Introanalysis
Introanalysis
Introduction to Analysis
Donald J. Estep
Department of Mathematics
Colorado State University
Fort Collins, CO 80523
[email protected]
http://www.math.colostate.edu/∼estep
Contents
Chapter 1. Introduction 1
1.1. Review of Real Numbers 1
1.2. Functions and Sets 3
1.3. Unions and Intersection 5
1.4. Cardinality 5
Chapter 3. Compactness 20
3.1. Compactness Arguments and Uniform Boundedness 20
3.2. Sequential Compactness 24
3.3. Separability 25
3.4. Notions of Compactness 28
3.5. Some Properties of Compactness 32
3.6. Compact Sets in Rn 33
Chapter 5. Sequences in Rn 42
5.1. Arithmetic Properties and Convergence 42
5.2. Sequences in R and Order. 45
5.3. Series 46
Introduction
is a set of numbers listed in a specific order. Here, “∞” denotes a neverending list.
The subscript “i” or “bike” is called the index.
Example 1.1.2.
1
= .111111
9
So,
(.1, .11, .111, ..., .11...1n , ...)
| {z }
n 10 s
1
2 1. INTRODUCTION
1
is a sequence of reals that approximates 9 in the sense that
1
− .1...1n < 10−n
9
for all n.
Sequences can have many forms.
Example 1.1.3.
(2, 4, 6, ...) = {2n}∞
n=1
(−1, 1, −1, 1, ...) = {(−1)n }∞
n=1
1 1 1 ∞
(1, , , ...) = { }n=1
2 3 n
(1, 1, 1, ...) = {1}∞
n=1
3
(3, 4.182, , −14.68, ...) = {?}∞n=1
5
Some of these sequences have the special property that as the index increases,
the corresponding numbers become closer and closer to one number.
Definition 1.1.4. If the numbers in the sequence {an } approach a number a
as n increases, then we call a the limit of the sequence and write
After making this definition, the usual thing to do is to prove some basic prop-
erties. Some of these are:
Theorem 1.1.10. Limits of Sequences of Real Numbers
(1) Limits are unique.
(2) Arithmetic Properties: Suppose an → a and bn → b. Then
(a) an + bn → a + b.
(b) an bn → ab.
An important concept related to convergence is the notion of a Cauchy se-
quence.
Definition 1.1.11. A sequence {an } of real numbers is a Cauchy Sequence,
or is Cauchy, if for every > 0, there is an N such that
|ai − aj | < for i, j > N.
The interest in the Cauchy criterion (Def. 1.1.11) is that it does not require the
value of the limit, i.e., what the sequence converges to. The standard criterion for
convergence is practically useless. We are able to prove
Theorem 1.1.12. A sequence of real numbers converges if and only if it is
Cauchy.
Sequences and limits are inherent to the notion of the real number system.
Recall that rational numbers, in general, have infinite decimal expansions. But,
at least these decimal expansions are periodic, e.g., .123412341234... . Most real
numbers (in a precise sense) have infinite, non-repeating decimal expansions. These
are called the irrational numbers. All kinds of difficulties arise because
√ of this fact.
For example, how should we add the irrational numbers e and 2?
We add numbers with finite decimal expansions by starting at the rightmost
digit. But, that is impossible with irrationals! √What we do in practice is add
finite truncations,√or approximations, of e and 2, and count on this being an
approximation of 2 + e.
In fact, this is exactly the way we construct real numbers from the rationals,
which we understand better. Just as with ”+” above, the main point is to show the
real numbers inherit the usual properties of rational numbers. The tricky part is
that we can’t write down reals until we have defined them! This is why the Cauchy
criterion is so important.
All of these difficulties
√ carry over to defining functions of real numbers (how
should we compute e?) and the central role of convergence is why continuity is
such an important property.
In this course, we replace real numbers by a more abstract space, which includes
R and Rn as well as other collections like functions. Then we develop the basic
properties of the space and functions on the space.
f (C) = {f (c) : c ∈ C}
f −1 (C) = {a ∈ A | f (a) ∈ C}
{an }∞
n=1 = (a1 , a2 , ...)
Definition 1.2.7. Let A and B be sets and suppose that for each a ∈ A there
is associated a subset of B called Ca . The set whose elements are the sets Ca for
a ∈ A is denoted
{Ca | a ∈ A} = {Ca }a∈A = {Ca }
Example 1.3.2.
[
{−1, 2, 5, 8} {3, 5, 8, 10} = {−1, 2, 3, 5, 8, 10}
\
{−1, 2, 5, 8} {3, 5, 8, 10} = {5, 8}
Example 1.3.3. Let A = (0, 1] and Ca = (0, a) for a ∈ A. Then
Sa ⊂ Cb if and only if 0 < a ≤ b < 1
(1) C
(2) Ta∈A Ca = C1
(3) a∈A Ca = ∅ (the empty set)
S T
Exercise 1.3.4. What is a∈A Ca and a∈A Ca in Example 1.2.8?
T
Definition 1.3.5. If A and B are two sets and A B is not empty, then we
say A and B intersect. Otherwise, they are disjoint.
S T
There are many important properties of and that you can read on page
28 of Rudin’s “Principles of Mathematical Analysis”.
1.4. Cardinality
Cardinality refers to the “number of points” in a set. “Number of points” is in
quotes because we want to talk about infinite sets as well. Moreover, we want to
distinguish different kinds of infinite sets.
Example 1.4.1. { 1, 2, 5,π, -8}, Z , and R turn out to have different cardinal-
ities.
We start the discussion by developing a mechanism for comparing the number
of elements between two sets. We use functions:
Definition 1.4.2. If there is a 1-1 map of a set A onto a set B, then we say that
A and B are in 1-1 correspondence, have the same cardinality or cardinal
number, or are equivalent, and we write A B.
6 1. INTRODUCTION
Figure 1.1
Next we show that the rational numbers, Q, are countable. To do this, we show
the following:
Theorem 1.4.11. Let A be a countable set. Let Bn be the set of n-tuples
(a1 , ..., an ), where ai ∈ A for 1 ≤ i ≤ n, and the {ai } in an n-tuple need not be
distinct. Then Bn is countable.
Proof. By induction. Since B1 = A, B1 is countable. For the induction step,
assume Bn−1 is countable where n ∈ {2, 3, ...}. The elements of Bn correspond
to {(b, a) | b ∈ Bn−1 , a ∈ A}. In particular, for any fixed b ∈ Bn−1 , the set
{(b, a)} | a ∈ A} ≡ A and hence is countable. Thus, Bn is the countable union of
countable sets.
Corollary 1.4.12. Q is countable.
Proof. Consider A = Z and n = 2 in Theorem 1.4.11. Then Q = { ab | a, b ∈
Z}, corresponding to a subset of {(a, b) | a, b ∈ Z}. We know the latter set is
countable, so by Theorem 1.4.9, we have that Q is countable.
Next, we will prove a fact about the cardinality of R.
Theorem 1.4.13. R is not countable.
