0% found this document useful (0 votes)
13 views27 pages

sequence

Uploaded by

tiutiucap
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
13 views27 pages

sequence

Uploaded by

tiutiucap
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 27

MATH 3033 Real Analysis

Sequences and Series of Functions


Dr. Albert Ku

1 Convergence of Functions
Definition 1.1. Let U ⊆ Rm and fn : U → R for n ∈ N. We say the sequence of functions {fn } converges
pointwise on U to the limit function f : U → R if for each x ∈ U , {fn (x)} converges and

f (x) = lim fn (x)


n→∞

Equivalently, for any ε > 0 and x ∈ U , there is an integer N , which may depend on x as well as ε, such that

|fn (x) − f (x)| < ε if n > N

Definition 1.2. Let U ⊆ Rm and fn : U → R for n ∈ N. We say the sequence of functions {fn } converges
uniformly on U to the limit function f : U → R if for any ε > 0, there is an integer N such that

|fn (x) − f (x)| < ε for all x ∈ U if n > N

In fact, the above condition can also be stated as follows:

∥fn − f ∥U = sup {|fn (x) − f (x)| | x ∈ U } < ε if n > N

(In general, for any g : U → R, ∥g∥U = sup {|g(x)| | x ∈ U } is called the sup-norm of g on U .)

Remarks 1.1.

1. If {fn } convereges uniformly on U to f , then it converges pointwise on U to f . However, the converse


is not true.
2. The definition of pointwise convergence of functions implies for each x0 , fn (x0 ) approximates f (x0 )
for sufficiently large n. It does not imply that any particular fn is a good approximation of f over the
entire domain.

1
Example 1.1.
x
1. Let fn (x) = for x ∈ R. Then {fn } converges uniformly to the limit f = 0 on R because for
1 + n2 x2
any x ∈ R,
|x| 1 2n|x| 1
|fn (x) − f (x)| = = · ≤ .
1 + n2 x2 2n 1 + n2 x2 2n
In this example, {fn } is a good approximation of f for sufficiently large n.

nx
2. Let gn (x) = and g(x) = 0 for x ∈ R. Then for any x ̸= 0,
1 + n2 x2
n|x| n|x| 1
|gn (x) − g(x)| = 2 2
≤ 2 2 = → 0 as n → ∞,
1+n x n x n|x|

and since gn (0) = 0 for all n, gn(x) converges to the limit g(x) = 0 for all x ∈ R.
However, for any n ∈ N, gn n1 = 21 . Thus {gn } does not converge uniformly to g on R i.e. gn does
not approximate g over the entire real line.

More generally, we have the following test for uniform convergence of sequence of functions:
Theorem 1.1. (L-Test) Let {fn } be a sequence of functions on U ⊆ Rm such that it converges pointwise
to f on U . If there exists a sequence of real numbers {Ln } such that for any n ∈ N,

|fn (x) − f (x)| ≤ Ln for all x ∈ U

and lim Ln = 0, then {fn } converges uniformly to f on U .


n→∞

2
We will see that many nice properties of functions can be preserved in the limit of a uniformly convergent
sequence of functions. The following theorem shows that continuity is one such property:
Theorem 1.2. If {fn } converges uniformly to f on U and fn is continuous on U for all n ∈ N, f is continuous
on U .

Proof. Take any x0 ∈ U . For any n ∈ N and x ∈ U ,

|f (x) − f (x0 )| ≤ |f (x) − fn (x)| + |fn (x) − fn (x0 )| + |fn (x0 ) − f (x0 )|
ε
Suppose ε > 0. Since {fn } converges uniformly to f on U , we can choose n so that |fn (x) − f (x)| < 3 for
all x ∈ U . Therefore, we have

|f (x) − f (x0 )| < |fn (x) − fn (x0 )| + , x∈U
3
ε
Since fn is continuous at x0 , there exist δ > 0 such that |fn (x) − fn (x0 )| < 3 whenever |x − x0 | < δ. Apply
this to the above, we have
ε 2ε
|f (x) − f (x0 )| < + = ε whenever |x − x0 | < δ
3 3
Therefore, f is continuous at x0 .

Remarks 1.2. In the above theorem, the uniform convergence cannot be replaced by pointwise convergence.
The following exercise is an example of a sequence of continuous functions that converges pointwise to a
non-continuous function.

Exercise 1.1. For any n ∈ N, define fn : [0, 1] → R by fn (x) = xn for x ∈ [0, 1]. Show that {fn } converges
pointwise to the following function f on [0, 1]:
(
0 when 0 ≤ x < 1
f (x) =
1 when x = 1

(By the above theorem, {fn } does not converge uniformly on [0, 1].)

