Series and sequence mathematical topics
Series and sequence mathematical topics
Sequences
and Series
Theory and Practice
Sequences and Series
Ana Alves de Sá • Bento Louro
The translation was done with the help of an artificial intelligence machine translation
tool. A subsequent human revision was done primarily in terms of content.
Translation from the Portuguese language edition: “Sucessões e Séries: Teoria e Prática”
by Ana Alves de Sá and Bento Louro, © Ana Maria de Souza Alves de Sá and Bento
José Carrilho Miguens Louro 2014. Published by Escolar Editora. All Rights Reserved.
© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer
Nature Switzerland AG 2024
This work is subject to copyright. All rights are solely and exclusively licensed by the
Publisher, whether the whole or part of the material is concerned, specifically the rights of
reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in
any other physical way, and transmission or information storage and retrieval, electronic
adaptation, computer software, or by similar or dissimilar methodology now known or
hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc.
in this publication does not imply, even in the absence of a specific statement, that such
names are exempt from the relevant protective laws and regulations and therefore free
for general use.
The publisher, the authors and the editors are safe to assume that the advice and in-
formation in this book are believed to be true and accurate at the date of publication.
Neither the publisher nor the authors or the editors give a warranty, expressed or implied,
with respect to the material contained herein or for any errors or omissions that may have
been made. The publisher remains neutral with regard to jurisdictional claims in published
maps and institutional affiliations.
This Springer imprint is published by the registered company Springer Nature Switzerland
AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Any reader who has completed High School should be able to comprehend
Chap. 1, and the same holds true for Chap. 2, with the possible exception
of the Integral Test. For those unfamiliar with integral calculus, skipping
this section will not detract from their understanding of the rest of the
chapter. To understand Chap. 3, the reader must have basic knowledge of
differential and integral calculus. We go beyond the typical introduction
to function series by delving into the study of Fourier series. This topic
will prove to be especially useful when studying Differential Equations, but
no additional knowledge is required beyond what is already necessary for a
general understanding of function series.
At the end of each chapter, there are two lists of exercises: one titled ”Solved
Exercises” followed by the respective solutions, and another titled ”Proposed
Exercises,” which are intended for the reader’s practice and whose answers
can be found at the end of the book.
v
vi
The exercises featured in this book were used during practical classes and
exams for the subjects we taught. We extend our sincere appreciation to the
students whose inquiries aided in enhancing this publication, as well as to
our fellow educators who assisted us in teaching Calculus courses at Nova
School of Science & Technology. Their thorough review of the text and
copious suggestions were invaluable contributions.
This section consists of a list of notations. They are quite common. Our aim is to avoid ambiguity and
to help the reader who, here or there, may be used to a different notation.
N0 = N ∪ {0}
.
R+
.0 = R ∪ {0}
+
intervals on .R: (suppose .a, b ∈ R, with .a < b) are the following sets:
.
[a, b[ = {x ∈ R : a ≤ x < b}; [a, +∞[ = {x ∈ R : a ≤ x};
]a, b] = {x ∈ R : a < x ≤ b}; ]a, +∞[ = {x ∈ R : a < x};
[a, a] = {a}; ] − ∞, +∞[ = R
vii
viii
n
. ai = an , .an ∈ R and .n ∈ N
i=n
n
. ai = am + am+1 + · · · + an , .ai ∈ R, i = m, m + 1, . . . , n and .m, n ∈ N with .n > m
i=m
factorial of .n ∈ N0 : .0! = 1, 1! = 1 and .n! = n × (n − 1)!
(in a suggestive way, .n! = n × (n − 1) × · · · × 3 × 2 × 1)
n n!
number of p-combinations of n distinct objects (.n, p ∈ N0 , .n ≥ p): . =
p p! × (n − p)!
n
n k n−k
Newton Binomial: if .a, b ∈ R, .n ∈ N then .(a + b)n = a b
k
k=0
Method of Mathematical Induction: to prove that a proposition .P (n) is true for all .n ∈ N
proceed as follows:
1) prove that .P (1) is true
2) prove hereditary, that is, assume that the Induction Hypothesis .P (n) is true, and then
prove the Induction Thesis .P (n + 1) is true
integer part of .x ∈ R+ : .Int(x), the largest integer less than or equal to x
supremum of .X ⊂ R: a real number w such that w is an upper bound of X and .w ≤ z for all
z upper bound of X; we write .w = sup(X); if X is bounded from above, .sup(X) exists
maximum of .X ⊂ R: if .sup(X) ∈ X it is called the maximum of X and is denoted by .max(X)
infimum of .X ⊂ R: a real number u such that u is a lower bound of X and .u ≥ y for all y
lower bound of X; we write .u = inf(X); if X is bounded from below, .inf(X) exists
a function of the set X into the set Y is a relation between X and Y such that, to each
element .x ∈ X corresponds an unique element .y ∈ Y , and we will write .f : X → Y . X is the
domain of f and the set .{y ∈ Y : ∃x ∈ X, y = f (x)} is the range of f
ix
a) .g (x) = 0, ∀x ∈ I \ {a}.
L’Hôpital’s Rule (another version): Let .b ∈ R, f and g two differentiable functions on .]b, +∞[
such that:
Preface v
2 Numerical Series 67
2.1 Generalization of the Addition Operation . . . . . . . . . . . . . . . . . . . . . . . . . 67
2.2 Definition of Series: Convergence – General Properties . . . . . . . . . . . . . . . . . 69
2.3 Alternating Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
2.4 Absolute Convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
2.5 Series of Nonnegative Terms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
2.6 Products of Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
2.7 Solved Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
2.8 Proposed Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
x
Contents xi
Bibliography 337
Index 339
Sequences of Real
Numbers 1
Example 1.1.1 The sequences of general terms an = n2 and bn = cos(n) Figure 1.1: The graph of the
sequence un = n/(n + 1)
are illustrated in Fig. 1.2.