Proof. The proof has one tricky point, which is that a real number can have
two decimal expansions. For example, .20 = .19.
It suffices to show that the set of numbers {x | 0 < x < 1} = (0, 1) is uncount-
able. If these numbers were countable, then there would be a sequence {sn }∞ n=1
that gives them all. We show this is impossible by constructing a number in (0, 1)
but not in {sn }∞n=1 .
We write each sn as a decimal expansion:
sn = 0.dn1 dn2 dn3 ...
where dni ∈ {0, 1, 2, ..., 9} for each i and n. Now define y = 0.e1 e2 e3 ..., where
(
1 if dnm 6= 1
em = .
2 if dnm = 1
/ {sn }∞
Notice y ∈ (0, 1), but y ∈ n=1 (see Example 1.4.14).
Example 1.4.14. Let
s1 = .123456789...
s2 = .246812461...
s3 = .691284823...
s4 = .444444444...
..
.
y = .2121...
Now y is different than s1 in the first digit, s2 in the second, s3 in the third, and
so on. Note that situations like sn = .199999... and y = .2000... cannot occur by
the choice of {em }.
1.4. CARDINALITY 9
Metric Spaces
The real numbers have properties that make it natural to discuss sequences.
Indeed, the reals are defined in terms of sequences. We would like to create an
abstract notion of a space in which it makes sense to talk about sequences and in
which sequences play as an important of a role as they do in the real numbers. In
a sense, we would like to abstract the ”sequential nature” of the reals, as opposed
to their other properties.
Our abstract space should certainly include the reals as an example. It should
also include Rn . So, we have to lose the order properties (this is, properties pertain-
ing to the relation “¡”) in our abstract space. But, Rn itself has special properties
that will not be present in the abstract notion. Therefore, Rn will always be a
special, important example of our abstract notion of a space.
(a) k x k≥ 0 and k x k= 0 ⇔ x = 0
(b) k αx k= |α| k x k
(c) (Triangle Inequality) k x + y k≤k x k + k y k
(d) (Cauchy-Schwarz Inequality)|x · y| ≤k x k k y k
(4) Distance Structure Finally, the norm induces a natural distance func-
tion. We define the distance between x and y in Rn as
d(x, y) =k x − y k
This distance function satisfies
(a) d(x, y) ≥ 0 and d(x, y) = 0 ⇔ x = y
(b) d(x, y) = d(y, x)
(c) d(x, z) ≤ d(x, y) + d(y, z)
for x, y, z ∈ Rn .
The distance function is certainly critical for talking about the con-
vergence of sequences of points in Rn : Convergence is tied to the notion
of points becoming closer and distance between points becoming smaller.
To define our abstract space, we skip past the other properties of Rn
and go right to distance.
(0,1)
d(x, y) =k x − y k
(-1,0) (1,0)
d(x, y) =k x − y k∞
d(x, y) =k x − y k1
(0,-1)
d(x,y)
Exercise 2.2.13. Verify that the function given in Def. 2.2.12 is a metric.
P∞
Notice that the terms in i=1 x2i are positive. This implies, among other
things, that
∞
X N
X
lim xi = 0 and x2i = lim x2i
i→∞ N →∞
i=1 i=1
In view of the latter fact, we can think of l2 as something like (Rn , k · k) where
“n → ∞”. (Recall the usual definitions of +, etc. for sequences.)
v
u∞
uX
Theorem 2.2.15. With d(x, y) = t (xi − yi )2 for x, y ∈ l2 , (l2 , d) is a
i=1
metric space.
v
u∞ 2
uX
Note: Actually, k x k= t xi defines a norm on l2 , and
i=1
d(x, y) =k x − y k. However, we do not pursue that.
Proof. The proof is not so hard, except that we have to be careful to treat
the infinite sequences in a mathematically correct way. For example, to verify (1)
of Defn. 2.2.1, we want to show that d(x, y) ≥ 0 and d(x, y) = 0 ⇔ x = y (two
sequences are equal if and only if all the elements are equal) for x, y ∈ l2 .
However, we first have to show that d(x, y) is defined and finite! For this we
require the following basic inequality:
Lemma 2.2.16. Let a, b ∈ R. Then 2|a||b| ≤ a2 + b2 .
Proof: 0 ≤ (|a| − |b|)2 = |a|2 − 2|a||b| + |b|2 .
Now, for x, y ∈ l2 ,
XN
{ (xi − yi )2 }∞ n=1
i=1
is a monotone, nondecreasing sequence. If we prove the terms are bounded above
uniformly for all N , then it has a limit, i.e.,
∞
X
lim (xi − yi )2
N →∞
i−1
So,
max |f (x) − g(x)| ≤ max (|f (x) − g(x)| + |g(x) − h(x)|)
a≤x≤b a≤x≤b
≤ max |f (x) − g(x)| + max |g(x) − h(x)|,
a≤x≤b a≤x≤b
Nr1 (x)
Nr2 (x)
G2
G2
Figure 2.1
Compactness
open cover
a b c
A1 ⊃ A2 ⊃ A3 ⊃ ....
∞
\
Cantor asked: Under what conditions do we have An 6= ∅?
n=1
3.1. COMPACTNESS ARGUMENTS AND UNIFORM BOUNDEDNESS 23
In these two examples, the sets in the sequence are not open and bounded
respectively. Cantor proved
Theorem 3.1.28. Let {An } be a descending sequence of nonempty, closed, and
bounded subsets of R. Then
\∞
An 6= ∅.
n=1
Again, we present a more general result later. For now, we use this property
for our third proof of Theorem 3.1.7.
Proof. Suppose f is not uniformly bounded on A. Since A is bounded, A ⊂
[a, b] for some a, b. Divide [a, b] into two sub-intervals of length (b−a) 2 . f must be
unbounded on at least one of these sub-intervals. Call this sub-interval [a1 , b1 ].
Now repeat the division argument to get a new subinterval [a2 , b2 ] of [a1 , b1 ] of
length (b−a)
T
22 such that f is unbounded on [a2 , b2 ] A. Inductively, we obtain a
sequence {[an , bn ]} with (bn − an ) = b−a bn ] ⊂ [an−1 , bn−1 ] (so the sequence
2n , [an ,T
is descending), and f is unbounded on [an , bn ] A.
∞
\
By Theorem 3.1.28, there is a point x ∈ [an , bn ]. x is a limit point of A
n=1
(why?) and so x ∈ A. By local boundedness,
T there are δx and Mx such that
|f (y)| ≤ Mx for y ∈ (x − δx , x + δx ) A. But, for n sufficiently large, [an , bn ] ⊂
(x − δx , x + δx ), which gives a contractiction.
The last property does not really give a new compactness argument. It is
a restatement of the Bolzano-Weierstrass property in terms of Cauchy sequences.
(Recall Definition 1.1.11.)
