Lecture Notes On Topology 6
Lecture Notes On Topology 6
Huỳnh Quang Vũ
Current address: Faculty of Mathematics and Computer Science, University
of Science, Vietnam National University, 227 Nguyen Van Cu, District 5, Ho Chi
Minh City, Vietnam
E-mail address: [email protected]
URL: http://www.math.hcmus.edu.vn/∼hqvu
ABSTRACT. This is a set of lecture notes prepared for a series of introductory
courses in Topology for undergraduate students at the University of Science,
Vietnam National University–Ho Chi Minh City. It is written to be delivered
by a lecturer, namely by myself, tailored to the need of my own students. I did
not write it with self-study readers or other lecturers in mind.
In my experience many things here are much better explained in oral form
than in written form. Therefore in writing these notes I intend that more expla-
nations and discussions will be carried out in class. I hope by presenting only
the essentials these notes will be more suitable for classroom use. Some details
are left for students to fill in or to be discussed in class.
Since students in my department are required to take a course in Functional
Analysis, I try not to duplicate material in that course.
In general if you encounter an unfamiliar notion, look at the Index and the
Contents. A sign * in front of a problem notifies the reader that this is an im-
portant problem (which clarifies understranding or will be used later) although
might not appear to be so at first, so the reader should try it.
This set of lecture notes will be continuously developed (and corrected!). I
intend to keep the latest version freely available on my web page.
May 14, 2005 – July 10, 2011.
Contents
Bibliography 109
Index 111
3
CHAPTER 1
General Topology
1.1. Set
In General Topology we often work in very general settings, in particular we
often deal with infinite sets.
We will not define what a set is. That means we only work on the level of the
so-called “naive set theory”.
Even so we should be aware of certain problems in naive set theory. Until the
beginning of the 20th century, the set theory of George Cantor, in which set is
not defined, was thought to be a good basis for mathematics. Then some critical
problems were discovered.
Example 1.1.1 (Russell’s paradox). In 1901 Bertrand Russell discovered the fol-
lowing paradox. Consider the set S = {x | x ∈ / x} (the set of all sets which are
not members of themselves). Then whether S ∈ S or not is undecidable, because
answering yes or no to this question leads to contradiction. 1
This example shows that we were having too much liberty with the notion of
set. Axiomatic systems for the theory of sets have been developed since then. In the
Von Neumann-Bernays-Godel system a more general notion than set, called class
(lớp), is used. See [Dug66, p. 32].
In this course, in occasions where we deal with “set of sets” we often replace
the term “set” by the terms “class” or “collection” (họ).
Equivalent sets. Two sets are said to be equivalent if there is a bijection from
one to the other. When A and B are equivalent sometimes we write A ∼ B.
1A famous version of this paradox is the Barber paradox: In a village there is a barber. His job is
to do hair cut for a villager if and only if the villager does not cut his hair himself. The question is:
Does the barber cut his own hair?
1
2 1. GENERAL TOPOLOGY
Example 1.1.2. Two intervals [a, b] and [c, d] on the real number line are equiva-
lent. Similarly, two intervals (a, b) and (c, d) are equivalent. The bijection can be
d−c
given by a linear map x 7→ b−a (x − a) + c.
The interval (−1, 1) is equivalent to R via a map related to the tan function:
x
x 7→ √ .
1 − x2
1
x
0
1
−1
Countable sets.
Definition 1.1.3. A set is called countably infinite (vô hạn đếm được) if it is equiv-
alent to the set of all positive integers.
A set is called countable if it is either finite or countably infinite.
..
.
There are real numbers whose decimal presentations are not unique, such as
1
2 = 0.5000 . . . = 0.4999 . . . . Choose a number b = 0.b1 b2 b3 . . . such that bn 6= 0, 9
and bn 6= an,n . Choosing bn differing from 0 and 9 will guarantee that b 6= an
for all n (see more at 1.1.31). Thus the number b is not in the above table, a
contradiction.
Order. An order (thứ tự) on the set S is a relation R on S that is reflexive, anti-
symmetric and transitive.
Note that two arbitrary elements a and b do not need to be comparable; that is,
the pair (a, b) may not belong to R. For this reason an order is often called a partial
order.
When (a, b) ∈ R we often write a ≤ b. When a ≤ b and a 6= b we write a < b.
If any two elements of S are related then the order is called a total order (thứ
tự toàn phần) and (S, ≤) is called a totally ordered set.
Example 1.1.10. The set R of all real numbers with the usual order ≤ is totally
ordered.
Example 1.1.12 (Dictionary order). Let (S1 , ≤1 ) and (S2 , ≤2 ) be two ordered
sets. The following is an order on S1 × S2 : (a1 , b1 ) ≤ (a2 , b2 ) if (a1 < a2 ) or ((a1 =
a2 ) ∧ (b1 ≤ b2 )). This is called the dictionary order (thứ tự từ điển).
In an ordered set, the smallest element (phần tử nhỏ nhất) is the element that is
smaller than all other elements. More concisely, if S is an ordered set, the smallest
element of S is an element a ∈ S such that ∀b ∈ S, a ≤ b. The smallest element, if
exists, is unique.
A minimal element (phần tử cực tiểu) is an element which no element is smaller
than. More concisely, a minimal element of S is an element a ∈ S such that ∀b ∈
S, b ≤ a ⇒ b = a. A minimal element, if exists, may not be unique.
A lower bound (chặn dưới) of a subset of an ordered set is an element of the set
that is smaller than or equal to any element of the subset. More concisely, if A ⊂ S
then a lower bound of A in S is an element a ∈ S such that ∀b ∈ A, a ≤ b.
The definitions of largest element, maximal element, and upper bound are sim-
ilar.
|A| = |B| ⇐⇒ (A ∼ B)
PROOF. It comes from that any infinite set contains a countably infinite sub-
set, and 1.1.9.
Georg Cantor put forward the Continuum hypothesis: There is no cardinal be-
tween ℵ0 and c.
Theorem 1.1.15 (No maximal cardinal). The cardinal of a set is strictly less than
the cardinal of the set of all of its subsets, i.e. |A| < |2A |.
This implies that there is no maximal cardinal. There is no “universal set”, “set
which contains everything”, or “set of all sets”.
PROOF. Let A 6= 0/ and denote by 2A the set of all of its subsets.
(1) |A| ≤ |2A |: The map from A to 2A : a 7→ {a} is injective.
6 |2A |: Let φ be any map from A to 2A . Let X = {a ∈ A | a ∈
(2) |A| = / φ (a)}.
Suppose that there is x ∈ A such that φ (x) = X. Then the question whether
x belongs to X or not is undecidable. Therefore φ is not surjective.
There are cases where this axiom could be avoided. For example in the proof
of 1.1.6 we used the well-ordered property of Z+ instead. See for instance [End77,
p. 151] for further material on this subject.
Zorn lemma is often a convenient form of the Axiom of choice.
Cartesian product. Let {Ai }i∈I be a family of sets indexed by a set I. The Carte-
sian product (tích Decartes) ∏i∈I Ai of this family is defined to be the collection of
all maps a : I → i∈I Ai such that for each i ∈ I we have a(i) ∈ Ai . The statement
S
Problems.
1.1.23. Show that two planes with finitely many points removed are equivalent.
Hint: Use 1.1.22.
1.1.24. Give another proof of Theorem 1.1.7 by checking that the map Z+ × Z+ → Z+ ,
(m, n) 7→ 2m 3n is injective.
1.1.31. This result is used in 1.1.9. Show that any real number could be written in base d
with any d ∈ Z, d ≥ 2. However two forms in base d could represent the same real number,
as seen in 1.1.9. This happens only if starting from certain digits, all digits of one form are
0 and all digits of the other form are d − 1.
1.1.32 (2ℵ0 = c). We prove that 2N is equivalent to R.
(1) Show that 2N is equivalent to the set of all sequences of binary digits.
(2) Using 1.1.31, deduce that |[0, 1]| ≤ |2N |.
(3) Consider a map f : 2N → [0, 2], for each binary sequence a = a1 a2 a3 · · · define
f (a) as follows. If starting from a certain digit, all digits are 1, then let f (a) =
1.a1 a2 a3 · · · . Otherwise let f (a) = 0.a1 a2 a3 · · · . Show that f is injective.
Deduce that |2N | ≤ |[0, 2]|.
1.1.33 (R2 is equivalent to R). * Here we prove that R2 is equivalent to R, in other words,
a plane is equivalent to a line. As a corollary, Rn is equivalent to R.
(1) First method: Construct a map from [0, 1) × [0, 1) to [0, 1) as follows. In view
of 1.1.31, we only allow decimal presentations in which not all digits are 9 start-
ing from a certain digit. The pair of two real numbers 0.a1 a2 . . . and 0.b1 b2 . . .
corresponds to the real number 0.a1 b1 a2 b2 . . .. Check that this map is injective.
(2) Second method: Construct a map from 2N × 2N to 2N as follows. The pair of
two binary sequences a1 a2 . . . and b1 b2 . . . corresponds to the binary sequence
a1 b1 a2 b2 . . .. Check that this map is injective. Then use 1.1.32.
Note: In fact for all infinite cardinal ω we have ω 2 = ω, see [Dug66, p. 52], [Lan93, p. 888].
1.1.34 (Transfinite induction principle). An ordered set S is well-ordered (được sắp tốt)
if every non-empty subset A of S has a smallest element, i.e. ∃a ∈ A, ∀b ∈ A, a ≤ b.
For example with the usual order, N is well-ordered while R is not.
Ernst Zermelo proved in 1904 that any set can be well-ordered, based on the Axiom
of choice.
The following is a generalization of the Principle of induction.
Let A be a well-ordered set. Let P(a) be a statement whose truth depends on a ∈ A. If
(1) P(a) is true when a is the smallest element of A
(2) if P(a) is true for all a < b then P(b) is true
then P(a) is true for all a ∈ A.
Example 1.2.2. On any set X there is the trivial topology (tôpô hiển nhiên) {0,
/ X}.
On any set X there is the discrete topology (tôpô rời rạc) whereas any point
constitutes an open set. That means any subset of X is open, so the topology is 2X .
Thus on a set there can be many topologies.
Proposition 1.2.4. The statement “intersection of finitely many open sets is open”
is equivalent to the statement “intersection of two open sets is open”.
Example 1.2.5 (Metric space). Metric spaces are important examples of topolog-
ical spaces.
Recall that, briefly, a metric space is a set equipped with a distance between
every two points. Namely, a metric space is a set X with a map d : X × X 7→ R such
that for all x, y, z ∈ X:
(1) d(x, y) ≥ 0 (distance is non-negative),
10 1. GENERAL TOPOLOGY
The topology generated by this metric is called the Euclidean topology (tôpô
Euclid) of Rn .
Example 1.2.8 (Finite complement topology). The finite complement topology
on X consists of the empty set and all subsets of X whose complements are finite.
Proposition 1.2.9 (Dual description of topology). In a topological space X:
(1) 0/ and X are closed.
(2) A finite union of closed sets is closed.
(3) An intersection of closed sets is closed.
The above statement gives a dual (đối ngẫu) description of topology in term of
closed set instead of open set.
1.2. TOPOLOGICAL SPACES 11
Bases of a topology.
Definition 1.2.10. Given a topology, a collection of open sets is a basis (cơ sở) for
that topology if every non-empty open set is a union of members of that collection.
More concisely, let τ be a topology of X, then a collection B ⊂ τ is called a
basis for τ if any non-empty member of τ is a union of members of B.
So a basis of a topology is a subset of the topology that generates the entire
topology via unions. Thus specifying a basis is a more “economical way” to de-
scribe a topology.
Example 1.2.11. In a metric space the collection of all balls is a basis for the
topology.