Notes:
Figure 1.2: A bounded below sequence (a) and a bounded sequence (b)
Figure 1.3: A bounded above and unbounded below sequence (a) and a sequence not
bounded above or below (b)
|un | ≤ M, ∀n ∈ N.
.
4 Sequences of Real Numbers
n+2
Example 1.1.6 The sequence un = (−1)n is bounded (see Fig. 1.4):
n
n n + 2 2
.|un | = (−1) = 1 + ≤ 3, ∀n ∈ N.
n n
2n
Example 1.1.7 The sequence un = is strictly increasing. We can
3n + 7
prove this by computing and simplifying the expression un+1 − un :
5n
Example 1.1.8 The sequence un = is strictly decreasing. In fact,
n2 + 1
5(n + 1) 5n
un+1 − un = −
(n + 1)2 + 1 n2 + 1
5(n + 1)(n2 + 1) − 5n(n2 + 2n + 2)
. =
(n2 + 1)(n2 + 2n + 2)
5(n2 + n − 1)
= − < 0, ∀n ∈ N.
(n2 + 1)(n2 + 2n + 2)
u
. n+1 − un = (n + 1)2 − n2 = n2 + 2n + 1 − n2 = 2n + 1 > 0, ∀n ∈ N.
n + (−1)n
Example 1.1.12 The sequence un = is not monotonic (see
n2
Fig. 1.5).
⎧
⎪
⎪ n+1−1 n+1 n3 − (n + 1)3
⎪
⎨ (n + 1)2 − = < 0, if n is even
n2 n2 (n + 1)2
.un+1 − un =
⎪
⎪ n+1+1 n−1 n2 + n + 1
⎪
⎩ − = 2 > 0, if n is odd.
(n + 1) 2 n2 n (n + 1)2
6 Sequences of Real Numbers
n + (−1)n
Figure 1.5: The sequence un =
n2
We will now discuss two distinct types of sequences that hold significant
importance due to their properties: arithmetic sequences (also known as
arithmetic progressions) and geometric sequences (also known as geometric
progressions).
• Constant if .r = 0
a2 = a1 + r
a3 = a2 + r = a1 + 2r
a4 = a3 + r = a1 + 3r
. ..
.
an = an−1 + r = a1 + (n − 1) r
..
.
The general term of the arithmetic sequence with common difference r
is given by the expression
a = a1 + (n − 1) r, ∀n ∈ N.
. n
At first glance, the sum of the first terms of an arithmetic sequence may
seem like a time-consuming problem to solve.
However, according to legend, Gauss, one of the greatest mathematicians of
all time, solved the question of calculating the sum of the first 100 natural
numbers in a few minutes when he was still a child. What reasoning would Carl Friedrich Gauss
(1777–1855) was a Ger-
have allowed him to arrive so quickly at the result?
man mathematician. He
If we write the integers from 1 to 100, we find that: worked in a wide variety of
areas in both Mathematics
and Physics. He was a
• The sum of the first and last numbers is 101. gifted child. His first
mathematical discoveries
• The sum of the second and penultimate numbers is 101. were made in his teens.
His magnum opus, Disqui-
• The sum of the third and third-to-last numbers is 101, and so on. sitiones Arithmeticae, was
completed in 1798, when
There are precisely 50 pairs of numbers whose sum is a constant of value Gauss was only 21 years
old. (Source of image:
101, so it is concluded that the sum of the first 100 natural numbers is Oil portrait by Gottlieb
.101 × 50 = 5050. Wilhelm Emil Biermann)
a1 + an
Sn =
. × n.
2
Let us prove this using the method of induction.
For .n = 1, the equality is trivial. The property is proven if it is hereditary.
If we admit that equality is valid for n, then
8 Sequences of Real Numbers
a1 + an
Sn+1 = Sn + an+1 = × n + an+1
2
a1 + a1 + (n − 1)r
= × n + a1 + nr
2
(n − 1)r
= n a1 + × n + a1 + nr
2
n−1
= (n + 1)a1 + + 1 × nr
. 2
n+1
= (n + 1)a1 + × nr
2
nr
= (n + 1) a1 +
2
a1 + a1 + n r
= (n + 1)
2
a1 + an+1
= (n + 1) ,
2
which confirms that the equality holds for .n + 1.
a1 = k
.
an+1 = an × r, ∀n ∈ N,
where .k, r ∈ R\{0}, and r is known as the common ratio of the geometric
sequence.
1 1 1 1
For example, . , , , . . . is a geometric sequence with .r = .
10 100 1000 10
Similar to arithmetic sequences, it is possible to calculate any term of the
sequence knowing the first term and the common ratio. Considering the
definition of a geometric sequence with common ratio r, we obtain the
following:
a2 = a1 × r
a 3 = a2 × r = a 1 × r 2
a4 = a3 × r = a 1 × r 3
.
..
.
an = a1 × rn−1
..
.
1.1. Sequences of Real Numbers 9
• Increasing if .a1 > 0 and .r > 1 or if .a1 < 0 and .0 < r < 1
• Decreasing if .a1 > 0 and .0 < r < 1 or if .a1 < 0 and .r > 1
• Constant if .r = 1
There is a well-known legend that a Hindu prince, fascinated with the chess
game, decided to reward its inventor, who asked for this reward to be given
in wheat, receiving one grain for the first square of the chessboard, two
grains for the second square, four grains for the third, and so on, doubling
the number of grains from the last square considered. We conclude that the
number of wheat grains is represented by the sum of the first 64 terms of a
geometric sequence with common ratio 2.
