0% found this document useful (0 votes)
205 views10 pages

Ma 403: Real Analysis, Instructor: B. V. Limaye

(1) Pointwise convergence of a sequence of functions does not necessarily imply uniform convergence or preserve properties like continuity, differentiability, or integrability. (2) Uniform convergence is a stronger condition than pointwise convergence that requires the functions to converge uniformly across the entire domain. (3) Several examples are provided to illustrate that pointwise convergence alone is not sufficient to preserve properties or interchange limits, whereas uniform convergence is sufficient.

Uploaded by

Nikhil
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)
205 views10 pages

Ma 403: Real Analysis, Instructor: B. V. Limaye

(1) Pointwise convergence of a sequence of functions does not necessarily imply uniform convergence or preserve properties like continuity, differentiability, or integrability. (2) Uniform convergence is a stronger condition than pointwise convergence that requires the functions to converge uniformly across the entire domain. (3) Several examples are provided to illustrate that pointwise convergence alone is not sufficient to preserve properties or interchange limits, whereas uniform convergence is sufficient.

Uploaded by

Nikhil
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/ 10

UNIFORM CONVERGENCE

MA 403: REAL ANALYSIS, INSTRUCTOR: B. V. LIMAYE

1. Pointwise Convergence of a Sequence


Let E be a set and Y be a metric space. Consider functions f n : E → Y
for n = 1, 2, . . . . We say that the sequence (f n ) converges pointwise on
E if there is a function f : E → Y such that f n (p) → f (p) for every p ∈ E.
Clearly, such a function f is unique and it is called the pointwise limit of
(fn ) on E. We then write fn → f on E. For simplicity, we shall assume
Y = R with the usual metric.
Let fn → f on E. We ask the following questions:
(i) If each fn is bounded on E, must f be bounded on E? If so, must
supp∈E |fn (p)| → supp∈E |f (p)|?
(ii) If E is a metric space and each fn is continuous on E, must f be
continuous on E?
(iii) If E is an interval in R and each f n is differentiable on E, must f
be differentiable on E? If so, must fn0 → f 0 on E?
(iv) If E = [a, b] and each fn is Riemann integrable on E, must f be
Rb Rb
Riemann integrable on E? If so, must a fn (x)dx → a f (x)dx?
These questions involve interchange of two processes (one of which is
‘taking the limit as n → ∞’) as shown below.

(i) lim sup |fn (p)| = sup lim fn (p) .

n→∞ p∈E p∈E n→∞
(ii) For p ∈ E and pk → p in E, lim lim fn (pk ) = lim lim fn (pk ).
n→∞ n→∞ k→∞
d d  k→∞
(iii) lim (fn ) = lim fn .
n→∞ dx dx Zn→∞
Z b b 
(iv) lim fn (x)dx = lim fn (x) dx.
n→∞ a a n→∞

Answers to these questions are all negative.


Examples 1.1. (i) Let E := (0, 1] and define f n : E → R by

0 if 0 < x ≤ 1/n,
fn (x) :=
1/x if 1/n ≤ x ≤ 1.
Then |fn (x)| ≤ n for all x ∈ E, fn → f on E, where f (x) := 1/x.
Thus each fn is bounded on E, but f is not bounded on E.
(ii) Let E := [0, 1] and define fn : E → R by fn (x) := 1/(nx + 1). Then
each fn is continuous on E, fn → f on E, where f (0) := 1 and
f (x) := 0 if 0 < x ≤ 1. Clearly, f is not continuous on E.
(iii) (a) Let E := (−1, 1) and define fn : E → R by fn (x) := 1/(nx2 + 1).
Then each fn is differentiable on E and fn → f on (−1, 1),
1
2 MA 403: REAL ANALYSIS, INSTRUCTOR: B. V. LIMAYE

where f (0) := 1 and f (x) := 0 if 0 < |x| < 1. Clearly, f is not


differentiable on E.
(b) Let E := R and define fn : R → R by fn (x) := (sin nx)/n.
Then each fn is differentiable, fn → f on R, where f ≡ 0. But
fn0 (x) = cos nx for x ∈ R, and (fn0 ) does not converge pointwise
on R. For example, (fn0 (π)) is not a convergent sequence.
(c) Let E := (−1, 1) and define
[2 − (1 + x)n ]/n if −1 < x < 0,

fn (x) :=
(1 − x)n /n if 0 ≤ x < 1.
Then each fn is differentiable on E. (In particular, we have
fn0 (0) = −1 by L’Hôpital’s Rule.) Also, f n → f on (−1, 1),
where f ≡ 0. Also,
−(1 + x)n−1 if −1 < x < 0,