3
Theorem 1.3. (Cauchy critierion) The sequence of functions {fn }, defined on U , converges uniformly
on U if and only if for every ε > 0, there exists an integer N such that n, m ≥ N , x ∈ U implies

|fn (x) − fm (x)| < ε

Proof.
(⇒) Suppose {fn } converges uniformly on U , and let f be the limit function. Then there exists N ∈ N such
that n ≥ N , x ∈ U implies
ε
|fn (x) − f (x)| <
2
Then we have
ε ε
|fn (x) − fm (x)| ≤ |fn (x) − f (x)| + |f (x) − fm (x)| < + = ε
2 2
if n, m ≥ N and x ∈ U .

(⇐) Suppose the Cauchy criterion holds. Then for any x ∈ U , the sequence {fn (x)} converges, to a limit
which we may call f (x). Thus {fn } converges pointwise to f on U . We have to prove that the convergence
is uniform.
Let ε > 0, and choose N such that |fn (x)−fm (x)| < 2ε for all n, m ≥ N and x ∈ U . Also, since fn (x) → f (x)
as n → ∞, there exists a sufficiently large m (which might depend on x) such that m > N and
ε
|fm (x) − f (x)| <
2
Then for any n ≥ N and x ∈ U , we have
ε ε
|fn (x) − f (x)| ≤ |fn (x) − fm (x)| + |fm (x) − f (x)| < + =ε
2 2
Consequently, {fn } converges to f uniformly on U .

4

X
n
Definition 1.3. If {fn } is a sequence of functions defined on U ⊆ R , then fn is said to be an infinite
n=0

X
series of functions on U . The partial sums of fn are defined by
n=0

n
X
Fn = fk
k=0


X
If {Fn } converges pointwise to a function F on U , we say that fn converges pointwise to the sum F
n=0
on U , and write

X
F (x) = fn (x)
n=0

X
If {Fn } converges uniformly to a function F on U , we say that fn converges uniformly to the sum F
n=0
on U .

X
Example 1.2. The functions fk (x) = xk , k = 0, 1, 2, . . . define the infinite series xk on R. Its nth partial
k=0
sum is
 1 − xn+1

2 n , x ̸= 1
Fn (x) = 1 + x + x + · · · + x = 1−x
n + 1, x=1

It is easy to see that {Fn } converges pointwise to


1
F (x) = , if |x| < 1
1−x
Hence, we write

X 1
xk = , −1 < x < 1
1−x
k=0

For any fixed n, since the difference


xn+1
F (x) − Fn (x) =
1−x
can be made arbitrarily large by taking x close to 1, and so the convergence is not uniform on (−1, 1).
However, the series does converge uniformly on any interval [−r, r] with 0 < r < 1, since for any x ∈ [−r, r],

rn+1
|Fn (x) − F (x)| ≤ → 0 as n → ∞.
1−r

5
The following are two tests for uniform convergence of series:

X
Theorem 1.4. (Cauchy criterion for uniform convergence of series) A series fn converges uni-
n=0
formly on U if and only if for each ε > 0, there exists N ∈ N such that for any x ∈ U and m ≥ n > N ,

|fn (x) + fn+1 (x) + · · · + fm (x)| < ε

Proof. Follows from applying Cauchy criterion on the sequence of partial sum {Fn }.

Theorem 1.5. (Weierstrass M-test) Suppose |fn (x)| ≤ Mn for all x ∈ U i.e. fn is bounded by Mn .

X ∞
X
Then, if Mn converges, fn converges uniformly on U .
n=0 n=0


X
Proof. Since the partial sum of Mn converges, it is a Cauchy sequence, which means that for any ε > 0,
n=0
there exists N ∈ N such that if m ≥ n > N ,

Mn + Mn+1 + · · · + Mm < ε,

which implies that for any x ∈ U ,

|fn (x) + fn+1 (x) + · · · + fm (x)| ≤ |fn (x)| + |fn+1 (x)| + · · · + |fm (x)| ≤ Mn + Mn+1 + · · · + Mm < ε

X
By Theorem 1.4, fn converges uniformly on U .
n=0

∞ ∞
X 1 X sin(nx) 1
Example 1.3. The series 2 + n2
and 2
converges uniformly on R by considering Mn = 2 .
n=0
x n=1
n n

6

X ∞
X
Definition 1.4. fn is said to converge absolutely pointwise/uniformly on U if |fn | converges
n=0 n=0
pointwise/uniformly on U .


X
Exercise 1.2. If a series fn satisfies the Weierstrass M-test, then it converges absolutely uniformly.
n=0


X 1
Exercise 1.3. Consider the series x
.
n=1
n

(a) Given any p > 1, prove that the series converges uniformly on [p, ∞).

X 1
(b) Define ζ(x) = for any x > 1. Prove that ζ(x) is continuous on (1, ∞).
n=1
nx

7
2 Power Series
In this section, we will study a very special kind of series of functions:

X
Definition 2.1. Let a ∈ R and {an } be a sequence of real numbers. Then the series an (x − a)n is called
n=0
a power series. a is called the center of the power series.