Definition 1.2.13. A collection S ⊂ τ is called a subbasis (tiền cơ sở) for the topol-
ogy τ if the collection of finite intersections of members of S is a basis for τ.
Example 1.2.14. Let X = {1, 2, 3}. The topology τ = {0, / {1, 2}, {2, 3}, {2}, {1, 2, 3}}
has a basis {{1, 2}, {2, 3} {2}} and a subbasis {{1, 2}, {2, 3}}.
Example 1.2.15. The collection of all open rays, that are, sets of the forms (a, ∞)
and (−∞, a), is a subbasis for the Euclidean topology of R.
The collection of all open half-planes (meaning not consisting the separating
lines) is a subbasis for the Euclidean topology of R2 .
Comparing topologies.
Example 1.2.17. On a set the trivial topology is the coarsest topology and the
discrete topology is the finest one.
4Be careful that not everyone uses this convention. For instance Kelley [Kel55] uses this convention
but Munkres [Mun00] requires a neighborhood to be open.
12 1. GENERAL TOPOLOGY
Generating topologies. Suppose that we have a set and we want certain subsets
of that set to be open, how do find a topology for that purpose?
PROOF. To check that τ is a topology we only need to check that the inter-
section of two unions of finite intersections of members of S is a union of finite
intersections of members of S.
Let A and B be two collections of finite intersections of members of S. We have
( C∈A C) ∩ ( D∈B D) = C∈A, D∈B (C ∩ D). Since each C ∩ D is a finite intersection
S S S
Example 1.2.20. Let X = {1, 2, 3, 4}. The set {{1}, {2, 3}, {3, 4}} generates the
topology {0,
/ {1}, {3}, {1, 3}, {2, 3}, {3, 4}, {1, 2, 3}, {1, 3, 4}, {2, 3, 4}, {1, 2, 3, 4}}.
A basis for this topology is {{1}, {3}, {2, 3}, {3, 4}}.
Example 1.2.21 (Ordering topology). Let (X, ≤) be a totally ordered set with at
least two elements. The collection of subsets of the forms {β ∈ X | β < α} and
{β ∈ X | β > α} generates a topology on X, called the ordering topology.
Example 1.2.22. The Euclidean topology on R is the ordering topology with re-
spect to the usual order of real numbers. (This is just a different way to state
1.2.15.)
5In some textbooks to avoid adding the element X to S it is required that the union of all members of
S is X.
6In some textbooks to avoid adding the element X to B it is required that the union of all members of
B is X.
1.2. TOPOLOGICAL SPACES 13
Problems.
1.2.23. A collection B of open sets is a basis if for each point x and each open set O
containing x there is a U in B such that U contains x and U is contained in O.
1.2.24. Show that two bases generate the same topology if and only if each member of one
basis is a union of members of the other basis.
1.2.25. In a metric space the set of all balls with rational radii is a basis for the topology.
1
1.2.26. In a metric space the set of all balls with radii 2m , m ≥ 1 is a basis.
1.2.27 (Rn has a countable basis). * The set of all balls each with rational radius whose
center has rational coordinates forms a basis for the Euclidean topology of Rn .
1.2.29. Let d1 and d2 be two metrics on X. If there are α, β > 0 such that for all x, y ∈ X,
αd1 (x, y) < d2 (x, y) < β d1 (x, y) then the two metrics are said to be equivalent. Show that
two equivalent metrics generate the same topology.
Hint: Show that a ball in one metric contains a ball in the other metric with the same center.
1.2.30 (All norms in Rn generate the Euclidean topology). In Rn denote by k·k2 the
Euclidean norm, and let k·k be any norm.
(1) Check that the map x 7→ kxk from (Rn , k·k2 ) to R is continuous.
(2) Let Sn be the unit sphere under the Euclidean norm. Show that the restriction of
the map above to Sn has a maximum value β and a minimum value α. Hence
x
α ≤ kxk ≤ β for all x 6= 0.
2
Deduce that any two norms in Rn are equivalent, hence all norms in Rn generate the Eu-
clidean topology.
1.2.31. Is the Euclidean topology on R2 the same as the ordering topology on R2 with
respect to the dictionary order? If it is not the same, can the two be compared?
1.2.33. The collection of intervals of the form [a, b) generates a topology on R. Is it the
Euclidean topology?
1.2.35. * On the set of all integer numbers Z, consider all arithmetic progressions
Sa,b = a + bZ,
where a ∈ Z and b ∈ Z+ .
(1) Show that these sets form a basis for a topology on Z.
14 1. GENERAL TOPOLOGY
(2) Show that with this topology each set Sa,b is both open and closed.
(3) Show that the set {±1} is closed.
(4) Show that if there are only finitely many prime numbers then the set {±1} is
open.
(5) Conclude that there are infinitely many prime numbers.
1.3. CONTINUITY 15
1.3. Continuity
Continuous function.
Theorem 1.3.3. A map is continuous if and only if the inverse image of an open
set is an open set.
Proposition 1.3.5. A map is continuous if and only if the inverse image of a closed
set is a closed set.
Example 1.3.6 (Metric space). Let (X, d1 ) and (Y, d2 ) be metric spaces. Recall
that in the theory of metric spaces, a map f : (X, d1 ) → (Y, d2 ) is continuous at
x ∈ X if and only if
In other words, given any ball B( f (x), ε) centered at f (x), there is a ball B(x, δ )
centered at x such that f brings B(x, δ ) into B( f (x), ε).
It is apparent that this definition is equivalent to the definition of continuity in
topological spaces where the topologies are generated by the metrics.
In other words, if we look at a metric space as a topological space then continu-
ity in the metric space is the same as continuity in the topological space. Therefore
we inherit all results concerning continuity in metric spaces.
16 1. GENERAL TOPOLOGY
Example 1.3.7. Any two open intervals in the real number line R under the Eu-
clidean topology are homeomorphic.
Roughly speaking, in the field of Topology, when two spaces are homeomor-
phic they are the same.
1.3.10. * Suppose that f : X → Y and S is a subbasis for the topology of Y . Show that f is
continuous if and only if the inverse image of any element of S is an open set in X.
1.3.11. Define an open map to be a map such that the image of an open set is an open set.
A closed map is a map such that the image of a closed set is a closed set.
Show that a homeomorphism is both an open map and a closed map.
1.3.14. * Suppose that X is a normed space. Prove that the topology generated by the norm
is exactly the coarsest topology on X such that the norm and the translations (maps of the
form x 7→ x + a) are continuous.
18 1. GENERAL TOPOLOGY
1.4. Subspace
Subspace topology. Let (X, τ) be a topological space and let A be a subset of
X. The subspace topology on A, also called the relative topology (tôpô tương đối),
is defined to be the collection {A ∩ O | O ∈ τ}. With this topology we say that A is
a subspace (không gian con) of X.
Thus a subset of a subspace A of X is open in A if and only if it is a restriction
of a open set in X to A.
Example 1.4.5. The Euclidean line R can be embedded in the Euclidean plane R2
as a line in the plane.
Example 1.4.7. On the Euclidean line R, consider the subspace A = [0, 1) ∪ {2}.
Its interior is intA = (0, 1), the closure is clA = [0, 1] ∪ {2}, the boundary is ∂ A =
{0, 1, 2}, the set of all limit points is [0, 1].
Problems.
1.4.11. Let X be a topological space and let A ⊂ X. Then the subspace topology on A
is exactly the coarsest topology on A such that the inclusion map i : A 7→ X, x 7→ x is
continuous.
1.4.12. (a) Let f : X → Y be continuous and let Z be a subspace of X. Then the restriction
of f to Z is still continuous.
(b) Let f : X → Y be continuous and let Z be a space containing Y as a subspace.
Consider f as a function from X to Z, then it is still continuous.
1.4.13 (Gluing continuous functions). * Let X = A ∪ B where A and B are both open or
are both closed in X. Suppose f : X → Y , and f |A and f |B are both continuous. Then f is
continuous.
Another way to phrase this is the following. Let g : A → Y and h : B → Y be continuous
and g(x) = h(x) on A ∩ B. Define
g(x), x ∈ A
f (x) = .
h(x), x ∈ B
Then f is continuous.
Is it still true if the restriction that A and B are both open or are both closed in X is
removed?
1.4.18. In the Euclidean plane the upper half-plane {(x, y) ∈ R2 | y > 0} is homeomorphic
to the plane.
1.4.22. Show that N and Z are homeomorphic under the Euclidean topology.
Further, prove that any two discrete spaces having the same cardinalities are homeo-
morphic.
1.4.24. Show that any homeomorphism from Sn−1 onto Sn−1 can be extended to a homeo-
morphism from the unit disk Dn = B0 (0, 1) onto Dn .
1.4.25. The closure of a subspace is the union of the subspace and the set of all of its limit
points.
1.4.29. The map ϕ : [0, 2π) → S1 given by t 7→ (cost, sint) is a bijection but is not a
homeomorphism, under the Euclidean topology.
Hint: Compare the subinterval [1, 2π) and its image via ϕ.
1.4.30. Find the closures, interiors and the boundaries of the interval [0, 1) under the Eu-
clidean, discrete and trivial topologies of R.
1.4.31. * In a metric space X, a point x ∈ X is a limit point of the subset A of X if and only
if there is a sequence in A \ {x} converging to x.
Note: This is not true in general topological spaces.
1.4.32. * In a normed space, show that the boundary of the ball B(x, r) is the sphere
{y | d(x, y) = r}, and so the ball B0 (x, r) = {y | d(x, y) ≤ r} is the closure of B(x, r).
1.4.33. In a metric space, show that the boundary of the ball B(x, r) is a subset of the
sphere {y | d(x, y) = r}. Is the closed ball B0 (x, r) = {y | d(x, y) ≤ r} the closure of B(x, r)?
Hint: Consider a metric space consisting of two points.
Y X
1.4.34. Suppose that A ⊂ Y ⊂ X. Then A = A ∩Y . Furthermore if Y is closed in X then
Y X
A =A .
1.5. Connectedness
A topological space is said to be connected (liên thông) if is not a union of two
non-empty disjoint open subsets.
Equivalently, a topological space is connected if and only if its only subsets
which are both closed and open are the empty set and the space itself.
Example 1.5.2. The Euclidean real number line minus a point is no longer con-
nected.
PROOF. Let X be a topological space and let A and B be two connected sub-
spaces. Suppose that A ∩ B 6= 0.
/ Suppose that C is subset of A ∪ B that is both open
and closed in A ∪ B. Suppose that C 6= 0. / Then either C ∩ A 6= 0/ or C ∩ B 6= 0. /
Without loss of generality, assume that C ∩ A 6= 0.
/
Note that C ∩ A is both open and closed in A (we are using 1.4.4 here). Since
A is connected, C ∩ A = A. Then C ∩ B 6= 0, / so the same argument shows that
C ∩ B = B, hence C = A ∪ B.
Definition 1.5.7. Under the above equivalence relation, the equivalence classes are
called the connected components of the space.
24 1. GENERAL TOPOLOGY
Theorem 1.5.9. If two spaces are homeomorphic then they have the same num-
ber of connected components, i.e. there is a bijection between the collections of
connected components of the two spaces.
In particular, if two spaces are homeomorphic and one space is connected then
the other space is also connected.
Proposition 1.5.12. A connected subspace of the Euclidean real number line must
be an interval.
PROOF. In this proof we use the property that the set of all real numbers is
complete.
Let C be a non-empty proper, both open and closed subset of R.
Let x ∈/ C. The set D = C ∩ (−∞, x) is the same as the set C ∩ (−∞, x]. This
implies that D is both open and closed in R, and is bounded from above.
If D 6= 0,
/ let s = sup D. Since D is closed and s is a contact point of D, s ∈ D.