0
fn (x) =
−(1 − x)n−1 if 0 ≤ x < 1.
Further, fn0 → g on (−1, 1), where g(0) := −1 and g(x) := 0
for 0 < |x| < 1. Clearly, f 0 6= g.
(iv) (a) Let E := [0, 1] and define fn : [0, 1] → R by

1 if x = 0, 1/n!, 2/n!, . . . , n!/n! = 1,
fn (x) :=
0 otherwise.
Then each fn is Riemann integrable on [0, 1] since it is discon-
tinuous only at a finite number of points. 
1 if x is rational,
Also, fn → f on [0, 1], where f (x) :=
0 if x is irrational.
For if x = p/q with p ∈ {0, 1, 2, . . . , q} ⊂ N, then for all n ≥ q,
we have n!x ∈ {0, 1, 2, . . .} and so fn (x) = 1, while if x is an
irrational number, then fn (x) = 0 for all n ∈ N. We have seen
that the Dirichlet function f is not Riemann integrable.
(b) Let E := [0, 1], and define fn : [0, 1] → R by fn (x) := n3 xe−nx .
Then each fn is Riemann integrable and fn → f on [0, 1],
where f ≡ 0. (Use L’Hôpital’s Rule repeatedly to show that
limt→∞ t3 /est = 0 for any s ∈ R with s > 0.) However, using
Integration by Parts, we have
Z 1
1 1 1
xe−nx dx = 2 − 2 n − n for each n ∈ N,
0 n n e ne
R1
so that 0 fn (x)dx = n − (n/en ) − (n2 /en ) → ∞.
(c) Let E := [0, 1] and define fn : [0, 1] → R by fn (x) := n2 xe−nx .
As above, each fn is Riemann integrable, and fn → f on [0, 1],
R1
where f ≡ 0, and 0 fn (x)dx = 1 − (1/en ) − (n/en ) → 1, which
R1
is not equal to 0 f (x)dx = 0.

2. Uniform Convergence of a Sequence


In an attempt to obtain affirmative answers to the questions posed at
the beginning of the previous section, we introduce a stronger concept of
convergence.
UNIFORM CONVERGENCE 3