A very natural question about power series is to determine the set of values of x on which a given power
series converges pointwise (or uniformly). To answer this question, we first need the following definitions:

Definition 2.2. Given {an } a sequence of real numbers. The limit superior of {an } is the limit of
Mn = sup{am | m ≥ n} as n → ∞:

lim sup an = lim sup{am | m ≥ n} = lim Mn


n→∞ n→∞ n→∞

Similarly, the limit inferior of {an } is the limit of mn = inf{am | m ≥ n} as n → ∞:

lim inf an = lim inf{am | m ≥ n} = lim mn


n→∞ n→∞ n→∞

It is always true that lim inf an ≤ lim sup an .


n→∞ n→∞

Remarks 2.1.
1. If {an } is unbounded above, then Mk = ∞ for all k ∈ N and lim sup an = lim Mk = ∞.
n→∞ k→∞

2. If {an } is unbounded below, then mk = −∞ for all k ∈ N and lim inf an = lim mk = −∞.
n→∞ k→∞

3. Suppose {an } is bounded. By definition, {Mn } is a decreasing sequence and is bounded below. There-
fore, lim sup an = lim Mn exists. Similarly, {mn } is an increasing sequence and is bounded above.
n→∞ n→∞
Therefore, lim inf an = lim mn exists.
n→∞ n→∞

4. For any sequence {an }, we have mn ≤ an ≤ Mn for all n ∈ N. If lim sup an = lim inf an = L, then
n→∞ h→∞
lim an = L by sandwich theorem.
n→∞

8
Theorem 2.1. Let H = lim sup an and h = lim inf an for a bounded sequence {an }. For any ε > 0,
n→∞ n→∞

(a) there are infinitely many n ∈ N such that an ∈ (H − ε, H + ε), and there are at most finitely many
n ∈ N such that an > H + ε;
(b) there are infinitely many n ∈ N such that an ∈ (h − ε, h + ε), and there are at most finitely many n ∈ N
such that an < h − ε.
Proof. To prove part (a), we first prove that there are infinitely many n ∈ N such that an > H − ε for
any ε > 0. Suppose not, then there is an ε0 > 0 such that there are only finitely many n ∈ N such
that an > H − ε0 . This implies that there exists N ∈ N such that an ≤ H − ε0 for all n > N . Hence
Mn = sup{am | m ≥ n} ≤ H − ε0 for all n > N . Thus, H = lim Mn = H ≤ H − ε0 , a contradiction.
n→∞

We next prove that there are only finitely many n ∈ N such that an > H + ε for any ε > 0. For any given
ε > 0, since lim Mn = H, there exists N ∈ N such that Mn < H + ε for n > N . Since an ≤ Mn for any n,
n→∞
an < H + ε for n > N .

The proof of part (b) is similar.

Corollary 2.1. Let H = lim sup an and h = lim inf an for a bounded sequence {an }.
n→∞ n→∞

(a) There is a subsequence converging to h and a subsequence converging to H.


(b) If L = {l ∈ R| l is a limit of some convergent subsequence of {an }}, then h and H are respectively the
smallest and the largest element of L.
Proof.

(a) Since for each k ∈ N there are infinitely many n ∈ N such that an ∈ H − k1 , H + k1 , we can pick


successively for each k ∈ N an nk such that ank ∈ H − k1 , H + k1 and nk > nk−1 . The subsequence
{ank } clearly converges to H.
Similarly, we can construct a subsequence converging to h.
(b) Let l be the limit of some subsequence {ank }. It suffices to prove that h ≤ l ≤ H. For any ε > 0,
there exists K ∈ N such that ank > l − ε for all k > K. But there are only finitely many nk such that
ank > H + ε. Hence l − ε ≤ H + ε. This implies l ≤ H + 2ε for any ε > 0, i.e. l ≤ H.
Similarly, we can prove that h ≤ l.

9
Remarks 2.2.
1. Bolzano-Weierstrass Theorem (on R) follows immediately from part (a) of the above corollary.
2. By part (b), we have lim sup an = sup L and lim inf an = inf L. Therefore, we have the following nice
n→∞ n→∞
properties:
ˆ For any c ∈ R such that c > 0, we have

lim sup(can ) = sup(cL) = c sup L = c lim sup an


n→∞ n→∞

lim inf (can ) = inf(cL) = c inf L = c lim inf an


n→∞ n→∞

ˆ More generally, if {cn } is a sequence such that lim cn = c > 0, then we have
n→∞

lim sup(cn an ) = c lim sup an


n→∞ n→∞

lim inf (cn an ) = c lim inf an


n→∞ n→∞

10
This is the main result we will use in this section is as follows:
Theorem 2.2. (Strong Form of Root Test) Suppose {an } is a sequence in R.

