Lec 06 Topology
Lec 06 Topology
PAUL S CHRIMPF
S EPTEMBER 26, 2013
U NIVERSITY OF B RITISH C OLUMBIA
E CONOMICS 526
This lecture focuses on sequences, limits, and topology. Similar material is covered in
chapters 12 and 29 of Simon and Blume, or 1.3 of Carter.
In example 3, there are two accumulation points, −1 and 1, and you can find subse-
quences that converge to these points.
Lemma 1.2. Let a be an accumulation point of { xn }. Then ∃ a subsequence that converges to a.
Proof. We can construct a subsequence as follows. Let {ϵk } be a sequence that converges
to zero with ϵk > 0∀k, (for example, ϵk = 1/k). By the definition of accumulation point,
for each ϵk ∃ infinitely many xn such that
d ( x n , a ) < ϵk (1)
Pick any xn1 such that (1) holds for ϵ1 . For k > 1, pick nk ̸= n j for all j < k and such that
(1) holds for ϵk . Such an nk always exists because there are infinite xn that satisfy (1). By
construction, limk→∞ xnk = a (you should verify this using the definition of limit). □
Convergence of sequences is often preserved by arithmetic operations, as in the follow-
ing two theorems.
2
LIMITS AND TOPOLOGY OF METRIC SPACES
Theorem 1.1. Let { xn } and {yn } be sequences in a normed vector space V. If xn → x and
yn → y, then
xn + yn → x + y.
Proof. Let ϵ > 0 be given. Then ∃ Nx such that for all n ≥ Nx ,
d( xn , x ) < ϵ/2,
and ∃ Ny such that for all n ≥ Ny ,
d(yn , y) < ϵ/2.
Let N = max{ Nx , Ny }. Then for all n ≥ N,
d( xn + yn , x + y) = ∥( xn + yn ) − ( x + y)∥ ≤ ∥ xn − x ∥ + ∥yn − y∥
<ϵ/2 + ϵ/2 = ϵ.
□
Theorem 1.2. Let { xn } be a sequence in a normed vector space with scalar field R and let {cn }
be a sequence in R. If xn → x and cn → c then
xn cn → xc.
Proof. On problem set. □
In fact, in the next lecture we will see that if f (·, ·) is continuous, then lim f ( xn , yn ) =
f ( x, y). The previous two theorems are examples of this with f ( x, y) = x + y and f (c, x ) =
cx, respectively.
1.1. Series. Infinite sums or series are formally defined as the limit of the sequence of
partial sums.
Definition 1.5. Let { xn }∞ n
n=1 be a sequence in a normed vector space. Let sn = ∑i =1 xi
denote the sum of the first n elements of the sequence. We call sn the nth partial sum. We
define the sum of all the xi s as
∞
∑ xi ≡ nlim
→∞
sn
i =1
This is called a(n infinite) series.
Example 1.6. Let β ∈ R. ∑i∞=0 βi is called a geometric series. Geometric series appear
often in economics, where β will be the subjective discount factor or perhaps 1/(1 + r ).
Notice that
s n =1 + β + β2 + · · · + β n
=1 + β (1 + β + · · · + β n −1 )
= 1 + β ( 1 + β + · · · + β n −1 + β n ) − β n +1
s n (1 − β ) =1 − β n +1
1 − β n +1
sn = ,
1−β
3
LIMITS AND TOPOLOGY OF METRIC SPACES
so,
∞
∑ βi = lim sn
i =0
1 − β n +1
= lim
1−β
1
= if | β| < 1.
1−β
1.2. Cauchy sequences. We have defined convergent sequences as ones whose entries
all get close to a fixed limit point. This means that all the entries of the sequence are also
getting closer together. You might imagine a sequence where the entries get close together
without necessarily reaching a fixed limit.
Definition 1.6. A sequence { xn }∞ n=1 is a Cauchy sequence if for any ϵ > 0 ∃ N such that
for all i, j ≥ N, d( xi , x j ) < ϵ.
It turns that in Rn Cauchy sequences and convergent sequences are the same. This is a
consequence of R having the least upper bound property.