Let E be a set and consider functions f n : E → R for n = 1, 2, . . . . We


say that the sequence (fn ) of functions converges uniformly on E if there
is a function f : E → R such that for every  > 0, there is n 0 ∈ N satisfying
n ≥ n0 , p ∈ E =⇒ |fn (p) − f (p)| < .
Note that the natural number n0 mentioned in the above definition may
depend upon the given sequence (fn ) of functions and on the given positive
number , but it is independent of p ∈ E. Clearly, such a function f is
unique and it is called the uniform limit of (f n ) on E. We then write
fn ⇒ f on E. Obviously, fn ⇒ f on E =⇒ fn → f on E, but the converse
is not true : Let E := (0, 1] and define f n (x) := 1/(nx + 1) for 0 < x ≤ 1. If
f (x) := 0 for x ∈ (0, 1], then fn → f on (0, 1], but fn 6⇒ f on (0, 1]. To see
this, let  := 1/2, note that there is no n 0 ∈ N satisfying
1 1
|fn (x) − f (x)| = < for all n ≥ n0 and for all x ∈ (0, 1],
nx + 1 2
since 1/(nx + 1) = 1/2 when x = 1/n, n ∈ N.
A sequence (fn ) of real-valued functions defined on a set E is said to be
uniformly Cauchy on E if for every  > 0, there is n 0 ∈ N satisfying
m, n ≥ n0 , p ∈ E =⇒ |fm (p) − fn (p)| < .
Proposition 2.1. (Cauchy Criterion for Uniform Convergence of a
Sequence) Let (fn ) be a sequence of real-valued functions defined on a set
E. Then (fn ) is uniformly convergent on E if and only if (f n ) is uniformly
Cauchy on E.
Proof. =⇒) Let fn ⇒ f . For all m, n ∈ N and p ∈ E, we have
|fm (p) − fn (p)| ≤ |fm (p) − f (p)| + |f (p) − fn (p)|.
⇐=) For each p ∈ E, (fn (p)) is a Cauchy sequence in R, and so it converges
to a real number which we denote by f (p). Let  > 0. There is n 0 ∈ N
satisfying
m, n ≥ n0 , p ∈ E =⇒ |fm (p) − fn (p)| < .
For any m ≥ n0 and p ∈ E, letting n → ∞, we have |fm (p) − f (p)| ≤ . 
We have the following useful test for checking the uniform convergence of
(fn ) when its pointwise limit is known.
Proposition 2.2. (Test for Uniform Convergence of a Sequence)
Let fn and f be real-valued functions defined on a set E. If f n → f on
E, and if there is a sequence (an ) of real numbers such that an → 0 and
|fn (p) − f (p)| ≤ an for all p ∈ E, then fn ⇒ f on E.
Proof. Let  > 0. Since an → 0, there is n0 ∈ N such that n ≥ n0 =⇒ an < ,
and so |fn (p) − f (p)| <  for all p ∈ E. 
Example 2.3. Let 0 < r < 1 and fn (x) := xn for x ∈ [−r, r]. Thenfn (x) →
0 for each x ∈ [−r, r]. Since r n → 0 and |fn (x) − 0| ≤ r n for all x ∈ [−r, r],
(fn ) is uniformly convergent on [−r, r].

Let us now pose the four questions stated in the last section with ‘con-
vergence’ replaced by ‘uniform convergence’. We shall answer them one by
one, but not necessarily in the same order.
4 MA 403: REAL ANALYSIS, INSTRUCTOR: B. V. LIMAYE

Uniform Convergence and Boundedness.


Proposition 2.4. Let fn and f be real-valued functions defined on a set E.
If fn ⇒ f on E and each fn is bounded on E, then f bounded on E.
Proof. There is n0 ∈ N such that n ≥ n0 , p ∈ E =⇒ |fn (p)−f (p)| < 1. Also,
since fn0 is bounded on E, there is α0 such that p ∈ E =⇒ |fn0 (p)| ≤ α0 .
Hence p ∈ E =⇒ |f (p)| ≤ |f (p) − fn0 (p)| + |fn0 (p)| < 1 + α0 . 

The converse of the above result is not true, that is, each f n as well
as f bounded on E and fn → f 6=⇒ fn ⇒ f on E. For example, let
E := (0, 1], fn (x) := 1/(nx + 1) and f ≡ 0.
Given a set E, let B(E) denote the set of all real-valued bounded functions
defined on E. For f, g in B(E), define d(f, g) := sup{|f (p) − g(p)| : p ∈ E}.
Then it is easy to see that d is a metric on B(E), known as the sup-metric
on B(E). Also, by Proposition 2.2, for f n and f in B(E), we have fn ⇒ f on
E if and only if d(fn , f ) → 0, that is, (fn ) converges to f in the sup-metric on
B(E). Similarly, (fn ) is uniformly Cauchy on E if and only if d(f n , fm ) → 0
as n, m → ∞, that is, (fn ) is a Cauchy sequence in the sup-metric on B(E).
Thus Propositions 2.1 and 2.4 show that B(E) is a complete metric space.
Also, under the hypotheses of Proposition 2.4, we have

sup |fn (p)| − sup |f (p)| ≤ d(fn , f ) → 0,

p∈E p∈E

and so, supp∈E |fn (p)| → supp∈E |f (p)|.

Uniform Convergence and Integration.