S64 = 1 + 2 + 22 + 23 + . . . + 263 .
.
Sn = a1 + a 2 + a 3 + . . . + a n
. (1.1)
be the sum of the first n terms. Multiplying .Sn by the ratio r, we obtain
r Sn = r a1 + r a2 + r a3 + . . . + r an = a2 + a3 + a4 + . . . + an+1 . (1.2)
.
that is,
a1 − an+1 a1 − a1 r n 1 − rn
Sn =
. = = a1 .
1−r 1−r 1−r
10 Sequences of Real Numbers
wheat grains (which not even the richest prince would be able to pay).
Example 1.1.13 Let us consider Fig. 1.6, and let .(An ) be the sequence of
the areas of the semicircles delimited by the lines .C1 , .C2 , .C3 , .. . .. We show
that .(An ) is a geometric sequence and express its general term as follows:
1
Each semicircle has an area . π r2 , where r is the radius of the respective
2
circumference. In Fig. 1.6a, the radius of the circumference is r; in Fig. 1.6b,
r
each semicircle has a radius of . ; in Fig. 1.6c, each semicircle has a radius
2
r
of . ; and so on. Therefore, we can write the first terms of the sequence
4
.An .
1
A1 = π r2
2
1 r 2 r2
. A2 = π ×2=π
2 2 4
1 r 2 r2
A3 = π ×4=π .
2 4 8
1
By induction, it can be proven that .An+1 = ×An , that is, it is a geometric
2
1
sequence with common ratio . , so the general term of the sequence is
2
r2
.An = π .
2n
1.1. Sequences of Real Numbers 11
• Limits of Sequences
∀ L ∈ R+ ∃ p ∈ N : n > p ⇒ un > L.
.
Notes:
Theorem 1.1.2 Let u and v be sequences such that from a certain order,
un ≤ vn . Then,
a) un → +∞ ⇒ vn → +∞.
b) vn → −∞ ⇒ un → −∞.
n
1 1 1 1
u =
. n √ = 1 + √ + √ + ··· + √ .
k 2 3 n
k=1
1 1 1 1 √
1 + √ + √ + ··· + √ ≥ n × √ = n
.
2 3 n n
√
and n → +∞, we can state that un → +∞ (see Fig. 1.9).
1.1. Sequences of Real Numbers 13
1 1 1 1 1 1
(a) If ε = 80 then a − 80
< un < a + 80
(b) If ε = 500
then a − 500
< un < a + 500
whatever
whatever n > 4 n > 30
Figure 1.10: The value of p varies with the value of ε
14 Sequences of Real Numbers
5
The limit of this sequence is a = . To prove this fact using the definition,
9
we proceed as follows: Let ε > 0; we must show that there exists a natural
number, p, such that
5n 5
.if n > p then − < ε.
9n + 1 9
Note that
5n 5 −5 5
−
.
9n + 1 9 9(9n + 1) = 9(9n + 1) .
=
Solving inequality
5
. <ε
9(9n + 1)
5 − 9ε
for n, we obtain n > .
81ε
5 − 9ε
If we choose a natural number p greater than or equal to , then
81ε
5 − 9ε 5n 5
.n > p ⇒ n > ⇒ − < ε.
81ε 9n + 1 9
5 5
Note that if ε ≥ , we can choose p = 1. If ε < , we can choose
9 9
5 − 9ε 5 − 9ε
p = Int . This number p = Int is the smallest number
81ε 81ε
that makes the statement true.
1
Example 1.1.16 Let us prove that un = → 0. Given an arbitrary ε > 0,
n
we must show that there exists a natural number, p, such that
1 1
.if n > p then = < ε.
n n
1
Let p be a natural number greater than or equal to . Then, for n > p, we
ε
1 1 1
have < < ε. If we choose p = Int , this will be the smallest order
n p ε
from which the proposition is true for a given value of ε (see Fig. 1.11).
1
Similarly, we can prove that the sequence vn = α , α ∈ R+ , converges to
n
zero.
1.1. Sequences of Real Numbers 15
Notes:
The set R, with this order relation, is called extended real line.
We can extend the notion of neighborhood to R. Let ε > 0. If a ∈ R,
we call the ε neighborhood of a the set Vε (a) = ]a − ε, a + ε[ (which
coincides with the neighborhood in R). The set Vε (+∞) = 1ε , +∞
is called the ε neighborhood of +∞. The set Vε (−∞) = −∞, − 1ε
is called the ε neighborhood of −∞. With the definitions given pre-
viously, we can unify from a formal viewpoint Definitions 1.1.6 and
1.1.7: If a ∈ R,
and
.∃ q ∈ N : n > q ⇒ |un − b| < ε.
Let m > max{p, q}. Then,
|b − a| |b − um + um − a| |b − um | + |um − a|
ε = = ≤
.
3 3 3
ε+ε 2ε
< = <ε
3 3
and we conclude that ε < ε, which is impossible.
c) lim(un )p = (lim un )p , p ∈ N.
un lim un
d) lim = , if vn = 0, ∀n ∈ N, and lim vn = 0.
vn lim vn
e) lim(un )1/p = (lim un )1/p (if p is even, it must be un ≥ 0, ∀n ∈ N).
f) lim (un )vn = (lim un )lim vn , if un > 0, ∀n ∈ N, and the limits of
the sequences are not both zero.
h) (∃ p ∈ N ∀n ≥ p, un ≥ 0) ⇒ lim un ≥ 0.
i) (∃ p ∈ N ∀n ≥ p, un ≥ vn ) ⇒ lim un ≥ lim vn .
a) Let a and b be the limits of sequences (un ) and (vn ), respectively. Let
ε > 0 be arbitrary. By definition,
ε
.∃ p1 ∈ N : n > p1 ⇒ |un − a| <
2
and
ε
∃ p2 ∈ N : n > p2 ⇒ |vn − b| <
. .