Since D is open s must belong to an open interval contained in D. But then there
are points in D which are bigger than s, a contradiction.
If D = 0/ we let E = C ∩ (x, ∞), consider t = inf E and proceed similarly.
Theorem 1.5.14. A subset of the Euclidean real number line is connected if and
only if it is an interval.
Example 1.5.15. The Euclidean line R is itself connected. Since the Euclidean Rn
is the union of all lines passing through the origin, it is connected.
Example 1.5.21. All open sets in a normed space are locally path-connected.
Topologist’s sine curve. The closure in the Euclidean plane of the graph of the
function y = sin 1x , x > 0 is often called the Topologist’s sine curve. This is a classic
example of a space which is connected but is not path-connected.
1
sin(1/x)
0.8
0.6
0.4
0.2
-0.2
-0.4
-0.6
-0.8
-1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Denote A = {(x, sin 1x ) | x > 0} and B = {0} × [−1, 1]. Then the Topologist’s
sine curve is X = A ∪ B.
PROOF. Suppose that there is a path γ(t) = (x(t), y(t)), t ∈ [0, 1] from the
origin (0, 0) on B to a point on A, we show that there is a contradiction.
Let t0 = sup{t ∈ [0, 1] | x(t) = 0}. Then x(t0 ) = 0, t0 < 1, and for all t > t0
we have x(t) > 0. Thus t0 is the moment when the path γ departs from B. We
can see that the path jumps immediately when it departs from B. Thus we will
show that γ(t) cannot be continuous at t0 , by showing that for any δ > 0 there are
t1 ,t2 ∈ (t0 ,t0 + δ ) such that y(t1 ) = 1 and y(t2 ) = −1.
To find t1 , note that the set x([tt ,t0 + δ2 ]) is an interval [0, x0 ] where x0 > 0.
There exists an x1 ∈ (0, x0 ) such that sin x11 = 1: we just need to take x1 = π +k2π 1
2
with sufficiently large k. There is t1 ∈ (t0 ,t0 + δ2 ] such that x(t1 ) = x1 . Then y(t1 ) =
sin x(t11 ) = 1. We can find t2 similarly.
Problems.
1.5.27. Here is a different proof of 1.5.13. Suppose that A and B are non-empty, disjoint
subsets of (0, 1) whose union is (0, 1). Let a ∈ A and b ∈ B. Let a0 = a, b0 = b, and for
each n ≥ 1 consider the middle point of the segment from an to bn . If an +b
2
n
∈ A then let
an +bn an +bn
an+1 = 2 and bn+1 = bn ; otherwise let an+1 = an and bn+1 = 2 . Then:
(1) The sequence {an | n ≥ 1} is a Cauchy sequence, hence is convergent to a number
a.
7On the surface of the Earth at any moment there are two opposite places where temperatures are
same!
28 1. GENERAL TOPOLOGY
(2) The sequence {bn | n ≥ 1} is also convergent to a. This implies that (0, 1) is
connected.
1.5.30. * If f : R → R is continuous under the Euclidean topology then its graph is con-
nected in the Euclidean plane. Moreover the graph is homeomorphic to R.
1.5.35. Let X be a topological space. A map f : X → Y is called a discrete map if Y has the
discrete topology and f is continuous. Show that X is connected if and only if all discrete
maps on X are constant.
Use this criterion to prove some of the results in this section.
1.5.36. What are the connected components of N and Q on the Euclidean real number
line?
1.5.38. Show that if a space has finitely many components then each component is both
open and closed.
Is it still true if there are infinitely many components?
1.5.39. Suppose that a space X has finitely many connected components. Show that a map
defined on X is continuous if and only if it is continuous on each components.
Is it still true if X has infinitely many components?
1.5.42. If two space are homeomorphic then there is a bijection between the collections of
path-connected components of the two spaces. In particular, if one space is path-connected
then the other space is also path-connected.
1.5.43. The plane with a point removed is path-connected under the Euclidean topology.
1.5.44. The plane with countably many points removed is path-connected under the Eu-
clidean topology.
Hint: Let A be countable and a ∈ R2 \ A. There is a line `a passing through a that does not
intersect A (by an argument involving cardinalities of sets).
1.5.45. * The Euclidean line and the Euclidean plane are not homeomorphic.
Hint: Delete a point from R. Use 1.4.20 and 1.5.9.
1.5.47. Show that R with the finite complement topology and R2 with the finite comple-
ment topology are homeomorphic.
1.5.48. A topological space is locally path-connected if and only if the collection of all
open path-connected subsets is a basis for the topology.
1.5.50. * Classify the alphabetical characters up to homeomorphisms, that is, which of the
following characters are homeomorphic to each other as subsets of the Euclidean plane?
ABCDEFGHIJKLMNOPQRSTUVWXYZ
Note that the result depends on the font you use!
Do the same for the Vietnamese alphabetical characters.
AĂÂBCDĐEÊGHIKLMNOÔƠPQRSTUƯVXY
Hint: Use 1.4.13 to modify a letter part by part.
Further readings
Invariance of dimension. That the Euclidean spaces R2 and R3 are not homeomorphic
is not easy. It is a consequence of the following difficult theorem of L. Brouwer in 1912:
Theorem 1.5.51 (Invariance of dimension). If two subsets of the Euclidean Rn are home-
omorphic and one set is open then the other is also open.
This theorem is often proved using Algebraic Topology, see for instance [Mun00, p.
381], [Vic94, p. 34], [Hat01, p. 126].
PROOF. Suppose that m < n. It is easy to check that the inclusion map Rm → Rn ,
(x1 , x2 , . . . , xm ) 7→ (x1 , x2 , . . . , xm , 0, . . . , 0) is a homeomorphism onto its image A ⊂ Rn . If A
is homeomorphic to Rn then by Invariance of dimension, A is open in Rn . But A is clearly
not open in Rn .
Jordan curve theorem. The following is an important and deep result of plane topology.
Theorem 1.5.53 (Jordan curve theorem). A simple, continuous, closed curve separates
the plane into two disconnected regions. More concisely, if C is a subset of the Euclidean
plane homeomorphic to the circle then R2 \C has two connected components.
Theorem 1.5.54. There is a continuous curve filling a rectangle on the plane. More con-
cisely, there is a continuous map from the interval [0, 1] onto the square [0, 1]2 under the
Euclidean topology.
Note that this map cannot be injective, in other words the curve cannot be simple.
Such a curve is called a Peano curve. It could be constructed as a limit of an iteration
of piecewise linear curves.
1.6. SEPARATION 31
1.6. Separation
Definition 1.6.1. We define:
T1 : A topological space is called a T1 -space if for any two points x 6= y there
is an open set containing x but not y and an open set containing y but not
x.
T2 : A topological space is called a T2 -space or Hausdorff if for any two
points x 6= y there are disjoint open sets U and V such that x ∈ U and
y ∈ V.
T3 : A T1 -space is called a T3 -space or regular (chính tắc) if for any point x
and a closed set F not containing x there are disjoint open sets U and V
such that x ∈ U and F ⊂ V . 8
T4 : A T1 -space is called a T4 -space or normal (chuẩn tắc) if for any two
disjoint closed sets F and G there are disjoint open sets U and V such that
F ⊂ U and G ⊂ V .
These definitions are often called separation axioms because they involve “sep-
arating” certain sets from each other by open sets.
Proposition 1.6.2. A space is T1 if and only if a set containing exactly one point is
a closed set.
Example 1.6.6. The real number line under the finite complement topology is T1
but is not T2 .
Remark 1.6.7. There are examples of a T2 -space which is not T3 , and a T3 -space
which is not T4 , but they are rather difficult, see 1.6.16, 1.11.12, [Mun00, p. 197]
and [SJ70].
Now suppose that A and B are disjoint closed sets. Let U = {x | d(x, A) <
d(x, B)} and V = {x | d(x, A) > d(x, B)}. Then A ⊂ U, B ⊂ V , U ∩V = 0,
/ and both
U and V are open.
Proposition 1.6.10. A T1 -space X is normal if and only if given a closed set C and
an open set U containing C there is an open set V such that C ⊂ V ⊂ V ⊂ U.
Problems.
1.6.11. If a finite set is a T1 -space then the topology is the discrete topology.
1.6.16. Show that the set R with the topology generated by all the subsets of the form (a, b)
and (a, b) ∩ Q is Hausdorff but is not regular.
Hint: Consider the set of all irrational numbers.
1.6.17. * Let X be normal, let f : X → Y be surjective, continuous, and closed. Prove that
Y is normal.
1.7. CONVERGENCE 33
1.7. Convergence
In metric spaces we can study continuity of functions via convergence of se-
quences. In general topological spaces, we need to use a notion more general than
sequences, called nets. Roughly speaking in general topological spaces, sequences
(countable indexes) might not be enough to describe the neighborhood systems at
a point, we need something of arbitrary index.
A directed set (tập được định hướng) is a (partially) ordered set I such that for
any two indices i and j there is an index k that is greater than or is equal to both i
and j. In symbols: ∀i, j ∈ I, ∃k ∈ I, k ≥ i ∧ k ≥ j.
A net (lưới) (also called a generalized sequence) is a map from a directed set
to a space. In other words, a net on a space X with index set a directed set I is a
map x : I → X. It is an element of the set ∏i∈I X. Thus, writing xi = x(i), we often
denote the net as (xi )i∈I . The notation {xi }i∈I is also used.
Example 1.7.1. Nets with index set I = N with the usual order are exactly se-
quences.
Convergence.
Definition 1.7.3. A net (xi ) is said to be convergent (hội tụ) to x ∈ X if for each
neighborhood U of x there is an index i ∈ I such that if j ≥ i then x j belongs to U.
The point x is called a limit of the net {xi }i∈I , and we often write xi → x.
Example 1.7.5. Let X = {x1 , x2 , x3 } with topology {0, / X, {x1 , x3 }, {x2 , x3 }, {x3 }}.
The net (x3 ) converges to x1 , x2 , and x3 . The net (x1 , x2 ) converges to x2 .
If X has the trivial topology then any net in X is convergent to any point in X.
PROOF. The proof is simply a repeat of the proof for the case of metric spaces.
(⇒) Suppose that f is continuous at x. Let U is a neighborhood of f (x). Then
f −1 (U) is a neighborhood of x in X. Since (xi ) is convergent to x, there is an i ∈ I
such that for all j ≥ i we have x j ∈ f −1 (U), which implies f (x j ) ∈ U.
(⇐) We will show that if U is an open neighborhood in Y of f (x) then f −1 (U)
is a neighborhood in X of x. Suppose the contrary, then x is not an interior point
of f −1 (U), so it is a limit point of X \ f −1 (U). There is a net (xi ) in X \ f −1 (U)
convergent to x. Since f is continuous, f (xi ) ∈ Y \ U is convergent to f (x) ∈ U.
That contradicts the assumption that U is open.
Proposition 1.7.9. If a space is Hausdorff then a net has at most one limit.
PROOF. Suppose that a net (xi ) is convergent to two different points x and y.
Since the space is Hausdorff, there are disjoint open neighborhoods U and V of x
and y. There is i ∈ I such that for γ ≥ i we have xγ ∈ U, and there is j ∈ I such that
for γ ≥ j we have xγ ∈ U. Since there is a γ ∈ I such that γ ≥ i and γ ≥ j, the point
xγ will be in U ∩V , a contradiction.
Problems.
1.7.11. Let R have the finite complement topology. Let (xi )i∈R be a net of distinct points
in R, i.e. xi 6= x j if i 6= j. Where does this net converge to?
1.7.12. Suppose that τ1 and τ2 are two topologies on X. Show that if for all nets xi and all
τ1 τ2
points x, xi → x ⇒ xi → x then τ1 ⊃ τ2 .