Proposition 2.5. Let (fn ) be a sequence of real-valued functions defined on
[a, b]. If fn ⇒ f on [a, b] and each fn is Riemann integrable on [a, b], then
Rb Rb
f is Riemann integrable on [a, b] and a fn (x)dx → a f (x)dx.
Proof. Since fn ⇒ f and each fn is bounded, we see that f is bounded on
[a, b] by Proposition 2.4. For n ∈ N, let α n := d(fn , f ), where d denotes the
sup-metric on B([a, b]). For each n ∈ N and x ∈ [a, b], we have |f n (x) −
f (x)| ≤ αn , that is, fn (x) − αn ≤ f (x) ≤ fn (x) + αn , and so
L(fn ) − αn (b − a) ≤ L(f ) ≤ U (f ) ≤ U (fn ) + αn (b − a).
But since fn is Riemann integrable, we have L(fn ) = U (fn ), and hence
0 ≤ U (f ) − L(f ) ≤ 2αn (b − a) → 0 as n → ∞. Thus L(f ) = U (f ), that is,
f is Riemann integrable on [a, b]. Also,
Z b Z b Z b
fn (x)dx − αn (b − a) ≤ f (x)dx ≤ fn (x)dx + αn (b − a),
a a a
R
b Rb
that is, a fn (x)dx − a f (x)dx ≤ αn (b − a) → 0 as n → ∞. 

The converse of the above result is not true, that is, each f n as well
Rb
as f Riemann integrable on [a, b], fn → f on [a, b] and a fn (x)dx →
Rb
a f (x)dx 6=⇒ fn ⇒ f . For example, if fn (x) := 1/(nx + 1) for x ∈ [0, 1],
UNIFORM CONVERGENCE 5

f (0) := 1, f (x) := 0 for x ∈ (0, 1], then f n 6⇒ f on [0, 1], but f is integrable
and
Z 1 Z 1
ln(nx + 1) 1 ln(1 + n)
fn (x)dx = = →0= f (x)dx.
0 n 0 n 0
Uniform Convergence and Continuity.
Proposition 2.6. Let (fn ) be a sequence of real-valued functions defined on
a metric space E. If fn ⇒ f on E and each fn is continuous on E, then f
is continuous on E.
Proof. Let  > 0. There is n0 ∈ N such that p ∈ E =⇒ |fn0 (p)−f (p)| < /3.
Consider p0 ∈ E. Since fn0 is continuous at p0 , there is δ > 0 such that
p ∈ E, d(p, p0 ) < δ =⇒ |fn0 (p) − fn0 (p0 )| < /3. and hence
|f (p) − f (p0 )| ≤ |f (p) − fn0 (p)| + |fn0 (p) − fn0 (p0 )| + |fn0 (p0 ) − f (p0 )| < ,
establishing the continuity of f at p 0 ∈ E. 
The converse of the above result is not true, that is, each f n as well as f
continuous on a metric space E, fn → f on E 6=⇒ fn ⇒ f . For example,
let fn (x) := nxe−nx and f (x) := 0 for x ∈ [0, 1]. Since fn (0) = 0 and for
x ∈ (0, 1], fn (x) → 0 as n → ∞ by L’Hôpital’s Rule, we see that f n → f .
But there is no n0 ∈ N such that n ≥ n0 , x ∈ [0, 1] =⇒ |nxe−nx − 0| < 1,
since |nxe−nx | = e−1 for x = 1/n, n ∈ N. However, the following partial
converse holds.
Proposition 2.7. (Dini’s Theorem) Let (f n ) be a sequence of real-valued
functions defined on a compact metric space E. If f n → f on E, each fn and
f are continuous on E, and (fn ) is a monotonic sequence (that is, f n ≤ fn+1
for all n ∈ N, or fn ≥ fn+1 for all n ∈ N), then fn ⇒ f on E.
For a proof, see Theorem 7.13 of [3].
The following examples show that neither the compactness of the metric
space E nor the continuity of the function f can be dropped from Dini’s
Theorem: (i) E := (0, 1] and fn (x) := 1/(nx + 1), x ∈ E, (ii) E := [0, 1] and
fn (x) := xn , x ∈ E.
Uniform Convergence and Differentiation. Answers to the questions
regarding differentiation posed in the last section are not affirmative even
when fn ⇒ f on anp interval of R.
(a) Let fn (x) := x2 + (1/n2 ) and f (x) := |x| for x ∈ [−1, 1]. Since
p √ p 1
2 2 2 2 2
|fn (x) − f (x)| = x + (1/n ) − x ≤ x + (1/n ) − x =
2
n
for all n ∈ N and x ∈ [−1, 1], Proposition 2.2 shows that f n ⇒ f on [−1, 1].
Although each fn is differentiable on [−1, 1], the limit function f is not.
(b) In Example 1.1 (iii) (b), fn ⇒ f on R, each fn differentiable, but (fn0 )
does not converge pointwise.
(c) In Example 1.1 (iii) (c), fn ⇒ f on (−1, 1) and each fn as well as f is
differentiable on (−1, 1), and fn0 → g on (−1, 1), where g 6= f 0 .
However, if we assume the uniform convergence of the ‘derived sequence’
(fn0 ) along with the convergence of the sequence (f n ) at only one point of
the interval, we have a satisfactory answer.
6 MA 403: REAL ANALYSIS, INSTRUCTOR: B. V. LIMAYE

