). ⇐: The set A is the preimage of the open set h(A) under the continuous map h,
and thus open.

⇒: The set h(A) is the preimage g −1 (A) of the open set A under the continuous map g,
and thus open.

(2). Immediate by taking complement of open sets and using (1).

(3). By Theorem 3.5,


h(A) ⊆ h(A) and h(A) = g −1 (A) ⊆ g −1 (A) = h(A).
This proves h(A) = h(A).

(4). Recall that ci = kc by Theorem 2.27. Thus i = ckc. Since h is invertible, we have
hc = ch. we can deduce that
hi = hckc = chkc = ckhc = ckch = ih.

(5). ⇐: Suppose that x ∈ O ⊆ A, where O ∈ ΩX . Then h(O) ∈ ΩY by (1). Since


h(x) ∈ h(O) ⊆ h(A),
we find h(A) is a neighborhood of h(x).

⇒: Suppose that h(x) ∈ O′ ⊆ h(A), where O′ ∈ ΩY . Then g(O′ ) ∈ ΩX by (1). Since


x = gh(x) ∈ g(O′ ) ⊆ gh(A) = A,
we find A is a neighborhood of x. □

Remark 3.25. Homeomorphic spaces are completely identical from the topological point
of view: a homeomorphism X → Y establishes a one-to-one correspondence between
“all phenomena” in X and Y that can be expressed in terms of topological structures.

When the concept of topological space had not yet been developed, mathematicians
studied only subspaces of Euclidean spaces, and continuous maps and homeomorphisms
between Euclidean spaces. Christian Felix Klein, in his famous Erlangen Program, clas-
sified various geometries that had emerged up to that time, like Euclidean, Lobachevsky,
affine, and projective geometries, and defined topology as a part of geometry that deals
with properties preserved by homeomorphisms.
54 D.G.L. WANG

Figure 21. Christian Felix Klein (1849–1925) was a German mathe-


matician. His 1872 Erlangen Program, classifying geometries by their
underlying symmetry groups, was a highly influential synthesis of much
of the mathematics of the day.

Theorem 3.26. A map f : X → Y is a homeomorphism ⇐⇒ f is bijective and


determines a bijection between the topologies of X and Y .

Hint. Direct by the definition of a homeomorphism and by using Theorem 3.24 (1). □

Here is a useful criterion for homeomorphisms.

Theorem 3.27. Let f be a continuous bijection. Then

f is a homeomorphism ⇐⇒ f is closed. ⇐⇒ f is open.

Proof. We show the first equivalence. The forward direction is clear from Theo-
rem 3.24 (2). For the backward direction, we show by the definition of a homeomorphism.
Since f is bijective, the inverse g = f −1 exists. Suppose that f : X → Y . For any closed
set F ⊆ X, the preimage of F under g is g −1 (F ) = f (F ) is closed since f is closed. By
Theorem 3.5, g is continuous, and thus f is a homeomorphism. The second equivalence
can be shown in the same fashion. □

Example 3.28. Most of the following can be deduced by using Theorem 3.26.

1) Every bijection on an indiscrete space onto itself is a homeomorphism.


2) Every bijection on a discrete space onto itself is a homeomorphism.
3) Every bijection on the cofinite space onto itself is a homeomorphism.
55

4) A discrete space and a non-one-point indiscrete space are not homeomorphic.


5) Spaces containing different numbers of points are not homeomorphic.
6) The spaces Z, Q, R, RT1 , and the arrow are pairwise non-homeomorphic.

Proof. Here are properties that distinguish each of the spaces from the remaining
ones: Z is discrete, Q is countable, each proper closed subset of RT1 is finite, and
any two nonempty open sets in the arrow have nonempty intersection. □

Example 3.29. We have the following examples.

(1) The function h : R → (0, 1) defined by h(x) = ex /(1 + ex ) is a homeomorphism.

Proof. It is clear that h(x) ∈ (0, 1), h is invertible and


y
h−1 (y) = ln .
1−y
From the expressions of h and h−1 , we can infer that both of them are continuous.
Thus h is a homeomorphism. □

(2) Every nondegenerate affine transformation of Rn is a homeomorphism.

Proof. Recall that an affine transformation f : Rn → Rn can be defined by

y = f (x) = Ax + b,

where A is a matrix and b a vector. The map f is said to be nondegenerate if A is


invertible, whence

x = A−1 (y − b) = A−1 (y) − A−1 (b),

which means that f is a bijection and f −1 is also a nondegenerate affine transformation.


Finally, f and f −1 are continuous, because they are given in coordinates by linear
formulas. Hence f is a homeomorphism. □

(3) ∀ R > 0, the inversion

τ : Rn \ {0} → Rn \ {0}
x
x 7→ R· 2
|x|
is a homeomorphism.