We will base our fourth proof of Theorem 3.1.7 on the following theorem:
24 3. COMPACTNESS
Proof. First we prove closed. Suppose x were a limit point of K. Then there
exists a sequence of elements in K that converge to x. This sequence must have
a subsequence that converges to an element in K. However, the sequence and the
subsequence must have the same limit, so x must be in K.
Second, we prove bounded. Assume K is not bounded. Choose x0 ∈ K. For
each n ∈ N, let xn ∈ K with d(x0 , xn ) > n. Now, the sequence {xn } has the
property that every subsequence is unbounded, and hence cannot converge. So, K
cannot be sequentially compact.
3.3. Separability
We saw that sequential compactness imples being closed and bounded, and
the closed and bounded sets in R are special with respect to making compactness
arguments. Another property of real numbers is that the rational numbers are
dense. This turns out to be even more fundamentally important than is obvious.
It also turns out that sequential compactness is related to this property.
Definition 3.3.1. A metric space (X, d) is separable if it contains a count-
able, dense subset. A subset A ⊂ X of a metric space (X, d) is separable if it is
separable considered as a metric space with the metric d.
Example 3.3.2. (R, | · |) and (Rn , k · k) are separable. So are closed, bounded
subsets of these metric spaces.
Theorem 3.3.3. Let K ⊂ X be a sequentially compact subset of a metric space
(X, d). Then K is separable.
We prove this by introducing a useful notion and proving another theorem.
Definition 3.3.4. Let A ⊂ X, wehre (X, d) is a metric space. If for > 0
there are points x1 , ..., xn ∈ A such that
n
[
A⊂ N (xi ),
i=1
Example 3.3.5. Given any set of real numbers that is bounded and any > 0,
we can find an -net consisting of points with finite decimal expansions. This is
very important for computation with real numbers on computers.
Note that the existence of an -net for any > 0 is a stronger condition than
boundedness. Not only is K contained in some large ball, which S is obtained by
choosing some -net {x1 , ..., xn } and taking a ball that contains i N (xi ), but it
is contained in the union of a finite number of neighborhoods of any small size.
Example 3.3.6. Consider the closed unit “ball” in l2 (see Definition 2.2.14.):
v
u∞
uX
N1 (0) = {x ∈ l2 | t x2n ≤ 1}.
n=1
n−1
[
We can choose a sequence of points {xm }∞
m=1 in K such that xn 6∈ N (xm )
m=1
for all n. This sequence can have no convergent subsequence. This is because if
{xmk }∞
k=1 → x, then there exists l such that k > l implies d(x, xmk ) < 2 . But now
xl+1 ∈ N (xl ), contradicting our choice of xl+1 . This proves the contrapositive.
Exercise 3.3.10. This proof is stated a little more easily if we use the notion
of Cauchy sequences and the fact that a sequence that converges is Cauchy. Do
this.
1
Proof. Of Theorem 3.3.3. By Theorem 3.3.9, for = m, m ∈ N, there is an
-net for K. Call it
{xm,1 , xm,2 , ..., xm,nm }.
So,
n
[m
K⊂ N m1 (xm,i ).
i=1
The set of points
A = {x1,1 , ..., x1,n1 , x2,1 , ..., x2,n2 , x3,1 , ..., x3,n3 , x4,1 , ...}
is at most countable and dense in K.
A is at most countable because it is the countable union of finite sets. (The
sets may intersect, so the end result could be finite.)
We need a little more justification to say A is dense in K. We have to show
that every point in K − A is a limit point of A. Assume x ∈ K, x 6∈ A. (Note,
if K is finite, then x ∈ A is forced.) We construct a sequence of points in A that
converges to x.
For m ∈ N, choose xm = xm,i such that 1 ≤ i ≤ nm and x ∈ N m1 (xm,i ). Then
we see that {xi }∞ ∞
i=1 ∈ A and {xi }i=1 → x.
Summing up, so far we have that sequential compactness implies closed, bounded,
totally bounded, and separable.
There is one last fact about separability we will use. Separability is related to
open covers.
Definition 3.3.11. An open cover of a set A contained in a metric space
(X, d) is a collection of open subsets {Gα }α∈a of X such that
[
A⊂ Gα .
α∈a
1
n x
r
4
xl
Figure 3.1
We now show part of the Borel-Lebesgue Theorem. The rest will be proven in
Chapter 4.
Theorem 3.4.4. Borel-Lebesgue Theorem, part I
Let (X, d) be a metric space, and K ⊂ X. The following are equivalent.
(1) K is compact.
(2) Every collection of closed subsets of K with the finite intersection property
has a nonempty intersection.
(3) K is sequentially compact.
In light of this result, we only use the term “compact” and no longer refer to
“sequential compactness”. We may restate the results in Section 3.2 and 3.3:
Theorem 3.4.5. Let K ⊂ X be a compact subset of a metric space (X, d).
Then,
(1) K is closed.
(2) K is bounded.
(3) K is separable.
(4) K is totally bounded.
Proof. Of Theorem 3.4.4
We show that (1) → (2) → (3) → (1).
(1) → (2)
We assumeT K is compact. Let {Fα }α∈a be a collection of closed subsets of
K such that α∈a Fα = ∅. Consider the open sets {Gα }α∈a with Gα = Fαc , the
complement of Fα in X. Then,
[ [ \
Gα = Fαc = ( Fα )c = ∅c = X.
α α α
So, [
K⊂ Gα .
α
By compactness, [ [
K ⊂ Gα1 ... Gαn
for some α1 , ..., αn . Moreover, since Fαi ⊂ K,
F αi = K \ Gαi
so
\ \ \\
F α1 ... F αn = K \ Gα1 ... K \ Gαn
[ [
= K \ (Gα1 ... Gαn ) = K \ K = ∅.
Hence, {Fα }α∈a cannot have the finite intersection property.
Example 3.4.6. Consider [0, 1] ⊂ (R, | · |). Define
1 1
Fn = [ , ], n ∈ N,
n+3 n+1
so {Fn } = {[ 41 , 12 ], [ 51 , 13 ], [ 61 , 14 ], [ 17 , 15 ], ...} (see Figure 3.2).
Now, if Gn = Fnc , then
1 [ 1
Gn = (−∞, ) ( , ∞)
n+3 n+1
30 3. COMPACTNESS
1 1 1 1 1 1
0 7 6 5 4 3 2
1
F1
F2
F3
F4
Figure 3.2
or
1 [ 1
G1 = (−∞, ) ( , ∞)
4 2
1 [ 1
G2 = (−∞, ) ( , ∞)
5 3
1 [ 1
G3 = (−∞, ) ( , ∞)
6 4
1 [ 1
G4 = (−∞, ) ( , ∞)
7 5
∞
\ T
Clearly, Fn = ∅, since F4 F1 = ∅.
n=1
Moreover, plotting {Gn } shows that
[ [ [
[0, 1] ⊂ G1 G2 G3 G4 .
{G1 , G2 , G3 , G4 } is the finite subcover constructed in the proof (see Figure 3.3).