In other words, if convergence in τ1 implies convergence in τ2 then τ1 is finer than τ2 .
Is the converse true?
Hint: Consider the identity map on X.
1.7.14. * Let Y be Hausdorff and let f , g : X → Y be continuous. Show that the set {x ∈
X | f (x) = g(x)} is closed in X, by:
(1) using nets.
(2) not using nets.
Show that, as a consequence, if f and g agree on a dense (trù mật) subspace of X (meaning
the closure of that subspace is X) then they agree on X.
1.7.15. The converse statement of 1.7.9 is also true. A space is Hausdorff if and only if a
net has at most one limit.
Hint: Suppose that there are two points x and y that could not be separated by open sets.
Consider the directed set whose elements are pairs (Ux ,Vy ) of open neighborhoods of x and y, under
set inclusion. Take a net n such that n(Ux ,Vy ) is a point in Ux ∩Vy .
1.7.16 (Sequence is not adequate for describing convergence). * Let (A, ≤) be a well-
ordered uncountable set (see 1.1.34). If A does not have a biggest element then add an
element to A and define that element to be the biggest one. Thus we can assume now that
A has a biggest element, denoted by ∞. For a, b ∈ A denote [a, b] = {x ∈ A | a ≤ x ≤ b} and
[a, b) = {x ∈ A | a ≤ x < b}. For example we can write A = [0, ∞].
Let Ω be the smallest element of the set {a ∈ A | [0, a] is uncountable} (this set is
non-empty since it contains ∞).
(1) Show that [0, Ω) is uncountable, and for all a ∈ A, a < Ω the set [0, a] is count-
able.
(2) Show that every countable subset of [0, Ω) is bounded in [0, Ω).
(3) Consider [0, Ω] with the order topology. Show that Ω is a limit point of [0, Ω].
(4) However, show that a sequence in [0, Ω) cannot converge to Ω.
So this is an example of a space for which sequence is not adequate for describing conver-
gence.
S
Hint: (b) Let C be a countable subset of [0, Ω). The set c∈C [0, c) is countable while the set
[0, Ω) is uncountable.
(1) if A, B ∈ F then A ∩ B ∈ F,
(2) if A ⊂ B and A ∈ F then B ∈ F.
For example, given a point, the collection of all neighborhoods of that point is a filter.
A filter is said to be convergent to a point if any neighborhood of that point is an
element of the filter.
A filter-base (cơ sở lọc) is a collection G of non-empty subsets of X such that if
A, B ∈ G then there is C ∈ G such that G ⊂ (A ∩ B).
If G is a filter-base in X then the filter generated by G is defined to be the collection
of all subsets of X each containing an element of G: {A ⊂ X | ∃B ∈ G, B ⊂ A}.
For example, in a metric space, the collection of all open balls centered at a point is
the filter-base for the filter consisting of all neighborhoods of that point.
A filter-base is said to be convergent to a point if the filter generated by it converges to
that point.
(1) Show that a filter-base is convergent to x if and only if every neighborhood of x
contains an element of the filter-base.
(2) Show that a point x ∈ X is a limit point of a subspace A of X if and only if there
is a filter-base in A \ {x} convergent to x.
(3) Show that a map f : X → Y is continuous at x if and only if for any filter-base F
that is convergent to x, the filter-base f (F) is convergent to f (x).
Filter gives an alternative way to net for describing convergence. For more see [Dug66, p.
209], [Eng89, p. 49], [Kel55, p. 83].
1.8. COMPACT SPACE 37
Example 1.8.3. Any subset of R with the finite complement topology is compact.
Note that this space is not Hausdorff (1.6.6).
Theorem 1.8.8. A space is compact if and only if every collection of closed subsets
with the finite intersection property has non-empty intersection.
Compact sets in metric spaces. The following results have been studied in
Functional Analysis. It is useful for the reader to review them.
Proposition 1.8.9. A metric space is compact if and only if every sequence has a
convergent subsequence.
Example 1.8.12. In the Euclidean space Rn the closed ball B0 (a, r) is compact.
Theorem 1.8.13 (Lebesgue’s number). Let X be a compact metric space and let I
be an open cover of X. Then there exists a number ε > 0 such that any ball B(x, ε)
is contained in a member of the cover I.
1.8.19 (The extreme value theorem). * If X is a compact space and f : X → (R, Euclidean)
is continuous then f has a maximum value and a minimum value.
1.8.21. In a Hausdorff space a point and a disjoint compact set can be separated by open
sets.
Hint: See the proof of 1.8.4.
1.8.22. In a regular space a closed set and a disjoint compact set can be separated by open
sets.
1.8.23. In a Hausdorff space two disjoint compact sets can be separated by open sets.
Hint: Use 1.8.21.
1.8.25. * The set of n × n-matrix with real coefficients, denoted by M(n; R), could be
2
naturally considered as a subset of the Euclidean space Rn by considering entries of a
matrix as coordinates, via the map
(ai, j ) 7−→ (a1,1 , a2,1 , . . . , an,1 , a1,2 , a2,2 , . . . , an,2 , a1,3 , . . . , an−1,n , an,n ).
The Orthogonal Group O(n) is defined to be the group of matrices representing or-
thogonal linear maps of Rn , that are linear maps that preserve inner product. Thus
The Special Orthogonal Group SO(n) is the subgroup of O(n) consisting of all or-
thogonal matrices with determinant 1.
(1) Show that any element of SO(2) is of the form
!
cos(ϕ) − sin(ϕ)
.
sin(ϕ) cos(ϕ)
This is a rotation in the plane around the origin with an angle ϕ. Thus SO(2) is
the group of rotations on the plane around the origin.
(2) Show that SO(2) is path-connected.
(3) How many connected components does O(2) have?
(4) Show that SO(n) is compact.
(5) The General Linear Group GL(n; R) is the group of all invertible n × n-matrices
with real coefficients. Show that GL(n; R) is not compact.
(6) Find the number of connected components of GL(n; R).
1.8.26 (Point-wise convergence topology). Let X and Y be two spaces. Let ( fi )i∈I be a
net of functions from X to Y .
We say that the net ( fi )i∈I converges to a function f : X → Y point-wise if for any x ∈ X
the net ( fi (x))i∈I converges to f (x).
40 1. GENERAL TOPOLOGY
1.8.27 (Compact-open topology). * Let X and Y be two spaces. Let C(X,Y ) be the set
of all continuous functions from X to Y .
The topology on C(X,Y ) generated by all sets of the form
Proposition 1.9.2. If each bi is a basis for Xi then ∏ni=1 bi is a basis for the product
topology on ∏ni=1 Xi .
Definition 1.9.4. Let {(Xi , τi )}i∈I be a family of topological spaces. The product
topology on ∏i∈I Xi is the topology generated by the collection F consisting of all
sets of the form ∏i∈I Ui , where Ui ∈ τi and Ui = Xi for all except finitely many i ∈ I.
Theorem 1.9.6 (Product topology is the topology such that projections are con-
tinuous). The product topology is the coarsest topology on ∏i∈I Xi such that all
the projection maps pi are continuous. In other words, the product topology is the
topology generated by the projection maps.
42 1. GENERAL TOPOLOGY
Theorem 1.9.7 (Map to product space is continuous if and only if each compo-
nent map is continuous). A map f : Y → ∏i∈I Xi is continuous if and only if each
component fi = pi ◦ f is continuous.
Remark 1.9.8. However continuity of a map from a product space is not the same
as continuity with respect to each variable, as we have seen in Calculus.
Tikhonov Theorem.
9
Theorem 1.9.10 (Tikhonov theorem). A product of compact spaces is compact.
Example 1.9.11. Let [0, 1] have the Euclidean topology. The space ∏i∈Z+ [0, 1] is
called the Hilbert cube. By Tikhonov theorem the Hilber cube is compact.
In what follows we will overcome this difficulty by first enlarging the collection
F.
(1) We show that there is a maximal collection F̃ of subsets of X such that F̃
contains F and still has the finite intersection property. We will use Zorn
lemma for this purpose.11
Let K be the collection of collections G of subsets of X such that G
contains F and has the finite intersection property. On K we define an
order by the usual set inclusion.
Now suppose that L is a totally ordered subcollection of K. Let H =
G∈L G. We will show that H ∈ K, therefore H is an upper bound of L.
S
11This is a routine step; it might be easier for the reader to carry it out instead of reading.
44 1. GENERAL TOPOLOGY
(4) Let xi ∈ A∈F̃ pi (A) and let x = (xi )i∈I ∈ ∏i∈I [ A∈F̃ pi (A)]. We will show
T T
1.9.13. Show that the sphere S2 with the North Pole and the South Pole removed is home-
omorphic to the infinite cylinder S1 × R.
1.9.14. Show that a space X is Hausdorff if and only if the diagonal ∆ = {(x, x) ∈ X × X}
is closed in X × X, by:
(1) using nets,
(2) not using nets.
1.9.15. If Y is Hausdorff and f : X → Y is continuous then the graph of f (the set {(x, f (x)) | x ∈
X}) is closed in X ×Y .
1.9.16. If for each i ∈ I the space Xi is homeomorphic to the space Yi then ∏i∈I Xi is
homeomorphic to ∏i∈I Yi .
1.9.17. * Show that each projection map pi is a an open map, mapping an open set onto an
open set.
Hint: Only need to show that the projection of an element of the basis is open.
1.9.21. * Fix a point O = (Oi ) ∈ ∏i∈I Xi . Define the inclusion map f : Xi → ∏i∈I Xi by
O if j 6= i
j
x 7→ f (x) with f (x) j = .
x if j = i
If S ⊂ P then define Z(S) to be the set of all common zeros of all polynomials in S,
thus Z(S) = {x ∈ Fn | ∀p ∈ S, p(x) = 0}. Such a set is called an algebraic set.
(1) Show that if we define that a subset of Fn is closed if it is algebraic, then this
gives a topology on Fn , called the Zariski topology.
(2) Show that the Zariski topology on F is exactly the finite complement topology.
(3) Show that if both F and Fn have the Zariski topology then all polynomials on Fn
are continuous.
(4) Is the Zariski topology on Fn the product topology?
The Zariski topology is used in Algebraic Geometry.
1.9.28. Using the characterization of compact subsets of Euclidean spaces, prove the
Tikhonov theorem for finite products of compact subsets of Euclidean spaces.
1.9.29. Let (X, dX ) and (Y, dY ) be metric spaces. Prove that the product topology on X ×Y
is given by the metric
Using the characterization of compact metric spaces in terms of sequences, prove the
Tikhonov theorem for finite products of compact metric spaces.
Further readings
Strategy for a proof of Tikhonov theorem based on net. The proof that we will
outline here is based on further developments of the theory of nets and a characterization
of compactness in terms of nets.
Definition 1.9.30 (Subnet). Let I and I 0 be directed sets, and let h : I 0 → I be a map such
that
∀k ∈ I, ∃k0 ∈ I 0 , (i0 ≥ k0 ⇒ h(i0 ) ≥ k).
If n : I → X is a net then n ◦ h is called a subnet of n.
Definition 1.9.31. Universal net A net n in X is universal if for any subset A of X either n
is eventually in A or n is eventually in X \ A.
The proof of the last two propositions above could be found in [Bre93].
Then we finish the proof of Tikhonov theorem as follows.
1.10. Compactification
A compactification (compắc hóa) of a space X is a compact space Y such that
X is homeomorphic to a dense subset of Y .
Example 1.10.1. A compactification of the Euclidean interval (0, 1) is the Eu-
clidean interval [0, 1]. Another is the circle S1 . Yet another is the Topologist’s sine
curve {(x, sin 1x ) | 0 < x ≤ 1} ∪ {(0, y) | − 1 ≤ y ≤ 1} (see 1.5.1).