Hint. τ is an involution. Use Theorems 3.11 and 3.12 to show the continuity. □
56 D.G.L. WANG

(4) Let H = {z ∈ C : ℑz > 0} be the upper half plane. Let a, b, c, d ∈ R s.t. ad > bc.
Then the function
f: H →H
az + b
z 7→
cz + d
is a homeomorphism.

Hint. Since
 ad − bc
ℑ f (x + yi) = · y,
|cz + d|2
we have f (H) ⊆ H. On the other hand,
b − dz
f −1 (z) = −.
a − cz
Use Theorems 3.11 and 3.12 to show the continuity. □

(5) The graph of a continuous real-valued function defined on an interval is homeomorphic


to the interval.

Hint. The vertical projection determines a homeomorphism. □

Question 3.30. Let a < b. Prove that

(1) [0, 1] ∼
= [a, b];
(2) [0, 1) ∼
= [a, b) ∼
= (0, 1] ∼
= (a, b];
(3) (0, 1) ∼
= (a, b) ∼
= R;
(4) [0, 1) ∼
= [0, ∞) and (0, 1) ∼
= (0, ∞).

Proof. Use x 7→ a + (b − a)x and x 7→ tan(πx/2). □

Question 3.31. Let N = (0, 1) ∈ S 1 be the north pole of the unit circle. Prove that
S 1 \ {N } ∼
= R.

Proof. Consider the map f : (1/4, 5/4) → S 1 \ {N } defined by


f (x) = (cos 2πx, sin 2πx).
An alternative proof is as follows. Consider the map g : S 1 \ {N } → R defined by
x
g(x, y) = .
1−y
57

Its inverse is
z2 − 1
 
−1 2z
g (z) = , .
z2 + 1 z2 + 1

Homework 3.3. Let p ∈ S n . The punctured sphere S n \ {p} is homeomorphic to Rn .

Figure 22. Illustration for the stereographic projection.

Proof. Let p = (0, 0, . . . , 1) ∈ Rn+1 be the north pole of S n . Define two maps f : S n \
{p} → Rn and g : Rn → S n \ {p} by
!
x1 xn
f (x1 , . . . , xn+1 ) = , ..., , and
1 − xn+1 1 − xn+1
!
2x1 2xn |x|2 − 1
g(x1 , . . . , xn ) = , ..., , ,
|x|2 + 1 |x|2 + 1 |x|2 + 1
where |x|2 = ni=1 x2i . By Theorem 3.12, the maps f and g are continuous.
P

We claim that the map f is a homeomorphism. In fact, it is routine to check that


f ◦ g is the identity map on Rn , and that g ◦ f is the identity map on S n \ {p}. This
completes the proof. □

The map f : S n \{p} → Rn in the above proof is called the stereographic projection.
It is a particular mapping (function) that projects a sphere onto a plane; see Fig. 22. It
was known to Hipparchus (approx. BC 190–BC 120), and originally as the planisphere
projection. The term planisphere is still used to refer to celestial charts; see Fig. 23.
Example 3.32. D2 ∼
= I 2.

Proof. Let K = {(x, y) ∈ R2 : x, y ∈ [−1, 1]}. We have the homeomorphisms


I2 → K
(x, y) 7→ (2x − 1, 2y − 1)
58 D.G.L. WANG

Figure 23. A celestial map from the 17th century, by the Dutch cartog-
rapher Frederik de Wit.

and
K → D2
(x, y)
(x, y) 7→ p .
x2 + y 2
Geometrically, this means that each segment joining the origin with a point on the
contour of the square is linearly mapped to the part of the segment that lies within the
circle. □

Example 3.33. The following plane domains are homeomorphic.

(1) The whole plane R2 ;


(2) open square I 2 ;
(3) open strip {(x, y) ∈ R2 : x ∈ (0, 1)};
(4) open upper half-plane;
(5) open half-strip {(x, y) ∈ R2 : x > 0, y ∈ (0, 1)};
(6) open rectangle {(x, y) ∈ R2 : a < x < b, c < y < d};
59

(7) open quadrant: {(x, y) ∈ R2 : x, y > 0};


(8) the plane without the ray {(x, 0) ∈ R2 : x ≥ 0};
(9) open half unit ball;
(10) open sector {(x, y) ∈ R2 : x2 + y 2 < 1, x > y > 0};
(11) unit open ball;
(12) open angle: {(x, y) ∈ R2 : x > y > 0}.

Proof. Use polar coordinates for the last two. □

Example 3.34. Every simple (i.e., without self-intersections) closed polygon in R2


is homeomorphic to the circle S 1 . Every non-closed simple finite polyline in R2 is
homeomorphic to the unit interval I.