1
X
1. lim sup |an | n < 1 ⇒ an converges absolutely.
n→∞
n=1


1
X
2. lim sup |an | n > 1 ⇒ an diverges.
n→∞
n=1
1
Proof. To prove the strong form of root test, we let M = lim sup |an | n . If M < 1, then choose r such that
n n→∞ o
1
M < r < 1. Since M = lim Mk , where Mk = sup |an | | n ≥ k , there exists K ∈ N such that MK < r.
n
k→∞
1
This implies that |an | n < r, or equivalently, |an | < rn , for any n ≥ K. Hence, we have

X ∞
X
|an | ≤ rn < ∞
n=K n=K


X
which means that |an | converges.
n=1
1
If M > 1, then there exists a subsequence of |ank | nk that converges to M , which means that there exist
1
K ∈ N such that for all k > K, |ank | nk > 1. This implies that |ank | > 1 for all k > K and hence lim an ̸= 0.
n→∞

X
By the test for divergence, an diverges.
n=1

Remarks 2.3. Sometimes it is more convenient to use ratio test. In fact, if an ̸= 0 for all n ∈ N, we have
the following inequalites:

an+1 1 1 an+1
lim inf ≤ lim inf |an | n ≤ lim sup |an | n ≤ lim sup
n→∞ an n→∞ n→∞ n→∞ an

The above inequalities imply the following strong form of ratio test: If {an } is a sequence such that
an ̸= 0 for all n ∈ N. Then

an+1 X
1. lim sup <1⇒ an converges absolutely.
n→∞ an n=1


an+1 X
2. lim inf >1⇒ an diverges.
n→∞ an n=1

an+1 1
Moreover, if lim exists, then it is equal to lim |an | n . (Proof omitted)
n→∞ an n→∞

11
The following is the main theorem about the convergence of power series:

X
Theorem 2.3. For any power series an (x − a)n , there exists ρ such that 0 ≤ ρ ≤ +∞ and
n=0

(a) if 0 < r < ρ, the series converges absolutely uniformly on [a − r, a + r],

(b) if |x − a| > ρ, the series diverges at x.


Proof. Let ρ be defined by
1 1 1 1
ˆ Case 1: = lim sup |an | n if {|an | n } is bounded and lim sup |an | n ̸= 0,
ρ n→∞ n→∞
1 1
ˆ Case 2: ρ = ∞ if {|an | n } is bounded and lim sup |an | n = 0, and
n→∞
1
ˆ Case 3: ρ = 0 if {|an | } is not bounded above.
n

Obviously, we have
|an (x − a)n | ≤ |an |rn
for |x − a| ≤ r. Therefore, by Weierstrass M-test, the power series converges absolutely uniformly on

X
[a − r, a + r] if |an |rn converges.
n=0

1
X
By the strong form of root test, |an |rn converges if lim sup(|an |rn ) n < 1, which is equivalent to
n→∞
n=0

1
r lim sup |an | n < 1
n→∞

It clearly holds if r < ρ for Case 1 and 2. This proves part (a) of the theorem.

1 |x − a|
For Case 1 and 3, if |x − a| > ρ, then |x − a| lim sup |an | n = > 1. Then we have
n→∞ ρ
1 1
lim sup |an (x − a)n | n = |x − a| lim sup |an | n > 1
n→∞ n→∞

By the strong form of root test, the power series diverges.

12
Definition 2.3. The value of ρ (necessarily unique) in Theorem 2.3 is called the radius of convergence

X
of power series an (x − a)n .
n=0

Remarks 2.4.
1. No general statement can be made concerning convergence of a power series at x = a ± ρ i.e. some
series may converge at x = a ± ρ while some other series may diverge.
2. In general, a power series does not converge uniformly on (a − ρ, a + ρ). For example, ρ = 1 for the
P∞ 1
series n=0 xn and the series converges to for |x| < 1, diverges for |x| ≥ 1, and the series does
1−x
not converge uniformly on (−1, 1).


X
Theorem 2.4. If f (x) = an (x − a)n has radius of convergence ρ > 0, then f is differentiable for
n=0
|x − a| < ρ with

X
f ′ (x) = nan (x − a)n−1 .
n=1

The power series on the right also has radius of convergence ρ.



X
Proof. Here we only prove that the power series nan (x − a)n−1 has radius of convergence ρ. The proof
n=1
of the convergence to f ′ (x) will be deferred to the next section.


X
Let ρ1 be the radius of convergence of the power series nan (x − a)n−1 . Then we use the argument similar
n=1
to the proof of Therorem 2.3. Let r > 0. On [a − r, a + r], we have

|nan (x − a)n−1 | ≤ |nan |rn−1



X
Then by Weierstrass M-test, the power series converges uniformly on [a − r, a + r] if |nan |rn−1 converges.
n=1

1
X
By the strong form of root test, |nan |rn−1 converges if lim sup(|nan |rn−1 ) n < 1, which is equivalent to
n→∞
n=1

n−1 1 1 1
lim sup(r n n n |an | n ) = r lim sup |an | n < 1
n→∞ n→∞

Therefore, ρ1 = ρ.

13

X
Corollary 2.2. If f (x) = an (x − a)n has radius of convergence ρ > 0, then f is infinitely differentiable
n=0
for |x − a| < ρ with

X
f (k) (x) = n(n − 1) · · · (n − k + 1)an (x − a)n−k .
n=k
(k)
for any k ∈ N. In particular, f (a) = k!ak .
Proof. Apply Theorem 2.4 k times.