Back to the proof: (2) → (3)
Let {xm } be a sequence in K. Define
This is a descending sequence of closed sets with the finite intersection property,
since
\ \
xn0 ∈ Fn1 ... Fnk
when n0 = max{n1 , ..., nk } for any n1 , ..., T
nk ∈ N.
∞
Hence, there is an x ∈ K with x ∈ n=1 Fn . There are two possibilities. If
x = xn for infinitely many n, the we extract the subsequence consisting of repeated
valued of xn , which obviously converges to x = xn .
If x = xn for finitely many n, then we construct a subsequence that converges
to x as follows: Choose n1 large enough that x 6= xm for m ≥ n1 . Given nk ∈ N,
us the fact that
x ∈ {xnk +1 , xnk +2 , xnk +3 , ...}
1
is a limit point to choose nk+1 > nk with 0 < d(x, xnk +1 ) < k+1 . Clearly, {xnk } →
x.
3.4. NOTIONS OF COMPACTNESS 31
1 1 1 1 1 1
0 7 6 5 4 3 2
1
Figure 3.3
= { 21 , 13 , ...} {0}
S
F2
= { 31 , 14 , ...} {0}
S
F3
..
.
T
Fn = {0}
= { 21 , 1, 13 , ...} {0}
S
F2
= {1, 13 , 1, 14 , ...} {0}
S
F3
..
.
T
Fn = {1, 0}
G1
G2
G3
Figure 3.4
For each l, {Im,l }m∈N is a descending sequence of intervals in R1 . Hence, there are
real numbers xl such that
am,l ≤ xl ≤ bm,l
for 1 ≤ l ≤ n, m ∈ N. We have x = (x1 , x2 , ..., xn ) ∈ Im for all m.
Now we prove that the n-cell
I = {x ∈ Rn | an ≤ xm ≤ bm , 1 ≤ m ≤ n}
is compact. Set δ =k b − a k, so
k x − y k≤ δ
for all x, y ∈ I. Suppose there is an open cover {Gα }α∈a of I that contains no finite
subcover. Set
am + b m
cm =
2
for 1 ≤ m ≤ n. The intervals {[am , cm ], [cm , bm ]} determine 2n n-cells {Il } whose
union is I.
One of these n-cells, at least, cannot be covered by any finite collection from
{Gα }α∈a . Call this J1 . We next subdivide J1 into 2n n-cells in the same way.
Again, one of the resulting n-cells cannot be covered by any finite collection from
{Gα }α∈a , and we call this J2 .
Inductively, we obtain a sequence of n-cells {Jm }∞ m=1 with the properties
(1) J1 ⊃ J2 ⊃ ...
(2) Jm is not covered by any finite collection from {Gα }.
(3) If x ∈ Jm and y ∈ Jm , then k x − y k≤ 2−m δ.
T∞
By the discussion above, we know there is a point x ∈ m=1 JM , x ∈ Gα for
some α. Since Gα is open, there is an r > 0 with Nr (x) ⊂ Gα . If we choose m with
2−m δ < r, then Jm ⊂ Gα . But, this is a contradiction.
From this, it is easy to prove
Theorem 3.6.3. A set K ⊂ Rn is compact if and only if it is closed and
bounded.
Proof. By Theorem 3.4.5, if K is compact, then it is closed and bounded. On
the other hand, if K is bounded, then it is contained in some compact n-cell. Since
K is a closed subset of a compact set, Theorem 3.5.4 shows it is compact.
These last two theorems complete the proofs of Theorems 3.1.12(Bolzano-
Weierstrass property), 3.1.21(Heine-Borel property), and 3.1.28(Cantor Intersection
Property).
CHAPTER 4
Theorem 4.1.2. Let {xn } be a sequence in a metric space (X, d). {xn } con-
verges to x ∈ X if and only if every neighborhood of x contains all but finitely many
terms of {xn }
Proof. Suppose {xn } → x and let N (x) be a neighborhood of x. By con-
vergence, there is an N such that d(xn , x) < for n ≥ N . Hence, xn ∈ N (x) for
n ≥ N.
Now assume every neighborhood of x contains all but a finite number of {xn }.
For > 0, consider N (x). By assumption, there is an N > 0 such that xn ∈ N (x),
i.e., d(xn , x) < for n ≥ N .
We have
d(x, xnm ) ≤ d(x, y) + d(y, xnm )
≤ 2−m δ + 2−m δ = 21−m δ
We conclude that {xnm } → x.
Finally, we discuss the connection between boundedness (Defn. 2.3.24) and
convergence of sequences.
Definition 4.1.4. A sequence {xn } in a metric space (X, d) is bounded if its
range forms a bounded set in X. Otherwise, it is unbounded. Equivalently, {xn }
is bounded if and only if there exists A ⊂ X,a bounded set, such that xn ∈ A for
all n.
Example 4.1.5. In (R, | · |),
(1) { n1 }∞
n=1 converges and is bounded.
(2) {n2 }∞ n=1 diverges and is unbounded.
(3) {1 + (−1)n }∞ n=1 diverges and is bounded.
4.3. Completeness
Unfortunately, the converse to Theorem 4.2.5 just does not hold. Not every
Cauchy sequence in a metric space must converge to a point in the space.
Example 4.3.1. Consider (0, 1) ⊂ (R, | · |). { n1 } is a Cauchy sequence in (0, 1),
but does not converge to a limit in (0, 1).
Example 4.3.2. Consider Q ⊂ (R, | · |). {(1 + n1 )n } is a Cauchy sequence in Q
because we know that (1 + n1 )n → e in (R, | · |), but its limit e 6∈ Q.
38 4. CAUCHY SEQUENCES IN METRIC SPACES
Since R is complete, {fn (x)} converges to a real number. Define the function
f : [a, b] → R by
f (x) = lim fn (x) a ≤ x ≤ b.
n→∞
This is our candidate for the limit.
4.3. COMPLETENESS 39
Since fN ∈ M([a, b]), there is an M such that sup |fN (x)| ≤ M and so
a≤x≤b
sup |f (x)| ≤ M + 1.
a≤x≤b
Finally, we show {fn } → f in (M([a, b]), d). Note that we have pointwise con-
vergence by construction, but we do not automatically have uniform convergence,
which is the convergence notion in (M([a, b]), d), automatically.
Let > 0. There is an N such that
d(fm , fN ) < for m ≥ N.
This means
|fm (x) − fN (x)| <
for a ≤ x ≤ b and m ≥ N . Taking the limit as m → ∞ yields
|f (x) − fN (x)| < , a ≤ x ≤ b.
For n ≥ N ,
|fn (x) − f (x)| ≤ |fn (x) − fN (x)| + |fN (x) − f (x)|
< 2,
for a ≤ x ≤ b. Hence, d(fn , f ) < 2 for n ≥ N .
We can characterize completeness in terms of a generalization of the Cantor
Intersection Property (recall Definition 3.1.23).