Hint. Map each edge of the polygon homeomorphically to a suitable arc of S 1 . Map
each edge of the polyline homeomorphically to a suitable part of I. □

One of the classical problems in topology is the homeomorphism problem: to find


out whether or not two given topological spaces are homeomorphic. In order to prove
that two spaces are homeomorphic, it suffices to present a homeomorphism between
them. To prove that two spaces are not homeomorphic, one may look for a property or
a characteristic shared by homeomorphic spaces s.t. one of the spaces has it, while the
other one does not.

Definition 3.35. Topological properties or topological invariants: properties


that are shared by homeomorphic spaces.

A first example of topological properties is the cardinality of the set of points, as well
as the number of open sets. Less obvious topological properties are the main object of
the forthcoming sections.

Example 3.36 (Not proved at this stage). Here are some pairs of spaces which are not
homeomorphic.

(1) Rm ∼
̸= Rn if m ̸= n.
(2) Dp ∼
̸= Dq if p ̸= q.
(3) S p ∼
̸= S q if p ̸= q.
60 D.G.L. WANG

(4) The punctured plane R2 \ {(0, 0)} is not homeomorphic to the space
{(x, y) ∈ R2 : x2 + y 2 ≥ 1}
with an open hole.

Definition 3.37. A continuous map f : X → Y is a (topological) embedding if the


submap f : X → f (X) is a homeomorphism. Denoted f : X ,→ Y . We say that X is
embeddable in Y if there is an embedding f : X → Y .

Theorem 3.38. Any inclusion is an embedding. The composition of embeddings is an


embedding. □

Question 3.39. Find topological spaces X and Y s.t. X is embeddable in Y , and Y is


embeddable in X, but X ∼
̸= Y .

Answer. Take X = I and Y = R. □

Question 3.40. Is Q embeddable in Z?

Answer. No. Z is discrete but Q is not. □

Homework 3.4. Is a discrete space embeddable in an indiscrete space? How about


vice versa?

Definition 3.41. Given metric spaces (X, dX ) and (Y, dY ), with f : X → Y .

• f is an isometric embedding if dY (f (a), f (b)) = dX (a, b) ∀ a, b ∈ X.


• f is an isometry if it is a bijective isometric embedding.

Lemma 3.42. ∀ metric spaces (X, dX ) and (Y, dY ), ∃ a metric space Z s.t. X and Y
can be isometrically embedded in Z.

Proof. Consider the set W = X ⊔ Y of disjoint union of X and Y . Fix x0 ∈ X and


y0 ∈ Y . For x, y ∈ W , we define a function

dX (x, y),

 if x, y ∈ X;
f (x, y) = dY (x, y), if x, y ∈ Y ;


dX (x, x0 ) + dY (y, y0 ) + 1, if x ∈ X and y ∈ Y .

We shall show that f is a metric on W . The conditions of being positive-definite


and symmetric are clear. It remains to show the triangle inequality for f w.r.t. three
elements in W . If the three elements are all in X, or all in Y , then the subadditivity
61

of X or Y itself works. Otherwise, we can suppose that x1 , x2 ∈ X and y ∈ Y , and it


suffices to show that
f (x1 , x2 ) + f (x1 , y) ≥ f (x2 , y) and f (x1 , y) + f (x2 , y) ≥ f (x1 , x2 ).
that is,
dX (x1 , x2 ) + dX (x1 , x0 ) ≥ dX (x2 , x0 ) and
dX (x1 , x0 ) + dX (x2 , x0 ) + 2dY (y, y0 ) + 2 ≥ dX (x1 , x2 ),
both of which hold for the triangle inequality w.r.t. the elements x0 , x1 , x2 in X. □

Definition 3.43. Isometrically embedding two metric spaces in a single one, we can
consider the Hausdorff distance between their images. The infimum of such Hausdorff
distances over all pairs of isometric embeddings of metric spaces X and Y in metric
spaces is the Gromov-Hausdorff distance between X and Y .

Figure 24. Mikhail Leonidovich Gromov (also Mikhael Gromov,


Michael Gromov or Mischa Gromov) (1943–present), is a French-Russian
mathematician known for work in geometry, analysis and group theory.
He is a permanent member of IHÉS (English: Institute of Advanced
Scientific Studies, a French institute supporting advanced research in
mathematics and theoretical physics) in Paris and a professor of Mathe-
matics at New York Univ. Gromov has won the Abel Prize in 2009 “for
his revolutionary contributions to geometry”.

Puzzle 3.44. Is the Gromov-Hausdorff distance a metric on the set of metric spaces?
62 D.G.L. WANG

4. Connectedness

4.1. Connected spaces. Being connected means, intuitively, being all in one piece.

Definition 4.1. Given (X, Ω).