X
Given a power series an (x − a)n with radius of convergence ρ > 0, we can define f (x) to be the power
n=0
series and by Corollary 2.2, f is infinitely differentiable on (a − ρ, a + ρ) with f (k) (a) = k!ak . It is natural to
consider the following question: Suppose we are given a function f that is infinitely differentiable on an open
interval containing a, can it be expressed as a power series centered at a? This is the motivation behind the
following definition:
Definition 2.4. Let a ∈ R, ρ > 0 and f : (a − ρ, a + ρ) → R be a function that is infinitely differentiable.
The Taylor series of f centered at a is

X f (n) (a)
(x − a)n
n=0
n!

The partial sum of the Taylor series up to the k th term is called the Taylor polynomial of f of degree k
centered at a:
k
X f (n) (a)
(x − a)n
n=0
n!

14
It is easy to see that if an infinitely differentiable function f can be expressed as a power series, then the
power series must be the Taylor series of f . However, the following example shows that there exists an
infinite differentiable function that is not equal to its Taylor series.
Example 2.1. Let f : R → R be defined by
( 2
e−1/x when x ̸= 0
f (x) =
0 when x = 0

It can be shown that for any k ∈ N, f (k) (0) = 0. Therefore, the Taylor series of f centered at 0 is 0, which
is certainly not equal to f .

If an infinitely differentiable function equals its Taylor series, it is called an analytic function.

The following theorem expresses the error term explicitly when a function is approximated by its Taylor
polynomial:
Theorem 2.5. Let f : (a − ρ, a + ρ) → R be a function such that its (k + 1)th derivative is continuous at
a, where k is a non-negative integer. Then for any x ∈ (a − ρ, a + ρ),
k
X f (n) (a)
f (x) = (x − a)n + Rk+1 (x)
n=0
n!

where
f (k+1) (c)
(Mean-value form) Rk+1 (x) = (x − a)k+1 , for some c between a and x
(k + 1)!
1 x (k+1)
Z
(Integral form) Rk+1 (x) = f (t)(x − t)k dt
k! a
Proof. The mean-value form was already proved in MATH 2033. Here we will only prove the integral form.
For each x, we apply integration by parts repeatedly to obtain the following:
Z x
f (x) = f (a) + f ′ (t)dt
a
x
Z x

= f (a) + f (t)(t − x) + f ′′ (t)(x − t)dt
a a
Z x

= f (a) + f (a)(x − a) + f ′′ (t)(x − t)dt
a
Z x
′ ′′ (x − t)2 x (x − t)2
= f (a) + f (a)(x − a) − f (t) + f ′′′ (t) dt
2! a a 2!
x
f ′′ (a) (x − t)2
Z
= f (a) + f ′ (a)(x − a) + (x − a)2 + f ′′′ (t) dt
2! a 2!
···
k
f (n) (a) 1 x (k+1)
X Z
= (x − a)k + f (t)(x − t)k dt.
n=0
n! k! a

15
The following corollary gives a sufficient condition for an infinitely differentiable function to be analytic:
Corollary 2.3. Let f : (a − ρ, a + ρ) → R be an infinitely differentiable function. Moreover, if there exists
r ∈ (0, ρ) such that
ˆ for any n ∈ N, there exists Mn > 0 such that |f (n) (x)| ≤ Mn for all x ∈ [a − r, a + r], and
Mn r n
ˆ → 0 as n → ∞,
n!
then the Taylor series of f centered at a has radius of convergence at least r and converges uniformly to f
on [a − r, a + r].
Proof. Let fk (x) be the k th Taylor polynomial of f centered at a. Then by Theorem 2.5, for x ∈ [a − r, a + r],

f (k+1) (c)
|fk (x) − f (x)| ≤ (x − a)k+1 for some c between x and a
(k + 1)!
Mk+1 rk+1

(k + 1)!

Mk+1 rk+1
By assumption, → 0 as n → ∞. Hence, the Taylor series converges uniformly to f on [a−r, a+r].
(k + 1)!
Since the Taylor series is a power series, its radius of convergence must be at least r.

Example 2.2. Special functions like ex , sin x, cos x are analytic functions i.e. they are equal to their own
Taylor series because they satisfy the conditions in Corollary 2.3.

X xn
ex =
n=0
n!

X (−1)n x2n+1
sin x =
n=0
(2n + 1)!

X (−1)n x2n
cos x =
n=0
(2n)!

All the above Taylor series have radius of convergence ∞.