2
Let p = max{p1 , p2 }; then, if n > p,
But ||un | − |a|| ≤ |un − a|, so ||un | − |a|| < ε, ∀n > p. We conclude that
|un | → |a|.
Notes:
−2 + 4 cos(n)
a =
. n .
n
We know that −1 ≤ cos(n) ≤ 1, ∀n ∈ N; therefore,
. − 6 ≤ −2 + 4 cos(n) ≤ 2, ∀n ∈ N.
u = cos(nπ) = (−1)n
. n
a − ε < un ≤ wn ≤ vn < a + ε.
.
−2 + 4 cos(n)
a =
. n
n
from the previous example, now applying Theorem 1.1.8.
We know that −1 ≤ cos(n) ≤ 1, ∀n ∈ N, which implies that
6 −2 + 4 cos(n) 2
. − ≤ ≤ , ∀n ∈ N.
n n n
1 −2 + 4 cos(n)
Considering that → 0, we conclude that lim = 0 (see
n n
Fig. 1.13).
20 Sequences of Real Numbers
we obtain
a1 ap−1 ap
a0 + + · · · + p−1 + p
P (vn ) vn vn vn
. = .
Q(vn ) b1 bp−1 bp
b0 + + · · · + p−1 + p
vn vn vn
1
By hypothesis, vn → +∞; then → 0 for all s ∈ N. We conclude by
vns
Theorem 1.1.5 that
P (vn ) a0
. → .
Q(vn ) b0
Let p = q. Putting vnp in evidence in the numerator and vnq in the denomi-
nator, we obtain the equalities
P (vn ) a0
If p > q, vnp−q → +∞, so the limit of will be +∞ if > 0 and
Q(vn ) b0
a0
−∞ if < 0.
b0
P (vn ) a0
If p < q, vnp−q → 0, so →0· = 0.
Q(vn ) b0
If a ≤ −1, the limit does not exist. If a < −1, |an | → +∞.
Proof: Let a ∈ R such that |a| > 1. We intend to prove that |a|n → +∞.
Consider an arbitrary L > 0. Since log(|a|) is positive when |a| > 1, we
have
log(L)
|a|n > L ⇔ en log(|a|) > elog(L) ⇔ n log(|a|) > log(L) ⇔ n >
. .
log(|a|)
log(ε)
|an | = |a|n < ε ⇔ n log(|a|) < log(ε) ⇔ n >
. .
log(|a|)
22 Sequences of Real Numbers
If a = 1, then an = 1, ∀n ∈ N, so an → 1.
n
4n
. .
2n + 1 n∈N
Given n ∈ N,
n n
4 4n 4n 4 4n
3n ≥ 2n + 1 ⇒
. = ≤ ⇒ ≤ .
3 3n 2n + 1 3 2n + 1
n
4
But by Theorem 1.1.10, lim = +∞. Considering the last inequality,
3
n
4n
we conclude by Theorem 1.1.2 that lim = +∞.
2n + 1
Example 1.1.20 Let us evaluate the limit of the sequence of general term
n
n
.(an )n = .
5n + 1
1
The sequence (an ) is monotonically increasing and has a limit of , so
5
n n n
1 n 1 1 n 1
. ≤ ≤ ⇒ ≤ ≤ .
6 5n + 1 5 6 5n + 1 5
n n
1 1
But by Theorem 1.1.10, lim = lim = 0. From the previous
5 6
n
n
inequalities, we conclude by Theorem 1.1.8 that lim = 0.
5n + 1
1.1. Sequences of Real Numbers 23
Theorem 1.1.11
u1 + · · · + un
a) Let a ∈ R. If un → a, then → a.
n
√
b) If b ∈ R+ , then n
b → 1.
un+1 √
c) Let c ∈ R. If un > 0, ∀n ∈ N, and → c, then n un → c.
un
that is,
u1 − a + u2 − a + · · · + un − a
.∃ s ∈ N : n > s ⇒ < ε.
n
We know that ∃ p ∈ N : n > p ⇒ |un − a| < 2ε . Additionally,
u1 − a + u2 − a + · · · + up − a
≤
n
.
|u1 − a| + |u2 − a| + · · · + |up − a| Mp
≤ ≤ ,
n n
where M = max{|u1 − a|, |u2 − a|, · · · , |up − a|}. Let q ∈ N such that
n > q ⇒ Mn p < 2ε . Let s = max{p, q}. If n > s, we have
u1 − a + u2 − a + · · · + un − a
≤
n
|u1 − a| + |u2 − a| + · · · + |up − a| |up+1 − a| + · · · + |un − a|
. ≤ +
n n
ε ε
(n − p)
≤ + 2 ≤ ε.
2 n
Thus, if a ∈ R, the statement is proved.
Now, we show the theorem for a = +∞. Let L be an arbitrary positive real
number. We know that there exists p ∈ N : n > p ⇒ un > 2 L. Given
that
u1 + u2 + · · · + up
. → 0,
n
then
u1 + u2 + · · · + up −L
.∃ q ∈ N : n > q ⇒ > .
n 2
24 Sequences of Real Numbers
Additionally,
up+1 + up+1 + · · · + un 2 L (n − p)
. > .
n n
As n−p
n → 1, ∃ r ∈ N : n > r ⇒ n−p 3
n > 4 . Let s = max{p, q, r}. If n > s,
we have
u1 + u2 + · · · + un −L 2 L (n − p) −L 3
. > + > + 2 L = L.
n 2 n 2 4
Thus, the statement is proved if a = +∞.