Example 1.10.2. A compactification of the complex plane C is the sphere S2 , is
often called the Riemann sphere.
One-point compactification. In some cases it is possible to compactify a non-
compact space by adding just one point. It is called a one-point compactification.
For example [0, 1] is a compactification of [0, 1).
Let X be a space, let ∞ be not in X, and let X ∞ = X ∪ {∞}. Let us see what a
topology on X ∞ should be in order for X ∞ to be compact. First X ∞ contains X as a
subspace, so each open subset of X must be open in X ∞ . If an open subset of X ∞
contains ∞ then its complement in X ∞ must be closed in X ∞ , and hence is compact,
and is a subset of X.
The existence of such a topology is given in the following:
Theorem 1.10.3 (Alexandroff compactification). Let X be a space. Let ∞ be not
in X and let X ∞ = X ∪ {∞}. Define a topology on X ∞ as follow: an open set in X ∞
is an open subset of X, or is X ∞ \C where C is a closed compact subset of X. With
this topology X ∞ is compact and contains X as a subspace.
If X is not compact then X is dense in X ∞ , and in this case X ∞ is called the
Alexandroff compactification of X. 12
PROOF. We go through several steps.
(1) We check that we really have a topology.
Let {Ci }i∈I be a collection of closed compact sets in X. Then i (X ∞ \
S
Thus in a completely regular space a point and a closed set disjoint from it can
be separated by a continuous real function.
PROOF. The space Y is Hausdorff, by 1.9.22. Then the result follows from
1.6.13.
X
Φ / Φ(X)
{
{{
f
{{{ ˜
}{{ f
R
PROOF. A continuous extension of f , if exists, is unique, by 1.7.14.
Since p f ◦ Φ = f the obvious choice for f˜ is the projection p f .
Problems.
1.10.12. Find the one-point compactification of the Euclidean (0, 1) ∪ (2, 3).
1.10.15. Show that Q is not locally compact (under the Euclidean topology of R). Is its
Alexandroff compactification Hausdorff?
1.10.16. What is the one-point compactification of the Euclidean open ball B(0, 1)? Find
the one-point compactification of the Euclidean space Rn .
1.10.17. What is the one-point compactification of the Euclidean annulus {(x, y) ∈ R2 | 1 <
x2 + y2 < 2}?
1.10.20. If there is a topology on the set X ∞ = X ∪ {∞} such that it is compact, Hausdorff,
and containing X as a subspace, then X must be Hausdorff, locally compact, and there is
only one such topology–the topology of the Alexandroff compactification.
1.10.21. We could have noticed that the notion of local compactness as we have defined
is not apparently a local property. For a property to be local, every neighborhood of any
point must contain a neighborhood of that point with the given property (as in the cases of
local connectedness and local path-connectedness).
Show that for Hausdorff spaces local compactness is indeed a local property.
Further readings
By 1.10.26 if a space has a Hausdorff Alexandroff compactification then it also has a
Hausdorff Stone-Cech compactification.
In a certain sense, for a noncompact space the Alexandroff compactification is the
“smallest” Hausdorff compactification of the space and the Stone-Cech compactification
is the “largest” one. For more discussions on this topic see for instance [Mun00, p. 237].
1.11. URYSOHN LEMMA 53
Urysohn lemma.
F ⊂ U0 ⊂ U0 ⊂ U 1n ⊂ U 1n ⊂ U 2n ⊂ U 2n ⊂ U 3n ⊂ U 3n ⊂ · · · ⊂
2 2 2 2 2 2
⊂ U 2n −1
n
⊂ U 2n −1
n
⊂ U 2nn = U1 .
2 2 2
Let I = { 2mn
| m, n ∈ N; 0 ≤ m ≤ 2n }.
We have a family of open sets
{Ur | r ∈ I} having the property r < s ⇒ Ur ⊂ Us .
(2) We can check that I is dense in [0, 1] (this is really the same thing as that
any real number in [0, 1] can be written in binary form).
(3) Define f : X → [0, 1],
inf{r ∈ I | x ∈ U } if x ∈ U
r
f (x) = .
1 if x ∈
/U
54 1. GENERAL TOPOLOGY
(5) We have f (x) > a if and only if there is r ∈ I such that r > a and x ∈
/ Ur .
Thus {x | f (x) > a} = {x ∈ X \Ur | r > a} = r>a X \Ur .
S
(1) Suppose that f is neither bounded from below nor bounded from above.
Let h be a homeomorphism from (−∞, ∞) to (0, 1). Then the range
of f1 = h ◦ f is a subset of (0, 1), therefore it can be extended as in
the previous case to a continuous function g1 such that infx∈X g1 (x) =
infx∈F f1 (x) = 0 and supx∈X g1 (x) = supx∈F f1 (x) = 1.
If the range of g1 includes neither 0 nor 1 then g = h−1 ◦ g1 will be
the desired function.
It may happens that the range of g1 includes either 0 or 1. In this case
let C = g−11 ({0, 1}). Note that C ∩ F = 0./ By Urysohn lemma, there is
a continuous function k : X → [0, 1] such that k|C = 0 and k|F = 1. Let
g2 = kg1 + (1 − k) 12 . Then g2 |F = g1 |F and the range of g2 is a subset
of (0, 1) (g2 (x) is a certain convex combination of g1 (x) and 21 ). Then
g = h−1 ◦ g2 will be the desired function.
(2) If f is bounded from below then similarly to the previous case we can use
a homeomorphism h : [a, ∞) → [0, 1), and we let C = g−1 1 ({1}).
The case when f is bounded from above is similar.
Problems.
1.11.5. A normal space is completely regular. So: normal ⇒ completely regular ⇒ regu-
lar. In other words: T4 ⇒ T3 1 ⇒ T3 .
2
1.11.6. Show that the Tiestze extension theorem implies the Urysohn lemma.
1.11.7. The Tiestze extension theorem is not true without the condition that the set F is
closed.
1.11.8. Show that the Tiestze extension theorem can be extended to maps to the space
∏i∈I R where R has the Euclidean topology.
1.11.9. Let X be a normal space and F be a closed subset of X. Then any continuous map
f : F → Sn can be extended to an open set containing F.
Hint: Use 1.11.8.
K = {(x, y) ∈ R2 | y > 0}, together with sets of the form {p} ∪ D where p is a point on the
line L = {(x, y) ∈ R2 | y = 0} and D is an open disk in K tangent to L at p. This is called
the Niemytzki space.
(1) Check that this is a topological space.
(2) What is the subspace topology on L?
(3) What are the closed sets in H?
(4) Show that H is Hausdorff.
(5) Show that H is regular.
(6) Show that H is not normal.
Hint: Any real function on L is continuous. The cardinality of the set of such
function is cc . A real function on H is continuous if and only if it is continuous on
the dense subset of points with rational coordinates, so the cardinality of the set of such
functions is at most cℵ0 . Since cc > cℵ0 , the space H cannot be normal, by Tiestze
Extention Theorem.
Further readings
Geometric Topology
Theorem 2.1.1. Suppose that f : X → Y and Y has the quotient topology corre-
sponding to f . Let Z be a topological space. Then g : Y → Z is continuous if and
only if g ◦ f is continuous.
f
X ? /Y
? ??
?? g
g◦ f ?
Z
The following result will provide us a tool to identify some quotient spaces:
p
X D / X/ ∼
D DD
DD
DD h
f D!
Y
Example 2.1.3 (Gluing the two end-points of a line segment gives a circle).
More precisely [0, 1]/0 ∼ 1 is homeomorphic to S1 :
p
[0, 1] / [0, 1]/0 ∼ 1
KKK
KKK
K
f KKKK
h
%
S1
Here f is the map t 7→ (cos(2πt), sin(2πt)).
The torus T 2 is homeomorphic to a subspace of R3 (we say that the torus can be
embedded in R3 ). The subspace is the surface of revolution obtained by revolving
a circle around a line not intersecting it.
Suppose that the circle is on the Oyz-plane, the center is on the y-axis and the
axis for the rotation is the z-axis. Let a be the distance from the center of the circle
to the z-axis, b be the radius of the circle (a > b). Let S be the surface of revolution,
[0, 2π] × [0, 2π] / T 2 = ([0, 2π]/0 ∼ 2π) × ([0, 2π]/0 ∼ 2π)
WWWW
WWWW
WWWW
WWWW h
WWWW
f
WWWW
+
S
where f (s,t) = ((a + b cos(s)) cos(t), (a + b cos(s)) sin(t), b sin(s)).
p
We also obtain an implicit equation for this surface: ( x2 + y2 − a)2 + z2 = b2 .
Example 2.1.6 (Gluing the boundary circle of a disk together gives a sphere).
More precisely D2 /∂ D2 is homeomorphic to S2 . We only need to construct a con-
tinuous map from D2 onto S2 such that after quotient out by the boundary ∂ D2 it
becomes injective, see Figure 2.1.6.
FIGURE 2.1.3.
Example 2.1.7 (The Mobius band). Gluing a pair of opposite edges of a square
in opposite directions gives the Mobius band (dải, lá Mobius). More precisely the
Mobius band is X = [0, 1] × [0, 1]/ ∼ where (0,t) ∼ (1, 1 − t) for all 0 ≤ t ≤ 1.
The Mobius band could be embedded in R3 . It is homeomorphic to a sub-
space of R3 obtained by rotating a straight segment around the z-axis while also
turning that segment “up side down”. The embedding can be induced by the map
(s,t) 7→ ((a +t cos(s/2)) cos(s), (a +t cos(s/2)) sin(s),t sin(s/2)), with 0 ≤ s ≤ 2π
and −1 ≤ t ≤ 1.
62 2. GEOMETRIC TOPOLOGY
t
a
s/2
Example 2.1.8 (The projective plane). Identifying opposite points on the bound-
ary of a disk (they are called antipodal points) we get a topological space called
the projective plane (mặt phẳng xạ ảnh) RP2 .
Example 2.1.9 (The projective space). More generally, identifying antipodal points
of Sn , or homeomorphically, identifying antipodal boundary points of Dn gives us
the projective space (không gian xạ ảnh) RPn .
Example 2.1.10 (The Klein bottle). Identifying one pair of opposite edges of a
square and the other pair in opposite directions gives a topological space called the
Klein bottle . More precisely it is [0, 1] × [0, 1]/ ∼ with (0,t) ∼ (1,t) and (s, 0) ∼
(1 − s, 1).
This space cannot be embedded in R3 , but it can be immersed in R3 . An
immersion (phép nhúng chìm) is a local embedding. More concisely, f : X → Y is
an immersion if each point in X has a neighborhood U such that f |U : U → f (U)
is a homeomorphism. Intuitively, an immersion allows self-intersection (tự cắt).
Problems.
2.1. QUOTIENT SPACE 63
2.1.12. Show that RP2 is homeomorphic to the sphere with antipodal points identified,
that is, S2 /x ∼ −x.
2.1.15. Show that the following spaces are homeomorphic (one of them is the Klein bottle).
b a
a a a b
b b
2.1.16. Describe the space that is the sphere S2 quotient by its equator S1 .
2.1.18. The one-point compactification of the open Mobius band (the Mobius band without
the boundary circle) is the projective space RP2 .
2.1.21. In order for the quotient space X/ ∼ to be Hausdorff, a necessary condition is that
each equivalence class [x] must be a closed subset of X. Is this condition sufficient?
2.2. TOPOLOGICAL MANIFOLDS 65
Remark 2.2.2. In this section we assume Rn has the Euclidean topology unless we
mention otherwise.
an open set. Then the graph of f , the set {(x, f (x)) | x ∈ D} as a subspace of Rn+1
is an n-dimensional manifold.