• X is connected: X has only two clopen subsets, i.e., ∅ and X.


• X is disconnected: X is not connected.
• A (connected) component of X: a maximal nonempty connected subset of X.
• X is totally disconnected: X is disconnected, and every point of X is a component.

Whenever the connectedness of a specific topological space is considered, please ask


yourself “how far is the truth away from my intuition?”

Example 4.2. Examples for connected spaces.

(1) The empty set ∅.


(2) Any singleton space.
(3) Any indiscrete space.
(4) The cofinite space.
(5) The arrow.
(6) Any two-element non-discrete space.

Proposition 4.3. Any set whose interior is empty is totally disconnected.

Proof. Let A be a set in a space such that A◦ = ∅. Assume that B is a nonempty


proper clopen subset of A. Then A◦ contains B; thus A◦ ̸= ∅, a contradiction. This
proves that A is totally disconnected. □

Example 4.4. The following spaces are totally disconnected.

(1) Any discrete space with at least 2 elements.


(2) Q.
(3) R \ Q.
(4) The Cantor set.
63

Theorem 4.5. Let X be a topological space. Then the following 5 states are equivalent:

(1) X is disconnected.
(2) ∃ nonempty open sets A and B s.t. X = A ⊔ B.
(3) ∃ nonempty closed sets A and B s.t. X = A ⊔ B.
(4) ∃ nonempty clopen sets A and B s.t. X = A ⊔ B.
(5) ∃ a continuous surjection X → S 0 = {−1, 1}.

Proof. We first show that (1) to (4) are equivalent to each other. If X is disconnected,
then X has a nonempty proper clopen subset, say, A. Then the set B = Ac is nonempty
and open, and X = A ⊔ B. This proves that (1) implies (2) to (4). Conversely, if
X = A ⊔ B for some nonempty open A and B, then A = B c and B = Ac are closed.
Hence both A and B are clopen. Using these arguments, it is easy to see that each
of (2) to (4) implies (1).

(2) ⇒ (5): We can define an indicator map f : X → S 0 by



−1, if x ∈ A;
f (x) =
1, if x ∈ B.
Since A and B are nonempty, the function f is surjective. Since the preimage of every
subset of S 0 is open, we infer that f is continuous.

(5) ⇒ (4): Since the singleton {1} is a clopen subset of S 0 , the preimgae A = f −1 (1)
is clopen. Since f is surjective, we have A ̸= ∅. Along the same line, we can infer that
B = f −1 (−1) is clopen and nonempty. Moreover, X = A ⊔ B. This proves (4). □

For convenience, we call the last sufficient and necessary condition of disconnectedness
the S 0 -condition.

Theorem 4.6. The continuous image of a connected space is connected, i.e., continuous
surjection preserves the connectedness. As a consequence, the connectedness is a
topological property, and the number of components is a topological invariant.

Proof. Let f : X → Y be a continuous surjection. If Y is disconnected, by Theorem 4.5,


∃ a continuous surjection g : Y → S 0 . Since the composition g ◦ f : X → S 0 is a
continuous surjection, by Theorem 4.5, X is disconnected, a contradiction. Since
64 D.G.L. WANG

any homeomorphism is a continuous surjection, the connectedness is preserved by


homeomorphisms and thus a topological property. □

Question 4.7. Let Ω1 , Ω2 be topologies on a set X s.t. Ω1 ⊆ Ω2 .

1) Does the connectedness of (X, Ω1 ) imply the connectedness of (X, Ω2 )?

Answer. No. For instance, take Ω1 to be indiscrete and take Ω2 to be discrete. □

2) Does the connectedness of (X, Ω2 ) imply the connectedness of (X, Ω1 )?

Answer. Yes. Assume that (X, Ω1 ) is disconnected. Then (X, Ω1 ) has a nonempty
proper clopen subset, say, A. Since Ω1 ⊆ Ω2 , the set A is clopen in Ω2 . Hence (X, Ω2 )
is disconnected. □

When A ⊆ M ⊆ X, it is safe to say that “A is connected” without specifying


the ambient topology (ΩM or ΩX ) from which ΩA is induced. This is because the
connectedness of a space relies only on the topology of that space, while the topology
of A induced from ΩM coincides with that from ΩX ; see Theorem 2.18.

Theorem 4.8. Clopen sets “separate” connected sets. In other words, if A is clopen in
X, and if C is connected in X, then either C ⊆ A or C ⊆ X \ A.

Proof. If not, then C ∩ A is a nonempty clopen subset of C, contradicting the connect-


edness of C. □

Theorem 4.9. Let Γ be a family of connected sets. If a connected set C meets every
member in Γ, then ∪A∈Γ A ∪ C is connected.