For a = −∞, the procedure is similar but with obvious adaptations.
√ √
b) Let b > 1. In this case n b > 1, so we can write n b = 1 + hn , where,
for each n, hn > 0. Then, by Newton’s Binomial,
n
n k
b = (1 + hn )n =
. h > 1 + n hn ,
k n
k=0
Applying roots,
up+1 ε √ up+1 ε
. n p+1 c− < n un < n p+1 c+ .
c − 2ε 2 c + 2ε 2
By item b),
up+1 up+1
. n p+1 → 1 and n p+1 → 1;
c − 2ε c + 2ε
therefore,
up+1 ε ε up+1 ε ε
. n p+1 c− →c− and n p+1 c+ →c+ .
c − 2ε 2 2 c + 2ε 2 2
Notes:
√
1. It is easy to see, using item c) of the previous theorem, that n
n → 1.
2. The converse of item c) of the previous theorem is not true, that is,
√ un+1
n
un → c does not imply that → c (un > 0, ∀n ∈ N). For
un
example, consider the sequence un = e−n−(−1) :
n
√ (−1)n
.
n
un = e−1− n → e−1 ,
and
un+1 e−3 , if n is odd
= e−1+2(−1) =
n
.
un e, if n is even
has no limit.
Example 1.1.21 Let us calculate the limit of the sequence with general
√
term n 2n + 3n . The sequence un = 2n + 3n is positive and
n+1
2
+1
un+1 2n+1 + 3n+1 3
. lim = lim = lim n = 3.
n→+∞ un n→+∞ 2n + 3n n→+∞ 2 1 1
· +
3 3 3
1.1. Sequences of Real Numbers 27
Proof: For example, suppose that the sequence (un ) is increasing. The set
of terms of the sequence is bounded; therefore, it has a supremum, which we
denote by S. Let ε > 0 be arbitrary. There is an element of the sequence,
up , such that S − ε < up ≤ S. However, the sequence is increasing;
therefore, up ≤ un , ∀n > p. We then have S − ε < un ≤ S, ∀n > p, so
un → S.
Note: The converse is not true; that is, there are convergent sequences that
1
are not monotonic. For example, the sequence un = (−1)n converges to
n
0 and is not monotonic (Fig. 1.14).
(−1)n
Figure 1.14: The sequence un =
n
28 Sequences of Real Numbers
n
1
• The Sequence 1+
n n∈N
n
1
Let us consider the sequence with general term .un = 1+ . According
n
to Newton’s Binomial, it can be written as follows:
1
n
n
n 1
p
un = 1+ =
n p=0
p n
0 1 2 n
n 1 n 1 n 1 n 1
= + + + ··· +
0 n 1 n 2 n n n
n! 1 n! 1 n! 1
= 1+ · + · 2+ · 3+
1!(n − 1)! n 2!(n − 2)! n 3!(n − 3)! n
n! 1 n! 1
+ ··· + · + ·
(n − 1)!(n − (n − 1))! nn−1 n!(n − n)! nn
1 n! 1 1 n! 1 1 n! 1
= 1+ · · + · · 2+ · · 3+
1! (n − 1)! n 2! (n − 2)! n 3! (n − 3)! n
.
1 n! 1 1 n! 1
+ ··· + · · + · ·
(n − 1)! (n − (n − 1))! nn−1 n! (n − n)! nn
1 1 n(n − 1) 1 n(n − 1)(n − 2)
= 1+ + · 2
+ · +
1! 2! n 3! n3
1 n! 1 n!
+ ··· + · + ·
(n − 1)! nn−1 n! nn
1 1 1 1 1 2
= 1+ + 1− + 1− 1− +
1! 2! n 3! n n
1 1 2 n−1
+ ··· + 1− 1− ··· 1 − .
n! n n n
p
Since .0 < 1 − < 1, for any .1 ≤ p ≤ n − 1, we can write
n
1 1 1
.2 ≤ un ≤ 1 + 1 + + + ··· + .
2 3! n!
It is easy to prove by induction that
1 1
. < n−1 , ∀n ≥ 3,
n! 2
which allows us to write, taking into account the formula for the sum of the
terms of a geometric sequence,
n
1 1 1 1 − 12 1
.2 ≤ un ≤ 1 + 1 + + 2 + · · · + n−1 = 1 + < 1+ = 3.
2 2 2 1 − 12 1 − 12
1.1. Sequences of Real Numbers 29
We have just proved that the sequence is bounded. Let us see that it is
monotonic:
u
. n+1 − un =
1 1 1 1 1 2
= 1+ + 1− + 1− 1− +
1! 2! n+1 3! n+1 n+1 Leonhard Euler (1707–
1 1 2 n−1 1783) was a Swiss mathe-
+ ··· + 1− 1− ··· 1 − + matician who studied phi-
n! n+1 n+1 n+1 losophy and mathemat-
1 1 2 n ics at the University of
. + 1− 1− ··· 1 − − Basel. He was a professor
(n + 1)! n+1 n+1 n+1
in St. Petersburg (Russia)
1 1 1 1 1 2 and Berlin. His contribu-
− 1+ + 1− + 1− 1−+ tions are important in al-
1! 2! n 3! n n
most all areas of math-
1 1 2 n−1 ematics, namely, number
+ ··· + 1− 1− ··· 1 − .