16
Another important result is the generalization of Binomial Theorem:
Theorem 2.6. (Binomial Theorem) For any a ∈ R, x ∈ (−1, 1),

a(a − 1) 2 X a(a − 1) · · · (a − k + 1)
(1 + x)a = 1 + ax + x + ··· = 1 + xk
2 k!
k=1

Proof. Let f (x) = (1 + x)a . Then f (k+1) (x) = a(a − 1) · · · (a − k)(1 + x)a−k−1 . By Theorem 2.5, we have

1 x (k+1)
Z
|Rk+1 (x)| = f (t)(x − t)k dt
k! 0
k
a(a − 1) · · · (a − k) x x − t
Z 
= (1 + t)a−1 dt
k! 0 1 + t

x−t
Claim: ≤ |x| for any t in the closed interval with 0 and x as endpoints.
1+t
x−t −1 − x
Let g(t) = . g(0) = x and g ′ (t) = < 0. Hence g is decreasing and we have
1+t (1 + t)2

g(t) ≤ g(0) = x if x > 0

g(t) ≥ g(0) = x if x < 0


x−t
Therefore, we get ≤ |x| for any t in the closed interval with 0 and x as endpoints and we have
1+t
Z x
a(a − 1) · · · (a − k) k
|Rk+1 (x)| ≤ |x| (1 + t)a−1 dt = bk+1
k! 0

In order to show that lim |Rk+1 | = 0, we consider the following:


k→∞


bk+1 |a − k| X
lim = lim |x| = |x| < 1 ⇒ bk converges by ratio test
k→∞ bk k→∞ k
k=1

⇒ lim bk = 0 by term test ⇒ lim |Rk+1 (x)| = 0.


k→∞ k→∞

17
3 Uniform Convergence and Differentiation
Suppose {fn } is a sequence of differentiable functions such that it converges uniformly f . Then it is natural
to ask the following questions:
ˆ Is f differentiable?

ˆ If f is differentiable, will {fn′ } converge uniformly to f ′ ?

Let’s consider the following examples.


Example 3.1. For any n ∈ N, we define
sin(nx)
fn (x) = √
n
Then it can easily be shown that {fn } converges uniformly on R to f = 0. However, we have

fn′ (x) = n cos(nx)

so that {fn′ } does not converge pointwise to f ′ = 0. For instance, fn′ (0) = n → ∞ as n → ∞. But
f ′ (0) = 0.

Example 3.2. For any n ∈ N, we define


1 1


 −x − if − 1 ≤ x ≤ −

 2n n
n 2 1 1

fn (x) = x if − < x <
 2 n n
 x− 1 1


if ≤x≤1

2n n
{fn } is a sequence of differentiable functions on [−1, 1] and it can be shown that {fn } converges uniformly
on [−1, 1] to f (x) = |x|. But clearly f (x) = |x| is not differentiable at x = 0.

18
Therefore, uniform convergence of differentiable functions does not guarantee the differentiability of the limit
function. We need much stronger conditions. The following is the main theorem of this section:
Theorem 3.1. Suppose {fn } is a sequence of functions, differentiable on [a, b] and such that {fn (x0 )}
converges for some point x0 ∈ [a, b]. If {fn′ } converges uniformly on [a, b], then {fn } converges uniformly on
[a, b] to a function f , and for any x ∈ [a, b],

f ′ (x) = lim fn′ (x)


n→∞

Proof. First, we need to prove that {fn } converges uniformly on [a, b]. Let ε > 0. Since {fn (x0 )} converges
and {fn′ } converges uniformly on [a, b], there exists N ∈ N such that for m, n ≥ N ,
ε
|fn (x0 ) − fm (x0 )| <
2
and
ε
|fn′ (t) − fm

(t)| < (1)
2(b − a)
for any t ∈ [a, b]. Using mean value inequality on the function fn − fm and the above inequality, we have
ε ε
|fn (x) − fm (x) − (fn (t) − fm (t))| ≤ |x − t| ≤ (2)
2(b − a) 2

for any x, t ∈ [a, b]. Combining the inequalities, for any x ∈ [a, b],

|fn (x) − fm (x)| ≤ |fn (x) − fm (x) − (fn (x0 ) − fm (x0 ))| + |fn (x0 ) − fm (x0 )|
ε ε
< + =ε
2 2
By Cauchy criterion, {fn } converges uniformly on [a, b].

Now, we define f (x) = lim fn (x) for any x ∈ [a, b] i.e. {fn } converges uniformly to f on [a, b]. We need to
n→∞
show that for any x ∈ [a, b],
f ′ (x) = lim fn′ (x)
n→∞

Fix any x ∈ [a, b]. Define 


 fn (t) − fn (x)
if t ̸= x
ϕn (t) = t−x
f ′ (x) if t = x
n

By definition, ϕn (t) is continuous on [a, b]. Then by the inequalities (1) and (2) again, for t ∈ [a, b],
ε
|ϕn (t) − ϕm (t)| ≤
2(b − a)

Therefore, {ϕn } converges uniformly on [a, b] and moreover, the limit function ϕ is continuous on [a, b] by
f (t) − f (x)
Theorem 1.2. Obviously, for t ̸= x, {ϕn (t)} converges pointwise to ϕ(t) = . By the continuity of
t−x
the limit function, {ϕn (x)} converges to ϕ(x) = lim ϕ(t). Hence, we have
t→x

f (t) − f (x)
f ′ (x) = lim = ϕ(x) = lim ϕn (x) = lim fn′ (x)
t→x t−x n→∞ n→∞

19
Remarks 3.1. We are now ready to complete the proof of Theorem 2.4: Consider the power series f (x) =

X n
X
ak (x − a)k with radius of convergence ρ > 0. Let fn (x) = ak (x − a)k . Obviously, for 0 < r < ρ, fn is
k=0 k=0
differentiable on [a − r, a + r] and we already proved that