FIGURE 2.2.1. The sets with same colors are glued to form a
neighborhood of a point on the torus. Each such neighborhood
is homeomorphic to a sphere.
p
We can also view the torus as a surface in R3 , given by the equation ( x2 + y2 −
a)2 +z2 = b2 . As such it can be covered by the open subsets corresponding to z > 0,
z < 0, x2 + y2 < a2 , x2 + y2 > a2 .
Remark 2.2.9. The interval [0, 1] is not a manifold. It is a manifold with boundary.
Similarly the Mobius band is not a manifold, but is a manifold with boundary. We
will not give a precise definition of manifold with boundary here.
Problems.
2.2.10. Show that if two spaces are homeomorphic and one space is an n-dimensional
manifold then the other is also an n-dimensional manifold.
Further readings
Bernard Riemann proposed the idea of manifold in his Habilitation dissertation. A
translation of this article is available in [Spi99].
Two conditions are often added to the definition of a manifold: it is Hausdorff, and it
has a countable basis. The first condition is useful for doing Analysis on manifolds, and
the second condition guarantees the existence of Partition of Unity.
Connected sum. Let S and T be two surfaces. From each surface deletes an open
disk, obtaining a surface with a circle boundary. Glue the two surfaces along the
two boundary circles. The resulting surface is called the connected sum (tổng trực
tiếp) of the two surfaces, denoted by S#T .
◦ ◦
S \ D2 T \ D2 S #T
The connected sum does not depend on the choices of the disks.
Classification.
Notice that at this stage we have not yet been able to prove that those surfaces
are distinct.
2.3. CLASSIFICATION OF COMPACT SURFACES 69
1
0.8
1 0.6
0.4
0.2
0.5 0
-0.2
0 -0.4
-0.6
-0.5 -0.8
-1
-1
3
2
1
-3 0
-2
-1 -1
0
1 -2
2
3 -3
It is known that any surface can be triangulated, this was proved in the 1920’s.
70 2. GEOMETRIC TOPOLOGY
χ(S) = V − E + F
Theorem 2.3.3. The Euler Characteristics with respect to two triangulations of the
same surface are equal.
By this theorem the Euler Characteristics of a surface is defined and does not
depend on the choice of triangulation. Since the Euler Characteristics does not
change under homeomorphisms, just like the number of connected components, it
is said to be a topological invariant (bất biến tôpô). If two surfaces have different
Euler Characteristics, then they are not homeomorphic.
Problems.
2.3.5. (a) Show that RP2 with an open disk removed is the Mobius band.
(b) Show that T 2 #RP2 = K#RP2 , where K is the Klein bottle.
2.3.6. Show that gluing two Mobius band along their boundaries gives the Klein bottle. In
other words, RP2 #RP2 = K. 2
2There is a humorous poem:
A mathematician named Klein
Thought the Mobius band was divine
Said he, “If you glue
The edges of two,
You’ll get a weird bottle like mine.”
2.3. CLASSIFICATION OF COMPACT SURFACES 71
2.3.7 (Surfaces are homogeneous). A space is homogeneous (đồng nhất) if given two
points there exists a homeomorphism from the space to itself bringing one point to the
other point.
(1) Show that the sphere S2 is homogeneous.
(2) Show that the torus T 2 is homogeneous.
It is known that any manifold is homogeneous, see 3.7.4.
2.3.10. * Compute the Euler Characteristics of all connected compact without boundary
surfaces.
Deduce that the orientable surfaces S2 and Tg , for different g, are distinct, meaning not
homeomorphic to each other. Similarly the non-orientable surfaces Mg are all distinct.
Further readings
Proof of the Classification Theorem. Let S be a triangulated surface. Cut S along the
triangles. Label the edges by alphabet characters and mark the orientations of each edge.
In this way each edge will appear twice on two different triangles.
Take one triangle. Pick a second triangle which has one common edge with the first
one, then glue the two along the common edge following the orientation of the edge. Con-
tinue this gluing process in such a way that at every step the resulting polygon is planar.
This is possible if at each stage the gluing is done in such a way that there is one edge of
the polygon such that the entire polygon is on one of its side. The last polygon P is called
a fundamental polygon of the surface.
The boundary of the fundamental polygon consists of labeled and oriented edges.
Choose one edge as the initial one then follow the edge of the polygon in a predetermined
direction. This way we associate each polygon with a word w.
We consider two words equivalent if they give rise to homeomorphic surfaces.
Example 2.3.11. A fundamental polygon of the sphere and its associated word.
In the reverse direction, the surface can be reconstructed from an associated word. We
consider two words equivalent if they give rise to homeomorphic surfaces. In order to find
all possible surfaces we will find all possible associated words up to equivalence.
Lemma 2.3.14. A pair of the form aa−1 in w can be deleted, meaning that this action will
give an equivalent word corresponding to a homeomorphic surface.
a a−1
a
Lemma 2.3.15. The word w is equivalent to a word whose all of the vertices of the associ-
ated polygon is identified to a single point on the associated surface (w is “reduced”).
When there is only one P vertex left, we arrive at the situation in Lemma 2.3.14.
Lemma 2.3.16. A word of the form −a − a− is equivalent to a word of the form −aa−.
Lemma 2.3.17. Suppose that w is reduced. Assume that w has the form −aαa−1 − where
α is a non-empty word. Then there is a letter b such that b is in α but the other b or b−1 is
not.
PROOF. If all letters in α appear in pairs then the vertices in the part of the polygon
associated to α are identified only with themselves, and are not identified with a vertex
outside of that part. This contradicts the assumption that w is reduced.
Lemma 2.3.18. A word of the form −a − b − a−1 − b−1 − is equivalent to a word of the
form −aba−1 b−1 −.
Lemma 2.3.19. A word of the form −aba−1 b−1 − cc− is equivalent to a word of the form
−a2 − b2 − c2 −.
2.3. CLASSIFICATION OF COMPACT SURFACES 73
P P
b a a
b
c Q c
Q
Q
c
b P P
b
c Q
b P
a
c Q
c a
a c
a c a
c c
a
PROOF. Do the operation in the figure, after that we are in a situation where we can
apply Lemma 2.3.16 three times.
a a
b c b b c c b
a a
a a
c c c
b
c d
a
a
a
d
c c
d a
d c c
a
d
b
d
d
c c
c
d
a b
Also it is crucial from the proof of 2.3.16 that this step will not undo the steps before
it.
4. If there is no pair of the form −aαa−1 where α 6= 0,
/ then stop: w has the form
a21 a22 · · · a2g .
5. w has the form −aαa−1 where α 6= 0. / By 2.3.17 w must has the form −a − b −
a−1 − b−1 −, since after Step 3 there could be no −b − a−1 − b−.
6. Repeatedly apply 2.3.18 finitely many times until w no longer has the form −aαbβ a−1 γb−1 −
where at least one of α, β , or γ is non-empty.
7. If w is not of the form −aa− then stop: w has the form a1 b1 a−1 −1 −1 −1
1 b1 · · · ag bg ag bg .
8. w has the form −aba−1 b−1 − cc−. Use 2.3.19 finitely many times to transform w
to the form a21 a22 · · · a2g .
76 2. GEOMETRIC TOPOLOGY
2.4. Homotopy
Homotopy of paths. Recall that a path (đường đi) in a space X is a continuous
map α from the Euclidean interval [0, 1] to X. The point α(0) is called the initial
end point, and α(1) is called the final end point. In this section for simplicity of
presentation we assume the domain of a path is the Euclidean interval [0, 1] instead
of any Euclidean closed interval as before.
Let α and β be two paths from a to b in X. A homotopy (phép đồng luân) from
α to β is a family of maps Ft : X → X, t ∈ [0, 1], such that the map (t, s) 7→ Ft (s) is
continuous, F0 = α, F1 = β , and for each t the path Ft goes from a to b.
If there is a homotopy from α to β we say that α is homotopic (đồng luân) to
β , sometimes written α ∼ β .
F0 = α
Ft
a b
F1 = β
Example 2.4.1. In a normed space any two paths α and β with the same initial
points and end points are homotopic, via the homotopy (1 − t)α + tβ . This is also
true for any convex subset of a normed space.
Note that continuity of a map is not the same as continuity with respect to each vari-
able (see 1.9.8). However by an argument similar to 1.4.13 we can check directly
that the map H is continuous. It is a homotopy from α to γ.
2.4. HOMOTOPY 77
Example 2.4.5. An annulus S1 × [0, 1] has a deformation retract to one of its circle
boundary S1 × {0}.
Remark 2.4.6. Homotopy of paths is a special case of homotopy of maps, with the
further requirement that the homotopy fixes the initial point and the end point.
Definition 2.4.7. The space X is said to be homotopic to the space Y if there are
continuous maps f : X → Y and g : Y → X such that g ◦ f is homotopic to the
identity map idX and f ◦ g is homotopic to the identity map idY .
Example 2.4.9. The annulus and the Mobius band are both homotopic to the circle,
but are not homeomorphic to it.
78 2. GEOMETRIC TOPOLOGY
The fundamental group. Let α be a path from a to b. Then the inverse path of
α is defined to be the path α −1 (t) = α(1 − t) from b to a.
Let α be a path from a to b, and β be a path from b to c, then the composition
(hợp) of α with β is defined to be the path
α(2t), 0 ≤ t ≤ 21
γ(t) =
β (2t − 1), 1 ≤ t ≤ 1.
2
PROOF. Let F be the first homotopy and G be the second homotopy. Consider
Ht = Ft · Gt .
The dependence of the fundamental group on the base point is explained in the
following proposition.
Thus for a path-connected space the choice of the base point is not important
for the fundamental group.
Problems.
2.4.19. Show that the Mobius band has a deformation retract to a circle.
2.4.20. Show that if the space X is homotopic to the space Y and Y is homotopic to the
space Z then X is homotopic to Z.
2.4.22. Show that the homotopy type of the Euclidean plane with a point removed does
not depend on the choice of the point.
Further readings
One of most celebrated achievements in Topology is the resolution of the Poincaré
conjecture:
80 2. GEOMETRIC TOPOLOGY
The proof of this statement is the result of a cummulative effort of many mathemati-
cians, including Stephen Smale (for dimension ≥ 5, early 1960s), Michael Freedman (for
dimension 4, early 1980s), and Grigory Perelman (for dimension 3, early 2000s).
CHAPTER 3
Differential Topology
Remark 3.1.1. From now on in this chapter we assume that Rn has the Euclidean
topology.
Remark 3.1.4. In this part we only study smooth manifolds imbedded in Euclidean
spaces. Unless stated otherwise, manifolds mean smooth manifolds.
Example 3.1.8. The graph of a smooth function y = f (x) for x ∈ (a, b) (a smooth
curve) is a 1-dimensional manifold.
Example 3.1.9. The real number line R is a manifold, but t 7→ t 3 is not a parametriza-
tion of this manifold.
Example 3.1.10. On the coordinate plane, the union of the two axes is not a man-
ifold.
Problems.
3.1.16. The torus can be obtained by rotating around the z axis a circle on the xOz plane
not intersecting the z axis. Show that the torus is a smooth manifold.
Hint: Since S1 can be covered by four neighborhoods, the torus T 2 = S1 × S1 can be covered
p
by 4 × 4 neighborhoods. It can be shown that the torus has the equation ( x2 + y2 − b)2 + z2 = a2
where 0 < a < b.
3.1.17. (a) Consider the union of the curve y = sin 1x , x > 0 and the segment {(0, y)| − 1 ≤
y ≤ 1} (the topologist’s sine curve, see also Section 1.5.1). Is it a manifold?
(b) Consider the curve which is the union of the curve y = x3 sin 1x , x 6= 0 and y = 0
when x = 0. Is this a manifold?
Hint: Consider a neighborhood of a point on the curve on the y-axis. Can it be homeomorphic
to an open neighborhood in R?