n! n n n theory, differential equa-
tions, analysis, calculus
of variations, and rational
By associating terms, we obtain mechanics, as well as in
other scientific areas, such
as calculation of planetary
1 1 1 orbits, artillery and bal-
un+1 − un = 1− − 1− +
2! n+1 n listics, shipbuilding, and
navigation. In 1771, he
1 1 2 1 2
+ 1− 1− − 1− 1− + became blind, but this did
3! n+1 n+1 n n not prevent him from con-
tinuing an extraordinary
1 1 2 n−1
. +··· + 1− 1− ··· 1 − − scientific production. Af-
n! n+1 n+1 n+1 ter his death, St. Peters-
burg Academy continued
1 2 n−1
− 1− 1− ··· 1 − + to publish his unpublished
n n n works for 50 years. Eu-
1 1 2 n ler introduced the nota-
+ 1− 1− ··· 1 − . tion .f (x) for a function,
(n + 1)! n+1 n+1 n+1 e for the base of natural
logarithms, i for the root
p p of .−1, .π, . for summa-
Since .1 − > 1− > 0 and the last term is positive, we have tion, .Δy for finite differ-
n+1 n
.un+1 − un > 0, for all .n ∈ N. The sequence
ences, and many others.
He was the most prolific
n mathematician of all time.
1
u =
. n 1+ (Source of image: Oil por-
n trait by Jakob Emanuel
Handmann (1718–1781),
is bounded and monotonic. By Theorem 1.1.12 we conclude that it is Deutsches Museum)
convergent. The limit of this sequence is an irrational number denoted by
30 Sequences of Real Numbers
the letter e: n
1
. lim 1 + = e.
n
This number is usually referred to as Neper’s number or Euler’s constant.
x n
Theorem 1.1.13 Let .x ∈ R. Then .lim 1 + = ex .
n
John Napier (or Jhone
Neper) (1550–1617) was
a Scottish mathematician.
He has dedicated himself Theorem 1.1.14 Let .x ∈ R and u be a sequence such that .lim un = +∞
to studies in theology and un
x
mathematics. He is es- or .lim un = −∞. Then .lim 1 + = ex .
sentially known as the in- un
ventor of logarithms, al-
though he also has works
on spherical triangles. He
invented a calculating rule, Example 1.1.22 We know that .xn = n + 5 → +∞. For every .n ∈ N,
known as “Napier’s bones”
because it was made of 2n+1
2n+1 2n+1 n+5 n+5
ivory, which allowed mul- n+3 2 2
. an = = 1− = 1− .
tiplication, division, and n+5 n+5 n+5
the mechanical calculation
of square and cubic roots.
Using the previous theorem, we obtain the following:
(Source of image: Engrav-
ing by Samuel Freeman n+5
2
(1773–1857)) . lim 1 − = e−2 .
n+5
2n + 1
Since . → 2, then
n+5
• Subsequences
Notes:
and
∃p2 ∈ N : zn > p2 ⇒ |uzn − a| < ε.
.
Proof: Let
M = {p ∈ N : up < un , ∀n > p}.1
.
If .M is infinite, that is, if there exists .p1 < p2 < · · · < pn < · · · belonging
to .M,
.up1 < up2 < up3 < · · · < upn < · · · ,
Proof: By Lemma 1.1.1, the sequence .(un ) admits at least one monotonic
subsequence .(upn ). Because .(upn ) is monotonic and bounded, then by
Theorem 1.1.12 it is convergent.
1
Example 1.1.24 Consider the sequence .un = (−1)n + . The sublimits
n
1
of .(un ) are .−1 and 1 because 1 is the limit of subsequence .u2n = 1 +
2n
1
and .−1 is the limit of subsequence .u2n−1 = −1 + (see Fig. 1.15).
2n − 1
1
Figure 1.15: Sublimits of sequence .un = (−1)n +
n
34 Sequences of Real Numbers
3. If u is convergent, then S is a unit set, that is, it has only one element.
1, 1, 2, 1, 2, 3, 1, 2, 3, 4, 1, 2, 3, 4, 5, . . . ,
.
then .S = N.
β − 1/4 < a2 ≤ β;
.
and then,
1 1 1 1
β−
. < a2 − < u p2 < a2 + < β + .
2 4 4 2
1.1. Sequences of Real Numbers 35
Repeating this reasoning, for each .n ∈ N, there exists .upn such that
1 1
β−
. < u pn < β + .
n n
We have constructed a subsequence .(upn ) that converges to .β, so .β ∈ S.
Similarly, it can be shown that .α ∈ S.
Thus, there exists .n1 such that .|un1 − a| ≥ ε; there exists .n2 > n1 such
that .|un2 − a| ≥ ε; and there exists .nm > nm−1 such that .|unm − a| ≥ ε.
Because it is bounded, the subsequence .(unm ) has a convergent subsequence
.(unm ): .unm → b = a, since .|unm − a| ≥ ε for all terms. Therefore, we
s s s
have shown that S has an element b different from a, which is incompatible
with the hypothesis of .lim un = lim un = a.
1
Example 1.1.25 We prove that the sequence .un = is a Cauchy se-
n
quence. Let .m, n > p; then
1 1 1 1 1 1 2
.
n m ≤ n + m < p + p = p.
−
so
|un | < |up+1 | + 1, ∀n > p.
.
1.1. Sequences of Real Numbers 37
.|un | ≤ M, ∀n ∈ N.
a) By induction prove that 0 < un < 2, ∀n ∈ N. 12. Consider the sequence of real numbers defined re-
cursively
b) Prove that the sequence is increasing.
⎧
c) Prove that the sequence is convergent. ⎨ x1 = 2
. x 1
d) Evaluate the limit of the sequence. ⎩xn+1 = n + , ∀n ∈ N.