X
kak (x − a)k−1
k=1

is also a power series with radius of convergence ρ, which implies that {fn′ } converges uniformly on [a−r, a+r].
By Theorem 3.1,
X∞

f (x) = kak (x − a)k−1
k=1

for any x ∈ [a − r, a + r]. Since r can be freely chosen between 0 and ρ, the above equality holds for any
x ∈ (a − ρ, a + ρ).

20
4 Uniform Convergence and Integration
In this section, we first define rigorously the definite integral of a real-valued function that we learned before.
Definition 4.1. Let f : [a, b] → R be a bounded real-valued function. A partition P of [a, b] is a finite set
of points t0 , t1 , . . . , tn such that
a = t0 < t1 < · · · < tn = b
For j = 1, 2, . . . , n, define
Mj = sup f ([tj−1 , tj ]), mj = inf f ([tj−1 , tj ]),
and define
n
X n
X
L(f, P ) = mj (tj − tj−1 ) U (f, P ) = Mj (tj − tj−1 )
j=1 j=1

Then L(f, P ) is called the lower Riemann sum for f corresponding to P and U (f, P ) is the upper Rie-
mann sum for f corresponding to P .

Clearly, we have U (f, P ) ≥ L(f, P ).


If we write L(f ) = sup{L(f, P )| P is a partition}, U (f ) = inf{U (f, P )| P is a partition}, then it can be
shown that for any partition P and P ′ of [a, b],

L(f, P ) ≤ L(f ) ≤ U (f ) ≤ U (f, P ′ ).

Definition 4.2. L(f ) is called the lower Riemann integral of f on [a, b]; U (f ) is called the upper
Riemann integral of f on [a, b]. If L(f ) = U (f ), then f is said to be Riemman integrable on [a, b], and
Z b Z b
the common value of L(f ) and U (f ) is written f (x) dx or f , and is called the Riemann integral of
a a
f on [a, b].
Remarks 4.1. If f is bounded on [a, b], then there exist m and M such that m ≤ f (x) ≤ M for all x ∈ [a, b].
It follows easily that
m(b − a) ≤ L(f ) ≤ U (f ) ≤ M (b − a).
If f is Riemann integrable, then it follows that
Z b
m(b − a) ≤ f ≤ M (b − a).
a

21
The following theorem gives a condition for the Riemann integrability of a function:
Theorem 4.1. If f is bounded on [a, b], then f is Riemann integrable on [a, b] if and only if, given any
ε > 0, there is a partition P of [a, b] such that

U (f, P ) − L(f, P ) < ε. (3)

Remarks 4.2.
1. The following two general classes of functions are Riemann integrable:
ˆ Continuous functions on [a, b]
ˆ Monotonic functions on [a, b]

2. Define f : [0, 1] → R by (
1 if x is rational
f (x) =
0 if x is irrational
Then it can easily be shown that for any partition P of [0, 1], U (f, P ) = 1 and L(f, P ) = 0. Therefore,
we have
U (f ) − L(f ) = 1
and by Theorem 4.1, f is not Riemann integrable
3. In the next chapter, we will introduce Lebesgue Integral, which is more general than Riemann
integral. For example, f is not Riemann integrable but is “Lebesgue integrable”.

22
Now we are ready to prove the main theorem of this section:
Theorem 4.2. Let {fn } be a sequence of Riemann integrable functions defined on [a, b] such that it converges
uniformly to f on [a, b]. Then f is also Riemann integrable and
Z b Z b
lim fn = f
n→∞ a a

Proof. For any ε > 0, since {fn } converges uniformly to f on [a, b], there exists N ∈ N such that for n ≥ N ,
ε
|fn (x) − f (x)| < for all x ∈ [a, b] (4)
3(b − a)

Also, as fN is Riemann integrable, there exists a partition P of [a, b] such that


ε
U (fN , P ) − L(fN , P ) <
3
By (4), we have
ε ε
L(fN − , P ) < L(f, P ) ≤ U (f, P ) < U (fN + ,P)
3(b − a) 3(b − a)
Combining the inequalities, we have
ε ε
U (f, P ) − L(f, P ) < U (fN + , P ) − L(fN − ,P)
3(b − a) 3(b − a)

= U (fN , P ) − L(fN , P ) + (b − a)
3(b − a)
ε 2ε
< + =ε
3 3
Therefore, f is Riemann integrable. Also, for n ≥ N ,
Z b Z b Z b Z b
ε ε
fn − f ≤ |fn − f | < = <ε
a a a a 3(b − a) 3
(Z )
b Z b
which implies that fn converges to f.
a a

23
Example 4.1. It can easily be shown that for each r ∈ (0, 1), the geometric series

X 1
(−1)n x2n converges uniformly to on [−r, r].
n=0
1 + x2

Therefore, by Theorem 4.2, we can integrate the power series term by term and obtain the following:

X (−1)n x2n+1
converges uniformly to arctan x on [−r, r]
n=0
2n + 1

In particular, for any x ∈ (−1, 1),



X (−1)n x2n+1
arctan(x) =
n=0
2n + 1

Question: Does the above equality holds when x = 1?