Proposition 2.8. Let (fn ) be a sequence of real-valued functions defined


on [a, b]. If (fn ) converges at one point of [a, b], each f n is continuously
differentiable on [a, b] and (fn0 ) converges uniformly on [a, b], then there is
f : [a, b] → R such that f is continuously differentiable on [a, b], f n0 → f 0 on
[a, b] and in fact, fn ⇒ f on [a, b].
Proof. Let x0 ∈ [a, b] and c0 ∈ R be such that fn (x0 ) → c0 . Also, let each
fn be continuously differentiable and g : [a, b] → R be such that f n0 ⇒ g
on [a, b]. By Proposition 2.6, the function g is continuous on [a, b]. Define
f : [a, b] → R by
Z x
f (x) := c0 + g(t)dt for x ∈ [a, b].
x0

By part (ii) of the Fundamental Theorem of Calculus (FTC), f 0 exists on


[a, b] and f 0 (x) = g(x) for x ∈ [a, b]. Thus f is continuously differentiable
on [a, b] and fn0 → g = f 0 . Also, by part (i) of the FTC, we have
Z x
fn (x) = fn (x0 ) + fn0 (t)dt for x ∈ [a, b].
x0
Hence for n ∈ N and x ∈ [a, b], we obtain
Z x 
|fn (x) − f (x)| ≤ |fn (x0 ) − c0 | + fn0 (t) − g(t) dt

x0
≤ |fn (x0 ) − c0 | + |x − x0 | sup |fn0 (t) − g(t)|
t∈[a,b]

≤ |fn (x0 ) − c0 | + (b − a)d(fn0 , g).


Thus fn ⇒ f on [a, b] by Proposition 2.2. 
The converse of the above result is not true, that is, each f n as well
as f continuously differentiable on [a, b], f n ⇒ f on [a, b], fn0 → f 0 on
[a, b] 6=⇒ fn0 ⇒ f 0 . For example, let fn (x) := (nx + 1)e−nx /n and f (x) := 0
for x ∈ [0, 1]. Since fn0 (x) = −nxe−nx for each n ∈ N and all x ∈ [0, 1],
each fn is monotonically decreasing on [0, 1]. As f n (0) = 1/n, we obtain
|fn (x) − f (x)| ≤ 1/n for all x ∈ [0, 1], and so f n ⇒ f on [0, 1]. Also, we have
seen after the proof of Proposition 2.6 that f n0 → f 0 , but fn0 6⇒ f 0 on [0, 1].
Remark 2.9. Proposition 2.8 holds if we drop the word ‘continuously’ ap-
pearing (two times) in its statement, but then the proof is much more in-
volved. See Theorem 7.17 of [3].
The results in Propositions 2.4, 2.5, 2.6 and 2.8 are summarized in the
following theorem.
Theorem 2.10. (i) The uniform limit of a sequence of real-valued
bounded functions defined on a set is bounded.
(ii) The uniform limit of a sequence of Riemann integrable functions
defined on [a, b] is Riemann integrable, and its Riemann integral is
the limit of the sequence of termwise Riemann integrals, that is, if
(fn ) is uniformly convergent to f on [a, b] and each f n is Riemann
integrable on [a, b], then the function f is Riemann integrable on
Rb Rb
[a, b] and a f (x)dx = limn→∞ a fn (x)dx.
UNIFORM CONVERGENCE 7