There is a proof of this in Chapter 29.1 of Simon and Blume. If you want more practice
with the sort of proofs in this lecture, it would be good to read that section. The conver-
gence of Cauchy sequences in the real numbers is a consequence of the least upper bound
property that we discussed in lecture 1. Cauchy sequences do not converge in all metric
spaces. For example, the rational numbers are a metric space, and any sequence of ratio-
nals that converges to an irrational number in R is a Cauchy sequence in Q but has no
limit in Q. Having Cauchy sequences converge is necessary for proving many theorems,
so we have a special name for metric spaces where Cauchy sequences converge.
Completeness is important for so many results that complete version of vector spaces
are named after the mathematicians who first studied them extensively. A Banach space
is a complete normed vector space. A Hilbert space is a complete inner product
√ space.
Since any inner product space is a normed vector space with norm ∥ x ∥ = ⟨ x, x ⟩, any
Hilbert space is also a Banach space.
is a Banach space.
4
LIMITS AND TOPOLOGY OF METRIC SPACES
ℓ2 with
∞
⟨ x, y⟩ = ∑ xi yi
i =1
is a Hilbert space.
Showing that ℓ p is complete is slightly tricky because you have deal with a sequence of
xi ∈ ℓ p , each element of which is itself an infinite sequence. You should not worry if you
have difficulty following the rest of this example.
To show that ℓ p is complete, let {xn }∞ n=1 be a Cauchy sequence. Denote the elements of
xi by xi1 , xi2 , .... First, let’s show that for any n, x1n , x2n , ... is a Cauchy sequence in R. Let
ϵ > 0. Since {xn }∞ n=1 is Cauchy, ∃ Nϵ such that for all i, j ≥ Nϵ ,
xi − x j < ϵ.
Since ( )
∞
∑
p p
xi − x j = xim − x jm
m =1
All terms in the sum on the right are non-negative and the sum includes xin − x jn , so
p p
xin − x jn ≤ xi − x j
xin − x jn ≤ xi − x j
Therefore, xin − x jn < ϵ for all i, j ≥ Nϵ , i.e. x1n , x2n , ... is a Cauchy sequence in R. R is
complete, so it has some limit. Denote the limit by xn∗ .
Now we will show that x∗ = ( x1∗ , x2∗ , ...) is the limit of {xn }∞ n=1 . First, we should show
∗
that x ∈ ℓ . Let
p
m
s∗m = ∑ |xn∗ | p .
n =1
We need to show that lim s∗m
exists. Since {xn }∞
n=1 is Cauchy, ∃ j such that if i ≥ j,
xi − x j < 1. Using the triangle inequality,
∥ xi ∥ ≤ xi − x j + x j = 1 + x j ≡ M
for all i ≥ j and some fixed j. Thus, ∥xi ∥ ≤ M for some constant M and all i ≥ j. Then,
m
s∗m = lim ∑ |xin | p ≤ M p
i → ∞ n =1
2. O PEN SETS
Definition 2.1. Let X be a metric space and x ∈ X. A neighborhood of x is the set
Nϵ ( x ) = {y ∈ X : d( x, y) < ϵ.
A neighborhood is also called an open ϵ-ball of x and written Bϵ ( x ).
Definition 2.2. A set, S ⊆ X is open if ∀ x ∈ S, ∃ ϵ > 0 such that
Nϵ ( x ) ⊂ S.
For every point in an open set, you can find a small neighborhood around that point
such that the neighborhood lies entirely within the set.
Example 2.1. Any open interval, ( a, b) = { x ∈ R : a < x < b}, is an open set.
Example 2.2. Any linear subspace of dimension k < n in Rn is not open.
Theorem 2.1.
(1) Any union of open sets is open. (finite or infinite)
(2) The finite intersection of open sets is open.
Proof. Let S j , j ∈ J be a collection of open sets. Pick any j0 ∈ J. If x ∈ ∪ j∈ J S j , then there
must be ϵ j0 > 0 such that Nϵj ( x ) ⊂ S j0 . It is immediate that Nϵj ( x ) ⊂ ∪ j∈ J S j as well.