Definition 4.3.8. Let A be a nonempty subset of a metric space (χ, d). The
diameter of A is defined by
diam(A) = sup d(x, y).
x,y∈A
Notice that diam(A) can be infinite and the diameter of a set consisting of a
single point is zero.
We have
Theorem 4.3.9. A metric space is complete if and only if the intersection of
every descending sequence of nonempty, closed sets whose diameters approach zero
consists of a single point.
Proof. Let {Fn } be a sequence of closed sets in a metric space (χ, d) with
F1 ⊃ F2 ⊃ F3 ⊃ ... and diam(Fn ) → 0, where Fn 6= 0 for all n.
Choose xn ∈ Fn for n ≥ 1. Note that since the sets are descending,
sup d(x, y) ≥ sup d(x, y).
x,y,∈Fn x,y∈Fn+1
Given > 0, choose N such that diam(Fn ) < for n ≥ N . This implies that
d(xn , xm ) < for n, m ≥ N . Hence {xn } is a Cauchy sequence and there is a point
40 4. CAUCHY SEQUENCES IN METRIC SPACES
∞
\ ∞
\
x such that xn → x. We claim {x} = Fn . First note that if x, y ∈ Fn , then
n=1 n=1
x, y ∈ Fn for all n, and since d(x, y) ≤ diam(Fn ) for all n, d(x, y) = 0, i.e., x = y.
∞
\
Hence, Fn can consist of at most one point.
n=1
Now we claim x ∈ Fn for all n. If not, since Fnc is open for all n, there is an n
and an > 0 such that N (x) ⊂ Fnc . But, then d(x, y) ≥ for y ∈ Fn . But, this
means d(x, xm ) ≥ > 0 for m ≥ N , contradicting xn → x.
Now let {xn } be a Cauchy sequence in (χ, d). Choose N1 > 0 such that
n, m ≥ N1 implies d(xn , xm ) < 12 . Set xN1 as the first term in a subsequence. Given
N1 , N2 , ..., Nk−1 , choose Nk > Nk−1 such that m, n ≥ Nk implies d(xn , xm ) < 21k .
Let xNk be the k th term in the subsequence. Now
1
d(xNk , xNk+1 ) <
2k
since Nk+1 > Nk .
Define the sequence of closed “balls”
B1 = N 1 (xN1 )
B2 = N 12 (xN2 )
B3 = N 14 (xN3 )
..
.
Bk =N 1 (xNk )
2k−1
..
.
These closed sets are nonempty, since xNk ∈ Bk . Moreover, they are descending.
If y ∈ Bk+1 , then
d(y, xNk ) ≤ d(y, xNk+1 ) + d(xnk+1 , xNk )
1 1 1
≤ 2k
+ 2k
= 2k−1
,
T∞
so Bk+1 ⊂ Bk . By assumption, there is a unique point x ∈ k=1 Bk . Since
1
d(x, xNk ) ≤ 2k−1 , xNk → x. Since {xn } is Cauchy, xn → x.
We can characterize complete subsets of a complete metric space very nicely:
Theorem 4.3.10. Let (χ, d) be a complete metric space. A subspace Y ⊂ χ is
complete if and only if Y is closed.
Proof. Suppose Y is closed and {xn } is a Cauchy sequence in Y. Since χ is
complete, xk → x ∈ χ. But, Y is a closed, so x ∈ Y. This means Y is complete.
If Y is complete, then let x be a limit point of Y. There is a sequence {xn }
in Y with xn → x. This sequence converges in χ, so it is a Cauchy sequence in χ,
and therefore in Y. This means xn → x̃ ∈ Y and Y is closed.
We can restate
Theorem 4.4.3. Borel-Lebesguq Theorem II A subset of a metric space is
compact if and only if it is complete and totally bounded.
CHAPTER 5
Sequences in Rn
Theorem 5.1.1. Let {an } and {bn } be convergent sequences in (R, | · |) with
an → a and bn → b. Then,
(1) lim (an + bn ) = a + b.
n→∞
(2) lim can = ca and lim (c + an ) = c + a for any c ∈ R.
n→∞ n→∞
(3) lim an bn = ab.
n→∞
1
(4) limn→∞ an = a1 , provided a 6= 0 and an 6= 0 for all n.
such that n ≥ N1 implies |an − a| < min(|a|, 3B ) and n ≥ N2 implies
|bn − b| < min(|b|, 3B ). So, n ≥ max(N1 , N2 ) implies
1 1 |a − an | |a − an |
| − |= < |a|2 < .
an a aan
2
Proof. Let xm = (xm,1 , ..., xm,n ) and x = (y1 , ..., yn ). The claim is
(xm → x) ⇔ (xm,k → yk ), 1 ≤ k ≤ n.
|xm,k − yk | ≤k xm − x k, 1≤k≤n
Theorem 5.1.3. Suppose {xm } and {ym } are sequences in (Rn , k · k) and {am }
is a sequence in (R, | · |) such that xm → x, ym → y, am → a. Then
(1) lim (xm + ym ) = x + y.
m→∞
(2) lim xm · ym = x · y.
m→∞
(3) lim am xm = ax.
m→∞
(2) Let xm = (xm,1 , xm,2 , ..., xm,n ), ym = (ym,1 , ym,2 , ..., ym,n ), x = (c1 , c2 , ..., cn ), and y =
(d1 , d2 , ..., dn ) for all m ∈ N. Then
Xn
lim xm · ym = lim xm,k ym,k
m→∞ m→∞
k=1
n
X
= lim xm,k ym,k
m→∞
k=1
Xn
= cm dm by Theorem 5.1.1
m=1
= x · y.
(3) Using the notation above, notice am xm = (am xm,1 , ..., am xm,n ). By Theo-
rem 5.1.1, each of these components converges to ack , so by Theorem 5.1.2,
we have the desired result.
5.2. SEQUENCES IN R AND ORDER. 45
Example 5.2.6. Let {xn } = Q. Then lim sup xn = ∞ and lim inf xn = −∞.
n→∞ n→∞
46 5. SEQUENCES IN Rn
1 3 −4 5 −6 7
Example 5.2.7. {xn } = {(−1)n (1 + )} = {−2, , , , , , ...}.
n 2 3 4 5 6
Here, lim sup xn = 1 and lim inf xn = −1.
n→∞ n→∞
5.3. Series
Series are just a special kind of sequence.
Definition 5.3.1. Let {xn } be a sequence in (R| · |). The partial sums
associated to {xn } are defined
Xn
Sn = xm .
m=1
5.3. SERIES 47
The sequence of partial sums may or may not converge. If it does, we define
Definition 5.3.2. Let {xn } be a sequence in (R, | · |) and {Sn } the sequence
of partial sums. If {Sn } converges we define the series associated to {xn } as
X∞ Xn
xn = lim Sn = lim xm
n→∞ n→∞
n=1 m=1
and say the series converges. Otherwise, we say the series diverges.
The Cauchy Criterion for convergence becomes
∞
X
Theorem 5.3.3. A series xn converges if and only if for every > 0, there
n=1
is an N such that
m
X
| ak | < for m > n ≥ N.
k=n
We now study functions from one metric space to another. Recall Defini-
tion 1.2.1.