3.1.21. * Show that any diffeomorphism from Sn−1 onto Sn−1 can be extended to a diffeo-
morphism from Dn = B0 (0, 1) onto Dn .
Hint: See 1.4.24 and 3.1.20.
84 3. DIFFERENTIAL TOPOLOGY
d(g ◦ f )x = dgy ◦ d fx .
? V @@
~~~
f @@ g
~~ @@
~~ @
U /W
g◦ f
by ϕ(u, v) = (x(u, v), y(u, v), z(u, v)). Consider a point ϕ(u0 , v0 ) on the surface.
Near to (u0 , v0 ) if we fix v = v0 and only allow u to change then we get a parametric
curve ϕ(u, v0 ) passing through ϕ(u0 , v0 ). The velocity vector of the curve ϕ(u, v0 )
is a “tangent vector” to the curve at the point ϕ(u0 , v0 ), and is given by the partial
derivative with respect to u: ∂∂ϕu (u0 , v0 ). Similarly we have another “tangent vector”
∂ϕ
∂ v (u0 , v0 ). Then the “tangent space” of the surface at ϕ(u0 , v0 ) is the plane spanned
the above two tangent vectors (under some further conditions for this notion to be
well-defined).
Example 3.2.5. Consider a surface z = f (x, y). Then the tangent plane at (x, y, f (x, y))
consists of the linear combinations of the vectors (1, 0, fx (x, y)) and (0, 1, fy (x, y)).
Consider the circle S1 . Let (x(t), y(t)) be any curve on S1 . The tangent space
of S1 at (x, y) is spanned by the velocity vector (x0 (t), y0 (t)) if this vector is not 0.
Since x(t)2 +y(t)2 = 1, differentiating both sides with respect to t we get x(t)x0 (t)+
y(t)y0 (t) = 0, or in other words (x0 (t), y0 (t)) is perpendicular to (x(t), y(t)). Thus
the tangent space is perpendicular to the radius.
Proposition 3.2.6. The tangent space does not depend on the choice of parametriza-
tion.
PROOF. Consider
? M `AA 0
ϕ ~~~ AA ϕ
~~~ AA
~~ A
U / U0
ϕ 0−1
◦ϕ
Thus any tangent vector with respect to the parametrization ϕ is also a tangent
vector with respect to the parametrization ϕ 0 . We conclude that the tangent space
does not depend on the choice of parametrization.
Proposition 3.2.9. d fx (h) ∈ T N f (x) and does not depend on the choice of F.
WO /N F
~~> O
f ◦ϕ
~
ϕ
~~~ ψ
~
U −1 / V
ψ ◦ f ◦ϕ
Assume that ϕ(u) = x, ψ(v) = f (x), h = dϕu (w). The diagram induces that
for w ∈ Rm : d fx (dϕu (w)) = dFx dϕu (w) = dψv d(ψ −1 ◦ f ◦ ϕ)u (w) . From this
d(g ◦ f )x = dg f (x) ◦ d fx .
Problems.
3.2.20. (a) Calculate the derivative of the map f : (0, 2π) → S1 , f (t) = (cost, sint).
(b) Calculate the derivative of the map f : S1 → R, f (x, y) = ey .
The Implicit Function Theorem is obtained by setting F(x, y) = (x, f (x, y)) and
applying the Inverse Function Theorem to F.
Theorem 3.3.8 (Inverse Function Theorem for manifolds). Let M and N be two
manifolds of the same dimensions, and let f : M → N be smooth. If x is a regular
point of f then there is a neighborhood in M of x on which f is a diffeomorphism
onto its image.
PROOF. Consider
f
MO /N
O
ϕ ψ
ψ −1 ◦ f ◦ϕ
U /V
The theorem is essentially the Implicit Function Theorem carried over to man-
ifolds.
90 3. DIFFERENTIAL TOPOLOGY
O /W
g
Example 3.3.11. To be able to follow the proof more easily the reader can try to
work it out for an example, such as the case where M is the graph of the function
z = x2 + y2 , and f is the height function f ((x, y, z)) = z defined on M.
The n-sphere Sn is a subset of Rn+1 determined by the implicit equation ∑n+1 2
i=1 xi =
1. Since 1 is a regular value of the function f (x1 , x2 , . . . , xn+1 ) = ∑n+1 2
i=1 xi we con-
clude that Sn is a manifold of dimension n.
Lie groups. The set Mn (R) of n × n matrices over R can be identified with the
2
Euclidean manifold Rn .
Consider the map det : Mn (R) → R. Let A = [ai, j ] ∈ Mn (R). Since det(A) =
∑σ ∈Sn (−1)σ a1,σ (1) a2,σ (2) · · · an,σ (n) = ∑ j (−1)i+ j ai, j det(Ai, j ), we can see that det
is a smooth function.
Let us find the critical points of det. A critical point is a matrix A = [ai, j ] at
which ∂∂ adet
i, j
(A) = (−1)i+ j det(Ai, j ) = 0 for all i, j. In particular, det(A) = 0. So 0 is
the only critical value of det.
Therefore SLn (R) = det−1 (1) is a manifold of dimension n2 − 1.
Furthermore we note that the group multiplication in SLn (R) is a smooth map
from SLn (R) × SLn (R) to SLn (R). The inverse operation is a smooth map from
SLn (R) to itself, because of for example Formula (3.3.1). We then say that SLn (R)
is a Lie group.
3.3. REGULAR VALUES 91
Definition 3.3.12. A Lie group is a smooth manifold which is also a group, for
which the group operations are compatible with the smooth structure, namely the
group multiplication and inversion are smooth.
Let O(n) be the group of orthogonal n × n matrices, the group of linear trans-
formation of Rn that preserves distances.
3.3.15. Show that if a and b are both positive or both negative then f −1 (a) and f −1 (b) are
diffeomorphic.
3.3.17. Show that the height function on the sphere S2 has exactly two critical points. The
height function is the function (x, y, z) 7→ z.
3.3.20. Let dim(M) = dim(N), M be compact and S be the set of all regular values of
f : M → N. For y ∈ S, let | f −1 (y)| be the number of elements of f −1 (y). The map
S → N
y 7 → | f −1 (y)|.
92 3. DIFFERENTIAL TOPOLOGY
is locally constant.
Hint: Each x ∈ f −1 (y) has a neighborhood Ux on which f is a diffeomorphism. Let V =
f (Ux )] \ f (M \ Consider V ∩ S.
T S
[ x∈ f −1 (y) x∈ f −1 (y) Ux ).
3.3.22. Find a counter-example to show that 3.3.9 is not correct if regular value is replaced
by critical value.
3.3.25. Show that the set of all invertible n × n-matrices GL(n; R) is a Lie group and find
its dimension.
3.3.26. (a) Check that the derivative of the determinant map det : Mn (R) → R is repre-
sented by a gradient vector whose (i, j)-entry is (−1)i+ j det(Ai, j ).
(b) Determine the tangent space of SLn (R) at A ∈ SLn (R).
Hint: Use Problem 3.3.23.
(c) Let I be the n × n identity matrix. Show that the tangent space of SLn (R) at the
identity element I is the set of all n × n matrices with zero traces. This is called the Lie
algebra sln (R) of the Lie group SLn (R).
In general if G is a Lie group then the tangent space of G at its identity element is
called the Lie algebra of G, often denoted by g.
3.4. MANIFOLDS WITH BOUNDARY 93
We need to check that our definition of the boundary does not depend on the
choice of the parametrizations. An inspection shows that the question is reduced
to:
Lemma 3.4.3. Suppose that f is a diffeomorphism from the closed unit disk Dn =
B0 (0, 1) in Rn to itself. Then f maps interior points to interior points.
From now on when we talk about a manifold it may be with or without bound-
ary.
Example 3.4.7. Let f be the height function on S2 and let y be a regular value.
Then the set f −1 ((−∞, y]) is a disk with the circle f −1 (y) as the boundary.
94 3. DIFFERENTIAL TOPOLOGY
f
MO /R
yy<
y
yy
yy g
ϕ
yy
U ×V
f
y
x
ϕ g
FIGURE 3.4.1.
Since (U ×V ) \ g−1 (y) consists of two connected components, exactly one of
the two is mapped via g to (y, ∞). In order to be definitive, let us assume that it is
W = {(u, v) | v ≥ h(u)} that is mapped to [y, ∞). Then ϕ : W → N is a parametriza-
tion of a neighborhood of x in N. It is not difficult to see that W is diffeomor-
phic to an open neighborhood of 0 in Hm . In fact the map W → Hm given by
(u, v) 7→ (u, v − h(u)) would give the desired diffeomorphism. Thus x is a bound-
ary point of N.
The tangent space of a manifold with boundary M is defined as follows. It x is
an interior point of M then T Mx is defined as before. If x is a boundary point then
there is a parametrization ϕ : U → M, where U is an open neighborhood of 0 in
Hm and ϕ(0) = x. Then T Mx is still defined as dϕ0 (Rm ).
3.4. MANIFOLDS WITH BOUNDARY 95
x f
y
ϕ g̃
∂Hm
Ũ
T g̃−1 (y)u
FIGURE 3.4.2.
x f
y
U
ϕ g̃
∂Hm
T g̃−1 (y)u
Ũ
FIGURE 3.4.3.
p then the desired result follows from 3.4.8 applied to g̃−1 (y) and p. For that we
need to show that the tangent space T g̃−1 (y)u at u ∈ p−1 (0) is not contained in
∂ Hm . Note that since u ∈ p−1 (0) we have u ∈ ∂ Hm .
Since g̃ is regular at u, the null space of d g̃u on T Ũu = Rm is exactly T g̃−1 (y)u ,
of dimension m − n. On the other hand, g̃|∂ Hm is regular at u, which implies that the
null space of d g̃u restricted to T (∂ Hm )u = ∂ Hm has dimension (m − 1) − n. Thus
T g̃−1 (y)u is not contained in ∂ Hm .
Problems.
f
UO /V
O
ϕ ψ
g
U0 / V0
Theorem 3.5.5 (Smooth Brouwer Fixed Point Theorem). A smooth map from
the disk Dn to itself has a fixed point.
PROOF. Suppose that f does not have a fixed point, i.e. f (x) 6= x for all x ∈ Dn .
The line from f (x) to x will intersect the boundary ∂ Dn at a point g(x). Then
g : Dn → ∂ Dn is a smooth function and is the identity on ∂ Dn . That is impossible,
by 3.5.3.
Actually the Brouwer Fixed Point Theorem holds true for continuous map. The
proof use the smooth version of the theorem. Since its proof does not use smooth
techniques we will not present it here, the interested reader can find it in Milnor’s
book [Mil97]. Basically one approximates a continuous function by smooth ones.
Theorem 3.5.6 (Brouwer Fixed Point Theorem). A continuous map from the disk
Dn to itself has a fixed point.
Problems.
3.5.7. Show that a smooth loop on S2 (i.e. a smooth map from S1 to S2 ) cannot cover S2 .
Also, there is no smooth surjective map from R to Rn with n > 1. In other words, there is
no smooth space filling curves, in contrast to the continuous case (compare 1.5.54).
3.5.8. Prove the Brouwer fixed point theorem for [0, 1] directly.
3.5.9. Check that the function g in the proof of 3.5.5 above is smooth.
3.5.10. Show that the Brouwer fixed point theorem is not correct for (0, 1).
3.5.11. Show that the Brouwer fixed point theorem is not correct for open balls in Rn .
Hint: Let ϕ be a diffeomorphism from Bn to Rn . Let f be any smooth map on Rn . Consider
the map ϕ −1 f ϕ.