(iii) The uniform limit of a sequence of continuous functions defined on


a metric space is continuous.
(vi) If a sequence of continuously differentiable functions defined on [a, b]
is convergent at one point of [a, b] and if the ‘derived’ sequence is
uniformly convergent on [a, b], then the given sequence converges
uniformly on [a, b], the uniform limit is continuously differentiable
on [a, b] and its derivative is the limit of the sequence of termwise
derivatives, that is, if (fn ) converges at one point of [a, b], each f n is
continuously differentiable on [a, b] and (f n0 ) is uniformly convergent
on [a, b], then (fn ) converges uniformly to a continuously differen-
tiable function f on [a, b], and f 0 (x) = limn→∞ fn0 (x) for all x ∈
[a, b].

3. Uniform Convergence of a series


The reader is assumed to be familiar with the elementary theory of series
of real numbers. (See, for example, Chapter 9 of [1], or Chapter 3 of [3].)
Let (fk ) be a sequence of real-valued functions defined on a set E. Con-
sider the sequence (sn ) of real-valued functions on E defined by
n
X
sn := f1 + · · · + fn = fk .
k=1

Note: Just as the sequence (fk ) determines the sequence (sn ), so does (sn )
determine (fk ): If we let s0P= 0, then we have fk = sk − sk−1 for all k ∈ N.

We say that the series k=1 fk converges pointwise on E ifP the se-
quence (sn ) converges pointwise on E, and we say that the series ∞ k=1 fk
converges uniformly on E if the sequence (s n ) converges uniformly on
E.
P∞For n ∈ N, the function sn is called the nth partial sum of the series
k=1 fk and if sn → s, then the function s is called its sum.
Results about convergence / uniform convergence of sequences of func-
tions carry over to corresponding results about convergence / uniform con-
vergence of series of functions.

Proposition 3.1. (Cauchy Criterion for Uniform Convergence of a


Series) Let (fk )P be a sequence of real-valued functions defined on a set E.
Then the series ∞ k=1 fk converges uniformly on E if and only if for every
 > 0, there is n0 ∈ N such that
Xm
m ≥ n ≥ n0 , p ∈ E =⇒ fk (p) < .

k=n

Proof. Use Proposition 2.1 for the sequence (s n ) of partial sums. 

Proposition 3.2. (Weierstrass M-Test for Uniform Absolute Con-


vergence of a Series) Let (fk ) be a sequence of real-valued functions
defined on a set E. Suppose there is a sequence P∞(M k ) in R such that
|fk (p)|
P≤ Mk for all k ∈ N and all p ∈ E. If k=1 Mk is convergent,
then ∞ k=1 f k converges uniformly and absolutely on E.
8 MA 403: REAL ANALYSIS, INSTRUCTOR: B. V. LIMAYE

Proof. Note that


Xm X m m
X
fk (p) ≤ |fk (p)| ≤ Mk for all m ≥ n,


k=n k=n k=n
and use Proposition 3.1. 
P∞
Examples 3.3. (i) Consider the series k=0 xk , where x ∈ (−1, 1). If
r < 1, thenPthe series converges uniformly on {x ∈ R : |x| ≤ r} since
the series ∞ k
k=0 Mk is convergent, where Mk := r for k = 0, 1, . . .
(ii) For k ∈ N, let fP k 2
k (x) := (−1) (x + k)/k , where x ∈ [0, 1]. We show