Example 6.0.14. Suppose f ∈ C([0, π]). Recall that we define its Fourier sine
series as
∞
X
ak sin(kx), 0 ≤ x ≤ π
k=1
with
Z π
2
ak = f (s) sin(ks)ds.
π 0
This gives you some idea of the tremendous range of applications of functions
on metric spaces.
48
6.1. LIMIT OF A FUNCTION 49
for every sequence {xn } in A such that xn 6= x for all n and lim xn = x. We say
n→∞
that f has a limit at x.
Note: The convergence lim xn = x is in the metric of X, i.e.,
n→∞
lim dx (xn , x) = 0, while the convergence lim f (xn ) = y is in
n→∞ n→∞
the metric of Y, i.e., lim dy (f (xn ), y) = 0.
n→∞
Example 6.1.2. Note that in the function in Figure 6.1, lim f (s) = y 6= f (x).
s→x
This is one reason for having the condition xn 6= x on the sequence.
f(x) f
x A
Figure 6.1
Proof. This follows from the properties of limits of sequences of real numbers.
In the same way, we prove
Theorem 6.5.2. Let (X, d) be a metric space and A ⊂ X.
(1) Suppose f1 , ..., fn : A → R with the usual metric and let f : A → Rn be
defined by f (x) = (f1 (x), ..., fn (x)), where we take the usual metric on Rn .
Then f is continuous at x ∈ A, or on A, if and only if each component
fm is continuous at x ∈ A, or on A.
(2) If f and g are continuous maps from A into (Rn , k · k), then f + g, f · g
are continuous.
Example 6.5.3. If x ∈ Rn is written x = (x1 , ..., xn ), then the coordinate
functions
φm (x) = xm , 1 ≤ m ≤ n,
are continuous since
|φm (x) − φm (y)| ≤k x − y k, x, y ∈ Rn .
Example 6.5.4. Repeated applications of Theorem 6.5.2 shows that polyno-
mials p : Rn → R,
M1 X
X M2 Mn
X
p(x) = ... Cm1 m2 ...mn xm mn
1 ...xn
1
m1 =0 m2 =0 mn =0
p(x)
are continuous. Furthermore, all rational functions q(x) , where p, q are polyno-
mials, are continuous at all points where q 6= 0.
The latter result also requires Theorem 6.2.7.
Example 6.5.5. Let (X, d) be a metric space and z ∈ X. We can define
f : X → R using d as
f (x) = d(x, z), x ∈ X.
The triangle inequality implies
d(x, z) ≤ d(x, y) + d(y, z)
d(y, z) ≤ d(x, y) + d(x, z), x, y ∈ X
or
|d(x, z) − d(y, z)| ≤ d(x, y), x, y ∈ X
and therefore f is continuous on X.
We now consider what happens for Rn -valued continuous functions on compact
sets.
Definition 6.5.6. Let (X, d) be a metric space and f : A → Rn , with the
usual metric on Rn . f is bounded on A if there is a constant M such that
k f (x) k≤ M for all x ∈ A.
If f is bounded on X, we say it is bounded.
Theorem 6.5.7. Let (X, d) be a metric space, A ⊂ X. If f : A → (Rn , k · k)
is continuous and A is compact, then f is bounded and f (A) is closed.
Proof. See Theorem 6.4.1 and Theorem 3.6.3.
56 6. CONTINUOUS FUNTIONS ON METRIC SPACES
2
0, n ≤x≤1
57
58 7. SEQUENCES OF FUNCTIONS AND C([a, b])
f1
f4
fn , n big
Figure 7.1
f4
f3
f2
f1
1
1 2
2 3 1
Figure 7.2
This example shows that pointwise convergence can allow some pretty bad
behavior!
In most situations, we want to know whether or not some particular properties
of a sequence of functions is inherited by the limit.
Example 7.1.4. When talking about C([a, b]), we want to know if the limit of
a sequence of continuous functions is continuous.
a x1 x2 x3 b
x0 xn
b−a
∆x = n
a xM b
x
Figure 7.3
Yn (x) is the area of the rectangles shown above. Yn (x) is continuous on [a, b]
for all n if f is continuous, and we want to know if {Yn (x)} → {y(x)}, the solution
of 7.1, which is also continuous if f is continuous. In most cases, we do not know
y.
Example 7.1.6. Continuing Ex. 7.1.4, if the sequence {fn } in C([a, b]) converges
to f , we can rephrase the issue of inheriting continuity like this: Choose a ≤ x ≤ b.
Then for n ≥ 1,
lim fn (xm ) = fn (x)
m→∞
for all sequences {xm } in [a, b] with xm → x. Since fn → f pointwise,
f (x) = lim fn (x) = lim lim fn (xm )
n→∞ n→∞ m→∞
for all sequences {xm } in [a, b] with xm → x. For f to be continuous at x, we must
have
lim f (xm ) = f (x)
m→∞
for all such sequences. In other words, we require
(7.2) lim lim fn (xm ) = lim lim fn (xm ).
m→∞ n→∞ n→∞ m→∞
Example 7.1.7. Continuing Ex. 7.1.5, it is easy to see that given x ∈ [a, b], for
all sufficiently fine meshes, i.e., sufficiently large n,
Yn (x + h) − Yn (x)
lim ≈ f (xm−1 )
h→0 h
and as n → ∞, xm−1 → x, so
Yn (x + h) − Yn (x)
lim lim = f (x).
n→∞ h→0 h
This says Yn (x) is an approximate solution.
On the other hand, if Yn (x) → y(x) and we want to show that y solves 7.1,
then we want for x ∈ [a, b],
y(x + h) − y(x)
lim = f (x)
h→0 h
or
Yn (x + h) − Yn (x) Yn (x + h) − Yn (x)
lim lim = lim lim
h→0 n→∞ h n→∞ n→0 h
General Principle: Whenever there is more than one limiting process in some
situation, it is important to determine if the order of the limit matters.
m
Example 7.1.8. Consider { }m,n=∞ .
n + m m,n=1
m
lim lim = lim 0 = 0
m→∞ n→∞ n + m m→∞
m
lim lim = lim 1 = 1.
n→∞ m→∞ n + m n→∞
In fact, the limit of a sequence of continuous functions that converge pointwise
is not necessarily continuous.
Example 7.1.9. Consider {xn }∞ n=0 on [0, 1]. The sequence is in C([0, 1]). It
converges pointwise on [0, 1] to
(
n 0, 0 ≤ x ≤ 1,
χ1 (x) = lim x =
n→∞ 1, x=1
which is not continuous (see figure 7.4).
If x < 1, then given > 0, choose N > − log(x) (recall log(x) < 0), so xn <
for n ≥ N . However, xn = 1 for x = 1 and all n.
7.2. UNIFORM CONVERGENCE: C([a, b]) IS CLOSED, AND COMPLETE 61
1 x0
x1
x2
x3
Figure 7.4
i.e.,
d(xn , x1 ) = 1 for all n.