3.5.12. Find a smooth map from the solid torus to itself without fixed point.
3.6. Orientation
Orientation on vector spaces. On a finite dimensional real vector space, two
bases are said to determine the same orientation of the space if the change of bases
matrix has positive determinant. Being of the same orientation is an equivalence
relation on the set of all bases. With this equivalence relation the set of all bases is
divided into two equivalence classes. If we choose one of the two as the positive
one, then we say the vector space is oriented and the chosen equivalence class of
bases is called the orientation (or positive orientation). Thus any finite dimensional
real vector space is orientable (i.e. can be oriented) with two possible orientations.
PROOF. If the surface is orientable then its tangent spaces could be oriented
ru × rv
smoothly. That means the unit normal vector is well-defined on the
||ru × rv ||
surface, and it is certainly smooth.
Conversely, if there is a smooth unit normal vector N(p) on the surface then
we can parametrize a neighborhood of any point by a parametrization r(u, v) such
that ru × rv is in the same direction with N (if the given parametrization gives the
reversed orientation we can just switch the variables u and v). This can be done if
we choose the neighborhood to be connected, since then hru × rv , Ni will always be
positive in that neighborhood.
Example 3.6.6. The torus is orientable. This can be seen from an implicit equation
of the torus.
102 3. DIFFERENTIAL TOPOLOGY
3.6.7. Two diffeomorphic surfaces are either both orientable or both un-orientable.
Hint: d( f ◦ r) = d f ◦ dr and r0−1 ◦ r = ( f ◦ r0 )−1 ◦ ( f ◦ r).
3.7. BROUWER DEGREE 103
deg( f , y) = ∑ sign(d fx ).
x∈ f −1 (y)
Notice that the set f −1 (y) is a finite because M is compact (see 3.3.9).
This number deg( f , y) is called the Brouwer degree (bậc Brouwer) 1 or topo-
logical degree of the map f with respect to the regular value y.
From the Inverse Function Theorem 3.3.8, each regular value y has a neigh-
borhood V and each preimage x of y has a neighborhood Ux on which f is a dif-
feomorphism onto V , either preserving or reversing orientation. Therefore we can
interpret that deg( f , y) counts the algebraic number of times the function f covers
the value y.
Example 3.7.1. Consider f : R → R, f (x) = x2 . Then deg( f , 1) = 0. This could
be explained geometrically from the graph of f , as f covers the value 1 twice in
opposite directions at x = −1 and x = 1.
On the other hand, consider g(x) = x3 − x with the regular value 0. From the
graph of f we see that f covers the value 0 three times with positive signs at x = −1
and x = 1 and negative sign at x = 0, therefore we see rightaway that deg(g, 0) = 1.
Homotopy invariance. In this section we will show that the Brouwer degree does
not depend on the choice of regular values and is invariant under smooth homotopy.
Lemma 3.7.2. Let M be the boundary of a compact oriented manifold X, oriented
as the boundary of X. If f : M → N extends to a smooth map F : X → N then
deg( f , y) = 0 for every regular value y.
PROOF. (a) Assume that y is a regular value of F. Then F −1 (y) is a 1-
dimensional manifold of dimension 1 whose boundary is F −1 (y) ∩ M = f −1 (y),
by Theorem 3.4.8.
By the Classification of one-dimensional manifolds, F −1 (y) is the disjoint
union of arcs and circles. Let A be a component that intersects M. Then A is
an arc with boundary {a, b} ⊂ M.
We will show that sign(det(d fa )) = − sign(det(d fb )). Taking sum over all arc
components of F −1 (y) would give us deg( f , y) = 0.
1L. E. J. Brouwer (1881–1966) is a Dutch mathematician. He had many important contributions in
the early development of topology, and founded Intuitionism.
104 3. DIFFERENTIAL TOPOLOGY
We do not present a proof for this lemma. The reader can find a proof in
[Mil97, p. 22].
Then qt (z) is continuous with respect to the pair (t, z). Notice that if t 6= 0 then
qt (z) = t n p((1 − t)t −1 z), and q0 (z) = zn while q1 (z) = an .
If we restrict z to the set {z ∈ C | |z| = 1} = S1 then qt (z) has no roots, so |qqtt (z)
(z)|
is a continuous homotopy of maps from S1 to itself, starting with the polynomial
zn and ending with the constant polynomial |aann | . But these two polynomials have
different degrees, a contradiction.
Vector field.
Example 3.7.10. Let v : S1 → R2 , v((x, y)) = (−y, x), then it is a nonzero (not zero
anywhere) tangent vector field on S1 .
Similarly we can find a nonzero tangent vector field on Sn with odd n.
Theorem 3.7.11 (The Hairy Ball Theorem). There is no smooth nonzero tangent
vector field on Sn if n is even. 2
Problems.
3.7.12. Find the topological degree of a polynomial on R. Notice that although the domain
R is not compact, the topological degree is well-defined for polynomial.
3.7.14. Show that deg( f , y) is locally constant on the subspace of all regular values of f .
3.7.15. What happens if we drop the condition that N is connected in Theorem 3.7.5?
Where do we use this condition?
Compute deg(ri ).
Compute deg( f ).
2This result can be interpreted cheerfully as that on our head there must be a spot with no hair!
3.7. BROUWER DEGREE 107
Compute deg(r).
3.7.22. Show that any polynomial of degree n gives rise to a map from S2 to itself of degree
n.
3.7.24. If f , g : Sn → Sn be smooth such that || f (x) − g(x)|| < 2 for all x ∈ Sn then f is
smoothly homotopic g.
Hint: Note that f (x) and g(x) will not be antipodal points. Use the homotopy
(1 − t) f (x) + tg(x)
Ft (x) = .
||(1 − t) f (x) + tg(x)||
3.7.25. If f , g : Sn → Sn be smooth such that f (x) 6= g(x) for all x ∈ Sn then f is smoothly
homotopic to −g.
3.7.27 (Brouwer fixed point theorem for the sphere). Let f : Sn → Sn be smooth. If
deg( f ) 6= (−1)n+1 then f has a fixed point.
Hint: If f does not have a fixed point then f will be homotopic to the reflection map.
3.7.28. Show that any map of from Sn to Sn of odd degree carries some pair of antipodal
points to a pair of antipodal point.
108 3. DIFFERENTIAL TOPOLOGY
[AF08] Colin Adams and David Fransoza, Introduction to topology: Pure and applied, Prentice
Hall, 2008.
[Atl] Topology Atlas website, http://at.yorku.ca/topology/qintop.htm.
[Bre93] Glen E. Bredon, Topology and Geometry, Graduate Texts in Mathematics, vol. 139,
Springer, 1993.
[Cai94] George L. Cain, Introduction to general topology, Addison-Wesley, 1994.
[Con90] John B. Conway, A course in functional analysis, Springer-Verlag, 1990.
[dC76] Manfredo do Carmo, Differential geometry of curves and surfaces, Prentice-Hall, 1976.
[DFN85] Boris A. Dubrovin, Anatoly T. Fomenko, and Sergey P. Novikov, Modern Geometry–
Methods and Applications, Part II, Graduate Texts in Mathematics, vol. 104, Springer-
Verlag, 1985.
[DFN90] , Modern Geometry–Methods and Applications, Part III, Graduate Texts in Math-
ematics, vol. 124, Springer-Verlag, 1990.
[Duc01] Duong Minh Duc, Topo, Nha Xuat Ban Dai hoc Quoc gia Thanh pho Ho Chi Minh, 2001.
[Dug66] James Dugundji, Topology, Allyn and Bacon, 1966.
[End77] Herbert B. Enderton, Elements of set theory, Academic Press, 1977.
[Eng89] Ryszard Engelking, General topology, Heldermann Verlag, 1989.
[GP74] Victor Guillemin and Alan Pollack, Differential topology, Prentice-Hall, 1974.
[Hat01] Allen Hatcher, Algebraic Topology, Cambridge University Press, 2001.
[Hir76] Morris W. Hirsch, Differential Topology, Graduate Texts in Mathematics, vol. 33,
Springer, 1976.
[HY61] John G. Hocking and Gail S. Young, Topology, Dover, 1961.
[Kel55] John L. Kelley, General topology, Van Nostrand, 1955.
[KF75] A. N. Kolmogorov and S. V. Fomin, Introductory real analysis, Dover Publications Inc.,
New York, 1975.
[Lan93] Serge Lang, Algebra, 3 ed., Addison-Wesley, 1993.
[Lee00] John M. Lee, Introduction to topological manifolds, Springer, 2000.
[Mil97] John Milnor, Topology from the differentiable viewpoint, Princeton landmarks in Mathe-
matics and Physics, Princeton University Press, 1997.
[Mun00] James Munkres, Topology a first course, 2nd ed., Prentice-Hall, 2000.
[Ros99] Dennis Roseman, Elementary topology, Prentice-Hall, 1999.
[Rud86] Walter Rudin, Real and complex analysis, 3 ed., McGraw Hill, 1986.
[SJ70] Lynn A. Steen and J. Arthur Seebach Jr., Counterexamples in topology, Holt, Rinehart
and Winston, 1970.
[Spi65] Michael Spivak, Calculus on manifolds, Addison-Wesley, 1965.
[Spi99] , A comprehensive introduction to differential geometry, vol. 2, Publish or Perish,
1999.
[Vic94] James W. Vick, Homology theory: an introduction to algebraic topology, 2nd ed., Grad-
uate Texts in Mathematics, vol. 145, Springer-Verlag, 1994.
109
110 BIBLIOGRAPHY
[VINK08] O. Ya. Viro, O. A. Ivanov, N. Yu. Netsvetaev, and V. M. Kharlamov, Elementary topology
problem textbook, preprint, 2008.
Index
basis, 12 filter, 33
boundary, 18 filter-base, 33
Brouwer degree, 92 finite intersection property, 35
Brouwer Fixed Point Theorem, 88 fundamental group, 68
111
112 INDEX
smooth, 72 well-ordered, 9
metric space
equivalent, 14 completely regular, 46
Mobius band, 55, 90 first countable, 32
Hausdorff, 29
neighborhood, 11
homeomorphic, 15
net, 31
locally compact, 45
universal, 44
Niemytzki, 52
norm, 11
normal, 29
order quotient, 53
dictionary, 5 regular, 29
lower bound, 5 subspace, 17
minimal, 5 tangent, 76
smallest, 5 Space filling curve, 28
total, 5 sphere, 17
stereographic projection, 17
parametrization, 72 Stone-Cech compactification, 46
path, 25 subbasis, 12
composition, 67 support, 52
inverse, 67 surface, 60
Peano curve, 28 connected sum, 60
point fundamental polygon, 63
contact, 18 genus, 60
interior, 18 non-orientable, 60
limit, 18 orientable, 60
point-wise convergence topology, 37 two-sided, 89
projective plane, 56
projective space, 56 Tiestze extension theorem, 50
topological dimension, 27
regular point, 79
topological invariant, 62
regular value, 79
topological space
relation
connected, 21
equivalence, 5
connected component, 21
retract, 51, 87
Topologist’s sine curve, 26
retraction, 87
topology
Riemann sphere, 46
coarser, 12
Sard Theorem, 87 compact-open, 37
separation axioms, 29 discrete, 10
set Euclidean, 11
cardinal, 6 finer, 12
closed, 10 finite complement, 11
countable, 3 generated by subsets, 12
directed, 31 ordering, 13
equivalence, 2 product, 39
indexed family, 2 quotient, 53
open, 10 relative, 17
ordered, 5 subspace, 17
INDEX 113
trivial, 10
uniform convergence, 38
Zariski, 41
torus, 54
trianglulation, 61
union
disjoint, 40
Urysohn lemma, 49
smooth, 74
Urysohn Metrizability Theorem, 52
vector field, 94
Zorn lemma, 7
You may say I’m a dreamer
But I’m not the only one
I hope someday you’ll join us ...
John Lennon, Imagine.