0 0
Let S1 , .., Sk be a finite collection of open sets. For each i ∃ϵi > 0 such that Nϵi ( x ) ⊂ Si .
Let ϵ = mini∈{1,...,k} ϵi . Then ϵ > 0 since it is the minimum of a finite set of positive
numbers. Also, Nϵ ( x ) ⊂ Si for each i, so Nϵ ( x ) ⊂ ∩ik=1 Si . □
Definition 2.3. The interior of a set A is the union of all open sets contained in A. It is
denoted as int( A).
From the previous, theorem, we know that the interior of any set is open.
Example 2.3. Here some examples of the interior of sets in R.
(1) A = ( a, b), int( A) = ( a, b).
6
LIMITS AND TOPOLOGY OF METRIC SPACES
3. C LOSED SETS
A closed set is the opposite of an open set.
Definition 3.1. A set S ⊆ X is closed if its complement, X c , is open.
Theorem 3.1.
(1) The intersection of any collection of closed sets is closed.
(2) The union of any finite collection of closed sets is closed.
( )c
Proof. Let Cj , j ∈ J be a collection of closed sets. Then ∩ j∈ J Cj = ∪ j∈ J Cjc . Cjc are open, so
( )c
by theorem 2.1, ∪ j∈ J Cjc = ∩ j∈ J Cj = is open.
The proof of part 2 is similar. □
Example 3.1 (Closed sets). Some examples of closed sets include
(1) [ a, b] ⊆ R
(2) Any linear subspace of Rn
(3) {( x, y) ∈ R2 : x2 + y2 ≤ 1}
Closed sets can also be defined as sets that contain the limit of any convergent sequence
in the set. Simon and Blume use this definition. The next theorem shows that their defi-
nition is equivalent to ours.
Theorem 3.2. Let { xn } be any convergent sequence with each element contained in a set C. Then
lim xn = x ∈ C for all such { xn } if and only if C is closed.
Proof. First, we will show that any set that contains the limit points of all its sequences is
closed. Let x ∈ C c . Consider N1/n ( x ). If for any n, N1/n ( x ) ⊂ C c , then C c is open, and
C is closed as desired. If for all n, N1/n ( x ) ̸⊂ C c , then ∃yn ∈ N1/n ( x ) ∩ C. The sequence
{yn } is in C and yn → x. However, by assumption C contains the limit of any sequence
within it. Therefore, there can be no such x, and C c must be open and C is closed.
Suppose C is closed. Then C c is open. Let { xn } be in C and xn → x. Then d( xn , x ) → 0,
and for any ϵ > 0, ∃ xn ∈ Nϵ ( x ). Hence, there can be no ϵ neighborhood of x contained in
C c . C c is open by assumption, so x ̸∈ C c and it must be that x ∈ C. □
Definition 3.2. The closure of a set S, denoted by S (or cl(S)), is the intersection of all
closed sets containing S.
Example 3.2. If S is closed, S = S.
4. C OMPACT SETS
Definition 4.1. An open cover of a set S is a collection of open sets, { Gα } α ∈ A such that
S ⊂ ∪α∈A Gα .
Example 4.1. Some open covers of R are:
• {R}
• {(−∞, 1), (−1, ∞)
• {..., (−3, −1), (−2, 0), (−1, 1), (0, 2), (1, 3), ...}
• {( x, y) : x < y}
The first two are finite open covers since they consist of finitely many open sets. The third
is a countably infinite open cover. The fourth is an uncountably infinite open cover.
Example 4.2. Let X be a metric space and A ⊆ X. The set of open balls of radius ϵ centered
at all points in A is an open cover of A. If A is finite / countable / uncountable, then this
open cover will also be finite / countable / uncountable.
Open covers of the form in the previous example are often used to prove some property
applies to all of A by verifying the property in each small Nϵ ( x ). Unfortunately, this often
involves taking a maximum or sum of something for each set in the open cover. When
the open cover is infinite, it can be hard to ensure that the infinite sum or maximum stays
finite. When we have a finite open cover, we know that things will remain finite.