Proof. Let X = ∪A∈Γ A ∪ C. Suppose that X = U ⊔ V where U and V are clopen. By


Theorem 4.8, the connected set C is contained in either U or V . W.l.o.g. we can suppose
that C ⊆ U . For any A ∈ Γ, since A ∩ U ̸= ∅, we infer that A ⊆ U by Theorem 4.8.
Hence
[
V =V ∩X = (V ∩ A) = ∅.
A∈Γ

This proves the connectedness of X. □

An alternative proof of Theorem 4.9. If X is disconnected, by Theorem 4.5, ∃ a contin-


uous surjection f : X → S 0 . Let A ∈ Γ. Let xA ∈ C ∩ A. W.l.o.g., we can suppose that
65

f (xA ) = 1. Since xA ∈ C, we obtain f (x) = 1 for all x ∈ C by Theorem 4.6. Since


xA ∈ A, we obtain that f (x) = 1 for all x ∈ A by Theorem 4.6. Therefore,

f (x) = 1 ∀ x ∈ C ∪ A.

For any other A′ ∈ Γ, let x′ ∈ C ∩ A′ . Since x′ ∈ C, we have f (x′ ) = 0. By Theorem 4.6,


we find
f (x) = 1 ∀x ∈ A′ .
This proves that f (x) = 1 ∀ x ∈ X, contradicting that f is a surjection. □

Corollary 4.10. Let {Ak }k∈Z be a family of connected sets s.t. Ak meets Ak+1 , ∀ k.
Then the union ∪k∈Z Ak is connected.

Proof. By Theorem 4.9, we can show by induction that the set Bn = ∪nk=−n Ak is
connected, ∀ n ∈ N. Using Theorem 4.9 again, we obtain the connectedness of the
union ∪k∈Z Ak = ∪n∈N Bn . □

Theorem 4.11. Any point in a topological space is contained in a unique component. As


a consequence, two connected components either are disjoint or coincide. In particular,
any connected space has only one component.

Proof. The union of connected components which contains p is the unique component.
It is connected by Theorem 4.9. □

Theorem 4.12. A subset of R is connected ⇐⇒ it is an interval.

Proof. Let A ⊆ R. If A is not an interval, then ∃ x, z ∈ A and y ∈ Ac s.t. x < y < z.


Then ?? holds true, and the set A is disconnected. Below we present two proofs for the
connectedness of any interval.

By the bisection method. Assume that an interval J is disconnected. Then ∃


nonempty closed sets U and V s.t. J = U ⊔ V . W.l.o.g., we can suppose that ∃ a0 ∈ U ,
b0 ∈ V s.t. a0 < b0 . For n ∈ N, we write cn = (an + bn )/2 and define

(a , c ), if c ∈ V ;
n n n
(an+1 , bn+1 ) =
(cn , bn ), if cn ∈ U .

We obtain a sequence an in U and a sequence bn in V . It is clear that an increases


and bn decreases, and that sup an = inf bn . Since both U and V are closed, this limit
lies in U ∩ V , a contradiction.
66 D.G.L. WANG

By the S 0 -condition. Assume that A is a disconnected interval. By Theorem 4.5, ∃


a continuous surjection f : A → S 0 . Then
G
f −1 (1) = A ∩ J, and
J∈Γ+
G
f −1 (−1) = A ∩ J,
J∈Γ−

for some nonempty collections Γ+ and Γ− of disjoint open intervals, and

J− ∩ J+ = ∅, ∀ J+ ∈ Γ+ and J− ∈ Γ− .

It follows that
G
A = f −1 (−1) ∪ f −1 (1) = A ∩ J.
J∈ Γ− ∪Γ+

Since the collection Γ− ∪ Γ+ contains at least two disjoint open intervals, the set A is
not an interval, a contradiction. □

Corollary 4.13. Let X be a connected space. Let f : X → R be a continuous function.


Then the image f (X) is an interval of R.

Proof. By Theorem 4.6 and Theorem 4.12. □

Corollary 4.14. Any continuous function f : R → R maps intervals to intervals. □

Corollary 4.15 (Intermediate Value Theorem). Any continuous function

f : [a, b] → R

takes every value between f (a) and f (b). □

Definition 4.16. A set A ⊆ Rn is convex: ∀ a, b ∈ A, we have

[a, b] = {x ∈ Rn : x = (1 − t)a + tb for some t ∈ I} ⊆ A.

Theorem 4.17. Any convex set in Rn is connected. In particular, the sets Rn , B n and
Dn are connected.