2 xn
9. Consider the sequence
a) Prove that xn > 0, ∀n ∈ N.
⎧ √
⎨a1 = √2 b) Prove that xn > 2, ∀n ∈ N.
⎩an+1 = √2 an ,
.
∀n ∈ N. c) Prove that (xn ) is monotonic.
d) Show that (xn ) converges.
a) By induction show that e) Evaluate the limit of (xn ).
√
. 2 ≤ an < 2, ∀n ∈ N. 13. Consider the sequence of real numbers defined re-
cursively
b) By induction show that the sequence (an ) is ⎧
⎪
⎨ x1 = 3
increasing.
.
⎪ x2 + 3
c) Show that there exists a ≤ 2 such that an → ⎩xn+1 = n , ∀n ∈ N.
2 xn
a.
10. Let a ∈ R be a positive number. Consider the se- a) Show by induction that
quence of real numbers defined recursively √
.xn − 3 > 0, ∀n ∈ N.
⎧
⎪
⎨x1 = a b) Show that the sequence (xn ) is decreasing.
. xn
⎪
⎩xn+1 = , ∀n ∈ N. c) Show that the sequence (xn ) converges.
2 + xn
d) Find the limit of the sequence (xn ).
40 Sequences of Real Numbers
14. Prove by definition the following limits: The middle segment of each of the four segments is
25n 25 removed and replaced by two segments of the same
a) lim = length. The resulting curve has a length of 16 .
3n + 1 3 9
√
n
b) lim √ =1
n+1
√
2 n + 5−n
c) lim √ = 2
n+1
sin(n)
d) lim 1+ =1
n
e) lim n3 = +∞
n2
f) lim = +∞
n+1 Continuing this process ad infinitum, the result is the
15. The Koch curve is created from a line segment of Koch curve.
length 1. This segment is divided into three equal
parts. We then remove the middle segment. Next,
we replace the removed segment with two segments
of length 13 . The resulting shape can be seen in the
figure below.
SOLUTIONS
n ∞
1. a) Let an = . It is not possible to directly apply Theorem 1.1.5, as the indeterminate form ∞
occurs.
n2 + 3
To solve this, we can rewrite an as follows:
1 1
n2 ·
n n n
.an = = = .
n2 + 3 3 3
n2 1 + 2 1+ 2
n n
1 3
We know, from Example 1.1.16, that is a null sequence, which by Theorem 1.1.5, implies that 2 → 0.
n n
Again by Theorem 1.1.5,
. lim an = 0.
√
n
Since lim n = 1 (see Note 1 after Theorem 1.1.11), we can conclude that
n √
n
. lim + n = 1.
n2 + 3
Note: The procedure described here is equivalent to dividing both terms of the fraction by the power of n
with the highest exponent that appears in the expression. Alternatively, we could use Theorem 1.1.9, with
vn = n, keeping in mind that p = 1 and q = 2.
n+1
b) The following sequence is given by an = . Taking into account the note from the previous item, we
2n + 5
divide both members of the fraction that defines the sequence (an ) by n, which is the highest power in the
expression.
n+1 1
1+
n+1 n n
.an = = = .
2n + 5 2n + 5 5
2+
n n
1
Since is a null sequence, we can apply Theorem 1.1.5 and conclude that
n
1
. lim an = .
2
n3 + 3n2 + 5
c) Let an = . Since n3 is the highest power of n that appears in the fraction that defines the
n2 + 2
sequence (an ), we can divide the numerator and denominator by n3 :
n3 + 3n2 + 5 3 5
1+ + 3
n3 +3n2 +5 n3 n n
.an = = = .
n2 + 2 n2 + 2 1 2
+
n3 n n3
Using Theorem 1.1.5, we can conclude that
. lim an = +∞,
1
because is a null sequence and an > 0, ∀n ∈ N.
n
42 Sequences of Real Numbers
35n7 + 1
d) Let an = . If we divide both the numerator and denominator of the fraction that defines
71n7 + 5n6 + 1
the sequence (an ) by n raised to the highest power, n7 , we obtain
35n7 + 1 1
35 + 7
35n7+1 n7 n
.an = = = .
71n7 + 5n6 + 1 71n7 + 5n6 + 1 5 1
71 + + 7
n7 n n
1
Knowing that is a null sequence, we can use Theorem 1.1.5, to conclude that
n
35
. lim an = .
71
√
3
n2 + n + n
e) Given an = √
4 √ , we can simplify the expression by dividing both the numerator and denomi-
2n4 + 1 + n
nator by n as shown below:
√
3 2
n2 + n + n 3 n +n 3 1 1
√ +1 + 2 +1
3
n2 + n + n n n3 n n
.an = √4 √ = √ 4 √ = = .
2n4 + 1 + n 2n4 + 1 + n 4
4 2n + 1 n 4 1 1
n + 2+ 4 +
n4 n2 n n
By applying Theorem 1.1.5 to the above expression, we conclude that the limit is given by
√
2
. lim an = − .
2
√ √
1+2 n n+1
g) Let an = √ . To simplify the expression of (an ), we can divide both the numerator and
n+ 3n
denominator of the fraction that defines it by n. This gives us
√ √ √ √
1+2 n n+1 1+2 n n+1 1 1
√ √ √ √ √ +2 1+
1+2 n n+1 n n n n n
.an = √ = √ = = .
n+ 3n n+ 3n n 3 1
1+ 3 3 1+
n n n2
By applying Theorem 1.1.5, we can determine that
. lim an = 2.
1.2. Solved Exercises 43
√
3
1 − 27n3
h) Let an = . To find the limit of this sequence, we will manipulate the expression for an to
1 + 4n
simplify it.