Convergence in C([a, b]) is an example of uniform convergence.
Definition 7.2.2. Let (X, dx ), (Y, dy ) be metric spaces, A ⊂ X, and {fn } a
sequence of functions fn : A → Y for all n. {fn } converges uniformly to f on
A if for every > 0 there is an N such that
dy (fn (x), f (x)) < for x ∈ A and n ≥ N.
Compare this to Definition 7.1.1.
Example 7.2.3. The functions in Example 7.1.2 converge uniformly to 0 since
| n1 sin(mx) − 0| ≤ n1 for all 0 ≤ x ≤ π.
62 7. SEQUENCES OF FUNCTIONS AND C([a, b])
Example 7.2.5. The sequence {xn } does not converge uniformly to χ1 (x) on
[0, 1], but does on [0, 12 ].
Uniform convergence goes well with continuity.
Theorem 7.2.6. Let (X, dx ), (Y, dy ) be metric spaces and A ⊂ X. Suppose
{fn } is a sequence of functions with fn : A → Y continuous on A for all n and
fn → (f : A → Y) uniformly on A. Then f is continous of A.
Proof. Choose x ∈ A and > 0. We want to show we can make dy (f (y), f (x))
smaller than by making dx (x, y) small. Uniform convergence means that we can
make dy (f (x), fn (x)) and dy (f (y), fn (y)) small, so for y ∈ A, we write
dy (f (x), f (y)) ≤ dy (f (x), fn (x)) + dy (fn (x), fn (y)) + dy (fn (y), f (y)).
By uniform convergence, there is an N such that dy (f (x), fn (x)) < and
dy (f (y), fn (y)) < for n ≥ N , independent of x and y. Since fn is continuous on
A, there is a δ > 0 such that for any fixed n ≥ N ,
d(fn (x), fn (y)) < ,
for all y ∈ A, dx (x, y) < δ. Hence, using that value of n, we conclude
d(f (x), f (y)) < 3 for all y ∈ A, dx (x, y) < δ.
The functions in Example 7.1.3 show the converse does not hold: a sequence of
continuous functions can converge to a continuous function without the convergence
being uniform.
We now discuss the related topic of completeness. We state the Cauchy criterion
for uniform convergence.
Theorem 7.2.7. Let (X, dx ) and (Y, dy ) be metric spaces, Y complete, A ⊂ X,
and {fn } a sequence with fn : A → Y for all n. {fn } converges uniformly on A if
and only if for every > 0 there is an N such that
dy (fn (x), fm (x)) < for n, m ≥ N and x ∈ A.
Proof. Suppose {fn } converges uniformly on A to f . Given > 0, there is
an N such that
dy (fn (x), f (x)) < , x ∈ A, n ≥ N.
So,
dy (fn (x), fm (x)) ≤ dy (fn (x), f (x)) + dy (f (x), fm (x))
< 2
for x ∈ A, n, m ≥ N .
Conversely, suppose the Cauchy condition holds. The sequence of points {fn (x)}
is a Cauchy sequence in Y for each x ∈ A, and therefore has a limit in Y that we
call f (x). This defines f : A → Y. {fn } converges pointwise to f on A and we
have to show the convergence is uniform.
7.3. C([a, b]) IS SEPARABLE 63
Another way to state this result is that there is a sequence of polynomials {pn }
(of course in C([a, b])) that converges to f in C([a, b]), that is, uniformly.
This theorem is profoundly important. It is the reason, for example, that the
use of polynomials is so widespread in numerical analysis, i.e., approximation of
functions, integrals, solutions of differential equations, and so on.
Note: unlike Taylor’s polynomials, this result does not require
increasing smoothness of f to increase the accuracy of the poly-
nomial approximations.
To prove that C([a, b]) is separable, we first note that if
Xn X n
p(x) = am xm and p̃(x) = ãm xm
m=0 m=0
are two polynomials on [a, b], then
d(p, p̃) = sup |p(x) − p̃(x)|
a≤x≤b
n
X
≤c· |am − ãm |
m=0
≤ (n + 1) · c · max |am − ãm |
0≤m≤n
since the rationals are dense in R. (Note, however, that as the degree increases, the
coefficients generally must be approximated to increasing accuracy.) This means
that given a continuous function f on [a, b], and > 0, we can find a polynomial
with rational coefficients pn such that d(pn , f ) < . We first use Theorem 7.3.1 to
find a polynomial, with possibly real coefficients, that approximates f to within 2
and then construct a polynomial with rational coefficients that approximates the
first polynomial to within 2 .
Since the set of polynomials with rational coefficients is countable, this proves
Theorem 7.3.2. C([a, b]) is separable.
We first note that it suffices to prove Theorem 7.3.1 on [0, 1]. We can map [0, 1]
into [a, b] by y = (b − a)x + a and vice-versa by x = a−ya−b . If g is continuous on [a, b],
then f (x) = g((b − a)x + a) is continuous on [0, 1]. If pn approximates f to within
on [0, 1], then p̃n (y) = pn ( a−y
a−b ) is a polynomial that approximates g(y) to within
on [a, b].
We give a constructive proof that uses probability.
Definition 7.3.3. Recall for n ≥ m ≥ 0, the binomial coefficient n choose m,
n n!
=
m m!(n − m)!
Example 7.3.4.
4 4! 3 3!
= =6 = =1
2 2!2! 0 3!0!
n
m is the number of distince subsets with m objects that can be chosen from
a set of n objects. This is very important in probability.
Example 7.3.5. We compute the probability P of getting an ace of diamonds
in a poker hand of 5 cards chosen at random from a deck of 52 cards using
number of outcomes in the event
P(event) =
total number of possible outcomes
when all outcomes are equally likely.
The total number of 5 card poker hands is 52
5 . Obtaining a “good” hand
amounts to choosing any 4 cards
from the remaining 51 cards after getting an ace
of diamonds. So, there are 51
4 good hands.
51
4 5
P = 52 = .
5
52
It is straightforward to show that
n n n n n n
= , = = n, = = 1.
m n−m 1 n−1 n 0
There is also an important result called the binomial exansion.
Theorem 7.3.6. For n ∈ N and a, b ∈ R,
n
n
X n m n−m
(a + b) = a b .
m=0
m
7.3. C([a, b]) IS SEPARABLE 65
The binomial polynomials have several useful properties following from 7.4
and 7.5:
n
X
(7.6a) pn,m (x) = 1
m=0
n
X
(7.6b) mpn,m (x) = nx
m=0
n
X
(7.6c) m2 pn,m (x) = (n2 − n)x2 + nx
m=0
1
P √ (nlarge)
πn
Theorem 7.3.10. Law of Large Numbers Assume event E occurs with prob-
ability X and let m denote the number of times E occurs in n trials. Let > 0 and
δ > 0 be given. The probability that m
n differs from X by less than δ is greater than
1 − , i.e.,
m
P(| − x| < δ) > 1 − ,
n
for all n sufficiently large.