Proof. Let A be a convex set in Rn with x0 ∈ A. Then the interval [a, x0 ] is connected
for any a ∈ A. Then the union A = ∪a∈A [a, x0 ] is connected by Theorem 4.9. □

Question 4.18. Find all connected subspaces A, with |A| ≥ 2,

1) of the arrow;
67

Answer. Any subset. Assume that A ⊆ [ 0, ∞) is disconnected. By Theorem 4.5,


∃ a continuous surjection f : A → S 0 . Note that any open set in A consists of the
elements of A that are larger than some number M , and that any closed set in A
consists of the elements of A that are smaller than or equal to some number m.
Therefore, the clopen set f −1 (1) in A has to be either the empty set or the whole set
A, contradicting that f is surjective. □

2) of the cofinite space;

Answer. Any infinite subset. On one hand, any finite set A with |A| ≥ 2 is the
union of two nonempty finite subsets. Since a subset is finite ⇐⇒ it is closed. By
Theorem 4.5, we infer that A is disconnected. On the other hand, assume to the
contrary that an infinite set A is disconnected. By Theorem 4.5, ∃ a continuous
surjection f : A → S 0 . Then each of the preimages f −1 (±1) is closed in A. Hence
A = f −1 (−1) ∪ f −1 (1) is finite, a contradiction. □

Theorem 4.19. Every open set of R has a countable number of components.

Proof. Every open set of R is a union of disjoint open intervals, each of which contains
a representative rational point. □

Homework 4.1. Let A0 ⊇ A1 ⊇ · · · be an infinite decreasing sequence of closed


connected sets in the plane R2 . Is ∩n≥0 An connected?

Theorem 4.20. Suppose that both A ∪ B and A ∩ B are connected. Then both of A
and B are connected if either both A and B are open or both A and B are closed.

Proof. Consider the ambient space C = A ∪ B. From premise, C is connected. By the


symmetry position of A and B, it suffices to show that A is connected. Assume to the
contrary that A is disconnected. Then

A = U ⊔ V,

where U and V are nonempty clopen sets in A. Since A ∩ B is a connected subset of A,


we infer that either A ∩ B ⊆ U or A ∩ B ⊆ V . W.l.o.g., we suppose that A ∩ B ⊆ U .
Since A ∪ B is connected, in order to obtain a contradiction, it suffices to show that V
is clopen.

In fact, if both A and B are open (resp., closed), then both U and V are open (resp.,
closed) by Theorem 2.16. On the other hand, the complement V c = U ∪ B is the union
of open (resp., closed) sets and also open (resp., closed). □
68 D.G.L. WANG

Remark 4.21. The condition of being open or closed is necessary. This can be seen by
the example A = Q and B = R \ Q.

Theorem 4.22. If A is connected and everywhere dense in X, then X is connected.

Proof. Let X = U ⊔ V where U and V are open. Then


A = A ∩ X = A ∩ (U ⊔ V ) = (A ∩ U ) ⊔ (A ∩ V ).
It follows from the connectedness of A that either U ∩ A = ∅ or V ∩ A = ∅. By
Theorem 2.36, any everywhere dense set meets every nonempty open set. Therefore
U = ∅ or V = ∅. This proves the connectedness of X. □

An alternative proof of Theorem 4.22. Suppose not. By Theorem 4.5, ∃ a continuous


surjection f : X → S 0 . By Theorem 4.6, we can suppose w.l.o.g. that f (a) = 1 ∀ a ∈ A.
By Theorem 2.36, the open set f −1 (−1) meets A, a contradiction. □

Corollary 4.23. If A is connected and A ⊆ B ⊆ A, then B is connected. In particular,


the closure A is connected.

Proof. Since A is everywhere dense in A, we infer that A is connected. Since the set A
is everywhere dense in B. Thus B is connected by Theorem 4.22. □

Corollary 4.24. Any component of a topological space is closed.

Proof. Let C be a component. By Corollary 4.23, the closure C is connected. By


Theorem 4.11, C = C is closed. □

Example 4.25. S n is connected.

Proof. By Homework 3.3 and Theorem 4.17, we have S n \ {0} ∼


= Rn is connected. By
Corollary 4.23, S n is connected. □

Question 4.26. Are the following subsets of R2 connected?

(1) The set Q2 = {(x, y) ∈ R2 : x, y ∈ Q}.

Answer. No. Any projection is continuous; the continuous image of any connected
set is connected; the projection of Q2 onto one coordinate is Q, which is totally
disconnected. □

(2) The set {(x, y) ∈ R2 : {x, y} ∩ Q ̸= ∅}.


69

Answer. Yes! Any two points are joined by a polyline with at most 3 segments. □

(3) The set Q2 ∪ (Qc )2 , where (Qc )2 = {(x, y) ∈ R2 : x, y ̸∈ Q}.