24
5 Abel’s Limit Theorem
The following are some useful properties about uniform convergence:

Theorem 5.1. Suppose {fn } converges uniformly to f on U .


1. (Subset Property) If V ⊆ U , {fn } converges uniformly to f on V .

2. (Union Property) If {fn } also converges uniformly to f on W , then it converges uniformly to f on


U ∪ W.
3. (Bounded Multiplier Property) If g(x) is a bounded function on U , {gfn } converges uniformly to
gf on U .

4. (Substitution Property) If h : U0 → U , then {fn ◦ h} converges uniformly to f ◦ h on U0 .


Proof. Exercise

Are there specials cases where pointwise convergence of nice functions implies uniform convergence? When
the domain is a closed and bounded interval, we have the following theorem:


X
Theorem 5.2. (Abel’s limit theorem) Suppose the power series ak (x − a)k converges pointwise to
k=0
S(x) on a closed and bounded interval [u, v], then it converges uniformly to S(x) on [u, v]. Moreover, we
have
X∞ X∞
lim+ S(x) = S(u) = ak (u − a)k and lim− S(x) = S(v) = ak (v − a)k
x→u x→v
k=0 k=0

n
X
Proof. We first consider the special case when a = 0 and [u, v] = [0, 1]. Let sn = ak . Since lim sn =
n→∞
k=0

X
ak = S(1), {sn } is Cauchy and hence for any ε > 0, there exists K ∈ N such that for j > n ≥ K,
k=0

j
X ε
|sj − sn | = ak <
2
k=n+1

25
j
X
Define Aj = ak for j > n and An = 0. Then for any x ∈ [0, 1] and m > n ≥ K, we have
k=n+1

m
X m
X m
X m
X
a k xk = (Ak − Ak−1 )xk = Ak x k − Ak−1 xk
k=n+1 k=n+1 k=n+1 k=n+1
m
X m−1
X m−1
X
= Ak x k − Ak xk+1 = Am xm + Ak (xk − xk+1 )
k=n+1 k=n k=n+1
m−1
X
≤ |Am |xm + |Ak |(xk − xk+1 )
k=n+1
m−1
ε m ε X ε n+1
≤ x + (xk − xk+1 ) = x <ε
2 2 2
k=n+1


X
By Cauchy criterion for uniform convergence, ak xk converges uniformly to S(x) on [0, 1]. Moreover, since
k=0
the partial sum is continuous on [0, 1], S(x) is continuous on [0, 1] and we have

X
lim S(x) = S(0) = a0 and lim S(x) = S(1) = ak
x→0+ x→1−
k=0

For the general case, we have 


[c, v]
 if c ≤ u
[u, v] ⊆ [u, c] ∪ [c, v] if u < c < v

[u, c] if v ≤ c

Then by subset and union properties in Theorem 5.1, it suffices to consider the cases [c, v] and [u, c].

x−c
For [c, v], we consider the substitution t = h(x) = and bk = ak (v − c)k . Then we have
v−c

X ∞
X
ak (x − c)k = bk tk
k=0 k=0

and x ∈ [c, v] ⇔ t ∈ [0, 1]. By substitution property in Theorem 5.1 and the special case we have already

X
proved, ak (x − c)k converges uniformly on [c, v].
k=0

c−x
For [u, c], we consider the substitution t = h(x) = and bk = ak (c − u)k . Then we have
c−u

X ∞
X
ak (x − c)k = bk tk
k=0 k=0

and x ∈ [u, c] ⇔ t ∈ [0, 1]. By substitution property in Theorem 5.1 and the special case we have already

X
proved, ak (x − c)k converges uniformly on [c, v].
k=0

26
Remarks 5.1.
1. Abel’s limit theorem is false for other types of intervals. For example,

X 1
xk = on [0, 1)
1−x
k=0

1
However, the series does not converge uniformly to on [0, 1).
1−x
2. In previous section, for any x ∈ (−1, 1),

X (−1)n x2n+1
arctan(x) =
n=0
2n + 1


X (−1)n
Moreover, when x = 1, converges by alternating series test. Hence the power series con-
n=0
2n + 1
verges pointwise on [0, 1] and by Abel’s limit theorem,

π X (−1)n
= lim− arctan(x) =
4 x→1
n=0
2n + 1


X (−1)k−1
Exercise 5.1. Prove that = ln 2.
k
k=1
1
(Hint: Consider the Taylor series of and integrate.)
1+x

27

You might also like