that the series k=1 fk converges uniformly on [0, 1]. For k ∈ N,
let gk (x) := (−1)k x/k 2 , where x ∈P [0, 1]. Letting Mk := 1/k 2 for
k ∈ N, weP observe that the series ∞ k=0 Mk is convergent. Hence
the series ∞ k=1 g k (x) converges uniformly on [0, 1]. Also, the se-
ries ∞ k /k converges uniformly on [0, 1], being a convergent
P
k=1 (−1)
series of constants. P Since fk (x) = gk (x) + (−1)k /k for k ∈ N and
x ∈ [0, 1], the series ∞ k=1 fk (x) converges uniformly on [0, 1].
This example also shows that the converse of Weierstrass’ M -test
does not hold: If Mk := supx∈[0,1] |fk (x)| = (1 + k)/k 2 for k ∈ N,
then ∞ 2
P P∞
k=1 Mk does not converge, since k=1 1/k converges, but
P ∞
k=1 1/k diverges.
Proposition 3.4. (Dirichlet’s Test for Uniform Conditional Con-
vergence of a Series) Let (fk ) be a monotonic sequence of real-valued
functions defined on a set E such that f k ⇒ 0 on E. If (gk ) is a sequence
of
P∞ real-valued functions defined on E such that the partialP∞ sums of the series
g
k=1 k are uniformly bounded on E, then the series k=1 fk gk converges
P∞ k
uniformly on E. In particular, the series k=1 (−1) fk converges uniformly
on E.
Proof. For each p ∈ E, the series ∞
P
k=1 fk (p)gk (p) converges in R by Dirich-
let’s Test for conditional convergence of a series of real numbers. (See,
for example, P Proposition 9.20 of [1], or Theorem 3.42 of [3].)PFor p ∈ E,
∞ n
let H(p)Pn := k=1 fk (p)gk (p). Also, for n ∈ N, let Gn := k=1 gk and
Hn := k=1 fk gk . Further, let β ∈ R be such that |Gn (p)| ≤ β for all n ∈ N
and all p ∈ E. Then by using the partial summation formula
n
X n−1
X
fk gk = (fk − fk+1 )Gk + fn Gn for all n ≥ 2,
k=1 k=1
we have |H(p) − Hn (p)| ≤ 2β|fn+1 (p)| for all p ∈ E.PSince fn+1 ⇒ 0 on
E, it follows that Hn ⇒ H on E, that is, the series ∞ k=1 fk gk converges
uniformly on E.
In particular, letting gk (p) := (−1)k for all k ∈ N and p ∈ E, and noting
that |Gn (p)| ≤ 1 for allP n ∈ N and all p ∈ E, we obtain the uniform
convergence of the series ∞ k
k=1 (−1) fk on E. 
Example 3.5. Let E := [0, 1] and fk (x) := xk /k for k ∈ N and x ∈ [0, 1].
Then (fk ) is a momotonically decreasing sequence and since |f k (x)| ≤ 1/k
for k ∈ N and x P
∈ [0, 1], we see that fk ⇒ 0 on [0, 1] by Proposition 2.2.
Hence the series ∞ k k
k=1 (−1) x /k converges uniformly on [0, 1].
UNIFORM CONVERGENCE 9

Results regarding the boundedness, Riemann integrability, continuity and


differentiability of the sum function of a convergent series of functions can be
easily deduced from the corresponding results for the sequence of its partial
sums.
Theorem 3.6. (i) The sum function of a uniformly convergent series
of real-valued bounded functions defined on a set is bounded.
(ii) The sum function of a uniformly convergent series of Riemann in-
tegrable functions defined on [a, b] is Riemann integrable,
P∞ and the
series can be integrated term by term, that is, if k=1 fk is uni-
formly convergent on [a, b] and each f k is Riemann integrable on
[a, b], then the function ∞
P
f
k=1 k is Riemann integrable on [a, b] and
Z b X∞  ∞ Z b
X
fk (x) dx = fk (x)dx.
a k=1 k=1 a

(iii) The sum function of a uniformly convergent series of real-valued


continuous functions defined on a metric space is continuous.
(vi) If a series of continuously differentiable functions defined on [a, b]
is convergent at one point of [a, b] and if the ‘derived’ series is
uniformly convergent on [a, b], then the given series converges uni-
formly on [a, b], the sum function is continuously differentiable on
[a,
P∞ b] and the series can be differentiated term by term, that is, if
k=1 fk converges at onePpoint of [a, b], each f k is continuously dif-
∞ 0
P∞ on [a, b] and k=1 fk is uniformly convergent on [a, b],
ferentiable
then k=1 fk converges uniformly to a continuously differentiable
function, and
X ∞ 0 X ∞
fk (x) = fk0 (x) for all x ∈ [a, b].
k=1 k=1

Proof. The results follow by applying Theorem 2.10 to the sequence of par-
tial sums of the given series. 