Answer. Yes! Let lk = {(x, kx) : x ∈ R} and X = ∪k∈Q\{0} lk . Then each line lk is
connected. It is clear that X ⊂ Q2 ∪ (Qc )2 ⊂ X = R2 . By Corollary 4.23, the set
Q2 ∪ (Qc )2 is connected. □

Simple constructions assigning homeomorphic spaces to homeomorphic ones, e.g.,


deleting one or several points.

Example 4.27. R1 ̸∼
= Rn for n > 1. □

Example 4.28. The following spaces are pairwise non-homeomorphic:

I, [ 0, 1), R, S 1.

Proof. The number of distinct points removing each of which yields a connected space
for I, [0, 1), R and S 1 are pairwise distinct, i.e., 2, 1, 0 and ∞ resp. □

Homework 4.2. Let S be the set of points on the spiral


!
1
r = exp , ϕ ≥ 0,
1 + ϕ2

in the polar representation; see Fig. 25. We know that S is connected. Let T ⊆ S 1 .
Prove or disprove that S ∪ T is connected.

Figure 25. The spiral (r, ϕ).


70 D.G.L. WANG

Puzzle 4.29. Give a topological classification of the letters of the alphabet:

A, B, C, D, E, F, G, H, I, J, K, L, M, N, O, P, Q, R, S, T, U, V, W, X, Y, Z,

regarded as subsets of R2 , where the arcs comprising the letters are assumed to have
zero thickness.

Hint. The answer depends on the graphics of the letters. A possible answer is as follows:
There are nine classes for this problem.

(1) A and R;
(2) B;
(3) C, G, I, J, L, M, N, S, U, V, W, Z;
(4) D and O;
(5) E, F, T, Y ;
(6) H and K;
(7) P ;
(8) Q;
(9) X.

Theorem 4.30 (Extendability of connected sets). We have the following.

1) If A and B are connected and A ∩ B ̸= ∅, then A ∪ B is connected.

Proof. See Homework 4.3. □

2) Let A be a connected subset of a connected space X. Then A ∪ B is connected if B


is a clopen subset of Ac .

Proof. Let B be a clopen subset of Ac and let C = A ∪ B. Assume to the contrary


that C is disconnected. Then ∃ open sets U and V s.t.

C ⊆ U ∪ V, U ∩ C ̸= ∅, V ∩ C ̸= ∅, and U ∩ V ∩ C = ∅.

Let W = U ∪ V . By Theorem 2.16, both the sets U and V are clopen in W . Since
A is connected, A ⊆ W , we infer from Theorem 4.8 that either A ⊆ U or A ⊆ V .
71

W.l.o.g., let us suppose that A ⊆ U . Let


D = V ∩ C.
Then D ̸= ∅. We shall show that D is a clopen subset of X, which yields an
immediate contradiction to the premise that X is connected.
In fact, consider the topological space Ac with a subset V . Since B is open in Ac
from premise, the intersection D = V ∩ B is an open subset of V . Since V is open
in X, we infer by Theorem 2.16 that D is open in X. On the other hand, consider
the topological space Ac with a subset U c . Since B is closed in Ac from premise,
the intersection D = U c ∩ B is closed in U c . Since U c is closed in X, we infer by
Theorem 2.16 that D is closed in X. This completes the proof. □

Homework 4.3. If A and B are connected and A ∩ B ̸= ∅, then A ∪ B is connected.


72 D.G.L. WANG

Appendix A. Some elementary inequalities

Theorem A.1 (Jensen’s inequality). Let x1 , . . . , xn ∈ R be in the domain of a real


function f . Let wi ≥ 0 s.t. ni=1 wi = 1. Then we have
P

n
! n
X X
f w i xi ≤ wi f (xi ), if f is convex and
i=1 i=1
n
! n
X X
f w i xi ≥ wi f (xi ), if f is concave,
i=1 i=1

where each of the equalities hold if and only if either x1 = x2 = · · · = xn , or the function
f is linear.

Theorem A.2 (Weighted power mean inequality). Let p > 0. Suppose that wi ≥ 0
s.t. ni=1 wi = 1. Denote by
P
n
!1/p
X
Mp = wi xpi
i=1
the p-th weighted power mean of positive numbers xi > 0. Then the function Mp is
increasing in p, i.e., Mp ≤ Mq whenever p < q, where the equality holds if and only if
x1 = x2 = · · · = xn .