Definition 4.2. A set K is compact if every open cover of K has a finite subcover.
By a finite subcover, we mean that there is finite set Gα1 , ...Gαk such that S ⊂ ∪kj=1 Gα j .
Compact sets are a generalization of finite sets. Many facts that are obviously true of
finite sets are also true for compact sets, but not true for infinite sets that are not compact.
Suppose we want to show a set has some property. If the set is compact, we can cover it
with a finite number of small ϵ balls and then we just need to show that each small ball
has the property we want. We will see many concrete examples of this technique in the
next few weeks.
8
LIMITS AND TOPOLOGY OF METRIC SPACES
Example 4.3. R is not compact. {..., (−3, −1), (−2, 0), (−1, 1), (0, 2), (1, 3), ...} is an infinite
cover, but if we leave out any single interval (the one beginning with n) we will fail to
cover some number (n + 1).
Example 4.4. Let K = { x }, a set of a single point. Then K is compact. Let { Gα }α∈A be an
open cover of K. Then ∃ α such that x ∈ Gα . This single set is a finite subcover.
Example 4.5. Let K = { x1 , ..., xn } be a finite set. Then K is compact. Let { Gα }α∈A be an
open cover of K. Then for each i, ∃ αi such that xi ∈ Gαi . The collection { Gα1 , ...Gαn } is a
finite subcover.
Example 4.6. (0, 1) ⊆ R is not compact. {(1/n, 1)}∞
n=2 is an open cover, but there can be
no finite subcover. Any finite subcover would have a largest n and could not contain, e.g.
1/(n + 1).
Example 4.7. Let x ∈ V, a normed vector space. Let K = { x 21 , x 23 , x 34 , ...}. Then K is not
1 ) for n = 1, 2, .... Assuming x ̸ = 0, each
n
compact. Consider the open cover N∥ x∥ 1 ( x n+
3( n +2)2
of these neighborhoods contains exactly one point of K, so there is no finite subcover.
Before using compactness, let’s investigate how being compact relates to other proper-
ties of sets, such as closed/open.
Lemma 4.1. Let X be a metric space and K ⊆ X. If K is compact, then K is closed.
Proof. Let x ∈ K c . The collection { Nd( x.y)/3 (y)}, y ∈ K is an open cover of K. K is compact,
so there is a finite subcover, Nd( x,y1 )/3 (y1 ), ..., Nd( x,yn )/3 (yn ). For each i, Nd( x,yi )/3 (yi ) ∩
Nd( x,yi )/3 ( x ) = ∅, so
∩in=1 Nd(x,yi )/3 ( x )
is an open neighborhood of x that is contained in K c . K c is open, so K is closed. □
Lemma 4.2. Let X be a metric space, C ⊆ K ⊆ X. If K is compact and C is closed. Then C is also
compact.
Proof. Let { Gα }α∈A be an open cover for C. Then { Gα }α∈A plus C c is an open cover for K.
Since K is compact there is a finite subcover. Since C ⊆ K, the finite subcover also covers
C. Therefore, C is compact. □
The definition of compactness is somewhat abstract. We just saw that compact sets are
always closed. Another property of compact sets is that they are bounded.
Definition 4.3. Let X be a metric space and S ⊆ X. S is bounded if ∃ x0 ∈ S and r ∈ R
such that
d( x, x0 ) < r
for all x ∈ S.
A bounded set is one that fits inside an open ball of finite radius. For subsets of R this
definition is equivalent to there being a lower and upper bound for the set. For subsets of
a normed vector space, if S is bounded then there exists some M such that ∥ x ∥ < M for
all x ∈ S.
9
LIMITS AND TOPOLOGY OF METRIC SPACES
Theorem 4.1 (Heine-Borel). A set S ⊆ Rn is compact if and only if it is closed and bounded.
We saw that closed sets contain the limit points of all their convergent sequences. There
is also a relationship between compactness and sequences.
13