Note: This does not say that E occurs exactly Xn times, nor
that E must occur roughly Xn times.
Since lower bounds are difficult in general, we consider the complementary sum
giving the probability of what we don’t want:
X X
pn,m (X) = 1 − pn,m (x),
0≤m≤n 0≤m≤n
|m
n −X|≥δ |m
n −X|<δ
7.3. C([a, b]) IS SEPARABLE 67
that we estimate as
X 1 X m
pn,m ≤ 2 ( − X)2 pn,m (X)
δ n
0≤m≤n 0≤m≤n
|m
n −X|≥δ |m
n −X|≥δ
n
1 X
≤ (m − nX)2 pn,m (X)
n2 δ 2 m=0
n n n
1 X 2 X
2 2
X
≤ ( m p n,m (X) − 2nX mp n,m (X) + n X pn,m (X)).
n2 δ 2 m=0 m=0 m=0
Using 7.6a - 7.6c, the sums on the right simplify to nX(1−X). Since X(1−X) ≤
1
4 for 0 ≤ X ≤ 1,
X 1
(7.8) pn,m (x) ≤
4nδ 2
0≤m≤n
|m >δ
n −X|
and
X 1
pn,m (x) ≥ 1 − .
4nδ 2
0≤m≤n
|m
n −X|<δ
For given , δ > 0, we can insure (4nδ 2 )−1 < by choosing n > 1
4δ 2 .
Proof. Of Theorem 7.3.1. We first define the approximating polynomial,
named after the person who made this proof.
Definition 7.3.11. We partition [0, 1] by a uniform mesh with n + 1 nodes,
xm = m
n , m = 0, 1, ..., n. The Bernstein polynomial of order n for f on [0, 1] is
n
X
Bn (f, x) = Bn (x) = f (xm )pn,m (x).
m=0
2.75
2.5
2.25
2.0
1.75
1.5
1.25
1.0
which tends to zero like n1 on (0, 1]. This contrasts with interpolating polynomials
and Taylor polynomials, which both have the property that if f is a polynomial
then pn = f for n ≥ deg(f ).
Example 7.3.13. For ex on (0, 1] (see Figure 7.5),
B1 (x) = (1 − x) + ex
1
B2 (x) = (1 − x)2 + 2e 2 x(1 − x) + ex2
1 2
B3 (x) = (1 − x)3 + 3e 3 x(1 − x)2 + 3e 3 x2 (1 − x) + ex3
..
.
Theorem 6.4.2 implies f is uniformly continuous on [0, 1]. Given > 0, there
is a δ > 0 such that
|f (x) − f (xm )| <
2
for all x, xm in [0, 1] with |x − xm | ≤ δ. Given δ, by the way, we can find xm such
that |x − xm | ≤ δ for all sufficiently large n, since the rationals are dense in [0, 1].
Thus,
X X
| (f (x) − f (xm ))pn,m (x)| ≤ |f (x) − f (xm )|pn,m (x)
0≤m≤n 0≤m≤n
|x−xm |<δ |x−xm |<δ
X
≤ pn,m (x) = .
2 2
0≤m≤n
Now the second sum on the right in 7.9 is bounded after we realize that Theo-
rem 6.5.8 implies |f | is bounded on [0, 1] by some constant M . Hence, 7.8 implies
X X
| (f (x) − f (xm ))pn,m (x)| ≤ 2M pn,m (x)
0≤m≤n 0≤m≤n
|x−xm |≥δ |x−xm |≥δ
M
≤ .
2nδ 2
M
Given δ from the first estimate, we can force 2nδ 2 < 2 by taking n sufficiently
large.
K ⊂ C([a, b]) is bounded means there is a function g ∈ C([a, b]) and an M such
that
d(f, g) = sup |f (x) − g(x)| ≤ M
a≤x≤b
for all f ∈ K. Since such a g is itself bounded on [a, b], we see there is an M such
that
sup |f (x)| ≤ M for all f ∈ K.
a≤x≤b
This motivates
Definition 7.4.4. Let (X, d) be a metric space, A ⊂ X, and F a set of functions
from A into Rn with the usual metric. F is uniformly bounded on A if there is
an M such that
sup k f (x) k≤ M for all f ∈ F.
x∈A
Each of the fm is uniformly continuous on [a, b] and since there is a finite number
of {f1 , ..., fn }, there is a δ > 0 such that
|fm (x) − fm (y)| < for 1 ≤ m ≤ n, x, y ∈ [a, b], |x − y| < δ.
Choosing f ∈ K, we choose fm as above and write
|f (x) − f (y)| ≤ |f (x) − fm (x)| + |fm (x) − fm (y)| + |fm (y) − f (y)|
and with δ chosen as above,
|f (x) − f (y)| < 3,
for all x, y ∈ [a, b] with |x − y| < δ.
Since f was chosen arbitrarily, K is equicontinuous.
Now, we show that if K is uniformly bounded and equicontinuous, then it is
totally bounded. So, given any > 0, we construct a finite set of functions F such
that
[
K⊂ N (f ),
f ∈F
a x1 x2 x3 b
Y2M m = M
2M 1
2M ∆y 2M m
= m
< 5
Y0 = −M
Figure 7.6
Y2M m = M
Y0 = −M
x1 x2 x3 ... xn−1 xn
74 7. SEQUENCES OF FUNCTIONS AND C([a, b])
Let F be the set of continuous functions on [a, b] that are piecewise linear whose
“corner points” occur at points on the grid (see Figure 7.7).
< 5
Figure 7.8
We want to show g(x) is close to f (x) for all a ≤ x ≤ b (see Figure 7.8). Choose
a ≤ x ≤ b. Now xj ≤ x ≤ xj+1 for some 1 ≤ j ≤ n. By the equicontinuity of K
and choice of δ, we know |g(y) − g(xj )| < 5 for xj ≤ y ≤ xj+1 . It follows that
|f (xj+1 ) − f (xj )| ≤ |f (xj+1 ) − g(xj+1 )| + |g(xj+1 ) − g(xj )| + |g(xj ) − f (xj )|
3
< + + = .
5 5 5 5
Since f is linear on [xj , xj+1 ],
3
|f (y) − g(xj )| < , xj ≤ y ≤ xj+1 .
5
7.4. COMPACT SETS IN C([a, b]) 75
So,
|g(x) − f (x)| ≤ |g(x) − g(xj )| + |g(xj ) − f (xj )| + |f (xj ) − f (x)|
< .
F is an -net for K.
Example 7.4.14. We don’t have time for details, but a classic application of
the Arzela-Ascoli Theorem is to show that the forward Euler approximation
Y0 = y0
Yn = Yn−1 + ∆tf (Yn−1 ), n = 1, 2, ..., N,
T
where ∆t = N, for the initial value problem
(
y 0 = f (y), 0 ≤ t ≤ T
y(0) = y0
converges to y for 0 ≤ t ≤ T if f is continuous.