Proof. Consider the function f (x) = xq/p . Since


!
q q
f ′′ (x) = · − 1 · xq/p−2 > 0,
p p
the function f (x) is convex. By Jensen’s inequality we find
n
!q/p n
! n n
X p
X p
X p
X
w i xi =f w i xi ≤ wi f (xi ) = wi xqi .
i=1 i=1 i=1 i=1

Raising both sides to the power of 1/q yields the desired inequality. The equality holds
if and only if all the numbers xi are equal because that is the equality condition for
Jensen’s inequality. □

Definition A.3. Let H = {x ∈ R : x ≥ 1} ∪ {+∞}.

• Hölder conjugate of p ∈ H: q ∈ H s.t. 1/p + 1/q = 1, i.e., p + q = pq.

Lemma A.4 (Young’s inequality for products). Let p, q > 1 be Hölder conjugates
of each other. Suppose that a, b ≥ 0. Then
ab ≤ ap /p + bq /q,
73

where the equality holds if and only if ap = bq .

Proof. Let t = 1/p. Then 1 − t = 1/q. By using the concavity of the logarithm function,
we infer that

log tap + (1 − t)bq ≥ t log ap + (1 − t) log(bq ) = log(ab),


 

with the equality holds if and only if ap = bq . Young’s inequality follows by exponenti-
ating. □

Theorem A.5 (Hölder’s inequality). Let p, q > 1 be Hölder conjugates of each other.
For any x = (x1 , . . . , xn ) ∈ Cn and y = (y1 , . . . , yn ) ∈ Cn , we have
n
X
|xi yi | ≤ ∥x∥p · ∥y∥q ,
i=1

where the equality holds if and only if the vectors

xp = (xp1 , . . . , xpn ) and y q = (y1q , . . . , ynq )

are linearly dependent.

Proof. It is w.l.o.g. to suppose that ∥x∥p = ∥y∥q = 1. Lemma A.4 gives

(4) |xi yi | ≤ |xi |p /p + |yi |q /q

for any i, where the equality holds if and only if |xi |p = |yi |q . Adding Ineq. (4) for
P
i = 1, . . . , n, we obtain |xi yi | ≤ 1 as desired. □

The p = 2 case of Hölder inequality is the Cauchy-Schwarz inequality.

Theorem A.6 (Minkowski’s inequality). For x, y ∈ Rn and p ≥ 1, we have

∥x + y∥p ≤ ∥x∥p + ∥y∥p .

Proof. If p = 1, then Minkowski’s inequality reduces to the subadditivity of the absolute


value function. If ∥x + y∥p = 0, then Minkowski’s inequality holds for the nonnegativity
of any norm. Below we can suppose that

p>1 and ∥x + y∥p > 0.

Let q be the Hölder conjugate of p. Then (p − 1)q = p and 1/q = (p − 1)/p. Let

z = (|x1 + y1 |p−1 , . . . , |xn + yn |p−1 ).


74 D.G.L. WANG

Then we have
X 1/q X (p−1)/p
(p−1)q p
∥z∥q = |xi + yi | = |xi + yi | = ∥x + y∥p−1
p .

By Theorem A.5, we can deduce that


X X X
∥x + y∥pp = |xi + yi |p ≤ |xi | · |xi + yi |p−1 + |yi | · |xi + yi |p−1
≤ ∥x∥p · ∥z∥q + ∥y∥p · ∥z∥q
= (∥x∥p + ∥y∥p ) · ∥x + y∥p−1
p .

By canceling the positive factor ∥x + y∥p−1


p we obtain Minkowski’s inequality. □

Appendix B. Subadditivity and superadditivity

A function f : A → B is subadditive if both A and B are closed under addition and


f (x + y) ≤ f (x) + f (y) for all x, y ∈ A.

For example, the square root function · is subadditive:
√ √ √
x + y ≤ x + y for all x, y ≥ 0.
A function f : A → B is superadditive if both A and B are closed under addition and
f (x + y) ≥ f (x) + f (y) for all x, y ∈ A.
For example, the determinant is superadditive for nonnegative Hermitian matrix, that
is, if A and B are nonnegative Hermitian matrices over C, then
det(A + B) ≥ det(A) + det(B).

In particular, a subadditive sequence is a sequence {an }n≥0 such that


ai+j ≤ ai + aj for all i, j ≥ 0.
A superadditive sequence is defined in the same manner.

Lemma B.1 (Fekete’s subadditive lemma). For any subadditive sequence {an }n≥0 , the
limit limn→∞ an /n exists and is equal to the infimum inf an /n, which might be −∞.

School of Mathematics and Statistics, Beijing Institute of Technology, 102400


Beijing, P. R. China

Email address: kwgl@icloud.com


75

Figure 26. Michael Fekete (1886–1957) was a Hungarian-Israeli mathe-


matician. In 1955, he was awarded the Israel Prize for exact sciences.

You might also like