4. Two Celebrated Theorems on Uniform Approximation


We have seen in Proposition 2.6 that a uniform limit of a sequence of
continuous functions on a metric space is continuous. In this section, we
reverse the procedure and ask whether every continuous function on a closed
and bounded interval of R is the uniform limit of a sequence of some ‘special’
continuous functions.
For a function f : [0, 1] → R and n ∈ N, we define the nth Bernstein
polynomial of f by
n    
X n k k
Bn (f ) := f x (1 − x)n−k .
k n
k=0

Theorem 4.1. (Polynomial Approximation Theorem of Weierstrass)


If f : [0, 1] → R is continuous, then Bn (f ) ⇒ f on [0, 1]. Consequently,
every real-valued continuous function on [0, 1] is the uniform limit of a se-
quence of real-valued polynomial functions.
10 MA 403: REAL ANALYSIS, INSTRUCTOR: B. V. LIMAYE

For a proof, see Theorem 7.26 of [3], or Corollary 3.12 of [2].


Remark 4.2. Theorem 4.1 can be used to prove that every real-valued
continuous function on any closed and bounded interval [a, b] is the uniform
limit of a sequence of real-valued polynomial functions. Let φ : [0, 1] → [a, b]
be defined by φ(x) := (1 − x)a + xb for x ∈ [0, 1]. Then φ is a bijective
continuous function and its continuous inverse φ −1 : [a, b] → [0, 1] is given
by φ−1 (t) = (t − a)/(b − a) for t ∈ [a, b]. Given a continuous real-valued
function g on [a, b], consider the continuous function f := g ◦ φ defined on
[0, 1]. If (Pn ) is a sequence of polynomial functions such that P n ⇒ f on
[0, 1], and if we let Qn := Pn ◦ φ−1 , then since Qn (t) = Pn (t − a)/(b − a)
for t ∈ [a, b], each Qn is a polynomial function, and Qn ⇒ f ◦ φ−1 = g on
[a, b].
Instead of polynomials, let us now consider trigonometric polynomials for
approximating a function. They are given by
Xn
a0 + (ak cos kx + bk sin kx) for n ∈ N,
k=1
where a0 , a1 , a2 , . . . , b1 , b2 , . . . are real numbers. For a Riemann integrable
function f on [−π, π], we define the Fourier coefficients of f by
Z π
1
a0 (f ) := f (t)dt, and for k ∈ N,
2π −π
1 π 1 π
Z Z
ak (f ) := f (t) cos kt dt, bk (f ) := f (t) sin kt dt.
π −π π −π
The series a0 (f ) + ∞
P 
k=1 ak (f ) cos kx + bk (f ) sin kx of functions defined
on [−π, π] is called the Fourier series of the function f . For n = 0, 1, 2, . . .,
let sn (f ) denote the nth partial sum of this series, and consider the arith-
metic means of these partial sums given by
s0 (f ) + s1 (f ) · · · + sn (f )
σn (f ) := for n = 0, 1, 2 . . .
n+1
Theorem 4.3. (Trigonometric Polynomial Approximation Theo-
rem of Fejér) If f : [−π, π] → R is continuous and f (−π) = f (π), then
σn (f ) ⇒ f on [−π, π]. Consequently, every real-valued continuous function
on [−π, π] having the same value at −π and π is the uniform limit of a
sequence of real-valued trigonometric polynomial functions.
For a proof, see Theorem 8.15 and Exercise 8.15 of [3], or Theorem 3.13
of [2].

References
[1] S. R. Ghorpade and B. V. Limaye, A Course in Calculus and Real Analysis,
Springer International Ed., New Delhi, 2006.
[2] B. V. Limaye, Functional Analysis, New Age International, 2nd Ed., New Delhi,
1996.
[3] W. Rudin, Principles of Mathematical Analysis, 3rd Ed., McGraw Hill, New Delhi,
1976.

You might also like