First, we will divide both the numerator and denominator of the fraction that defines the sequence (an ) by
n, and next we simplify the expression. This gives us
√
1 − 27n3 3 1 − 27n
3 3
1
√ 3
− 27
3
1 − 27n3 n n3 n3
.an = = = = .
1 + 4n 1 + 4n 1 1
+4 +4
n n n
Therefore, by Theorem 1.1.5,
3
=− .
. lim an
4
√
n (−1)n + n
i) Consider the sequence (an ) given by an = √ . In this case, by dividing both the numerator
2 + n3 + 1
and denominator of the fraction that defines the sequence (an ) by n3/2 , we can simplify the expression as
follows:
√ √
n (−1)n + n (−1)n + n (−1)n
√ √ +1
n (−1)n + n n 3/2 n 1/2 n
.an = √ = √ = = .
2 + n3 + 1 2 + n3 + 1 2 n3 + 1 2 1
√ + √ + 1 +
n3/2 n3 n3 n3
n3
1
Since (−1)n is bounded and √ is a null sequence, we can apply Theorem 1.1.6 to conclude that
n
(−1) n
√ → 0. Moreover, applying Theorem 1.1.5, we can compute the limit of the sequence and obtain
n
. lim an = 1.
√
3
n n2 + 2
j) Let an = . By dividing both the numerator and the denominator of the fraction that defines
n2 + (−1)n n
2
sequence (an ) by n , we obtain
√ √
n 3 n2 + 2 3
n2 + 2 2
3 n +2 3 1 2
√
3 + 3
n n2 + 2 n2 n n3 n n
.an = = 2 = = = .
n2 + (−1)n n n + (−1)n n (−1)n (−1)n (−1)n
2
1+ 1+ 1+
n n n n
. lim an = 0.
2 n e1/n 2n
k) Let an = √ = √ · e1/n = bn · e1/n . By dividing both the numerator and
(−1)n+ n2 + 5 (−1)n + n2 + 5
denominator of the fraction that defines the sequence (bn ) by n, we obtain
2n 2 2 2
.bn = √ = √ = = .
(−1)n + n2 + 5 (−1)n + n2 + 5 (−1) n 2
n +5 (−1) n 5
+ + 1 +
n n n2 n n2
44 Sequences of Real Numbers
Hence,
. lim bn = 2.
1/n
Since lim e = 1, we can conclude that
2 n e1/n
. lim √ = 2.
(−1)n + n2 + 5
n n n2 1/n
n2 − 1 1 1
.an = = 1− = 1− , ∀n ∈ N.
n2 n2 n2
1
Since lim = 0, we can conclude by Theorem 1.1.5 that
n
. lim an = (e−1 )0 = 1.
⎡ ⎤2 n ⎡ 22n ⎤1/2n
5 5
2n 1− n
4n ⎢ 1− ⎥
4n − 5 ⎢ ⎥
=⎢
4
⎥ =⎢ ⎢
4n ⎥
, ∀n ∈ N.
.an =
4n + 3 ⎣ 3 ⎦ ⎣ 22n ⎥
⎦
4n 1 + n 3
4 1+
4n
and
22n n
3 3 4
. lim 1+ = lim 1 + n = e3 .
4n 4
1
But by Theorem 1.1.10, lim = 0; therefore,
2n
0
e−5
. lim an = = 1.
e3
and
4 n
. lim 1+ = e4 ;
n
therefore,
e2
. lim an = · 1 = e−2 .
e4
d) To find the limit of the sequence (an ), we can simplify the expression defining it by putting n in evidence
in the numerator of the expression that defines (an ) and 5n in the denominator. This gives us the following
expression:
⎡ ⎤n
2 2 n
n 1+ n 1+
⎢ n ⎥ 1 n
.an = ⎢ ⎥ = · , ∀n ∈ N.
⎣ 1 ⎦ 5 1 n
5n 1 + 1+
n n
n
1
We know from Theorem 1.1.10 that lim = 0. Additionally, by Theorem 1.1.13, we can find that
5
2 n
. lim 1 + = e2
n
and n
1
. lim 1+ = e.
n
Therefore,
2 n
n 1+
1 n e2
. lim an = lim · lim =0· = 0.
5 1 n e
1+
n
e) If we consider an as the general term of the sequence, we can write it as follows:
⎡ ⎤n ⎡ ⎤n
1 1 1/3 n
n 3n 1 + 1+ 1+
3n + 1 ⎢ 3n ⎥ ⎢ 3n ⎥ n
.an = =⎢
⎣ ⎥⎦ =⎢
⎣
⎥ = , ∀n ∈ N.
3n + 2 2 2 ⎦ 2/3 n
3n 1 + 1+ 1+
3n 3n n
and n
2/3
. lim 1+ = e2/3 .
n
Therefore,
e1/3
. lim an = = e−1/3 .
e2/3
46 Sequences of Real Numbers
f) We have
1 4n 1 n 4
4n−2 3−2− 1−
2n + 1 n n
.an = 3− = = , ∀n ∈ N.
n 1 2 1 2
3−2− 1−
n n
By Theorem 1.1.13, we know that n
1
. lim 1− = e−1 .
n
Thus, we can conclude that 4
. lim an = e−1 = e−4 .
g) We have
and 2n
1
. lim 1+ = e.
2n
Therefore, we can find that
1/2
e5
. lim an = = e2 .
e
π
h) Knowing that lim arctan(x) = , we can conclude1 that lim earctan(n) = eπ/2 . Now, let us consider
2 x→+∞
n 2
n +3 2
an = . Putting n in evidence in this expression yields
2n2 + 1
⎡ ⎤n 1/n
3 3 n 3 n2
n n2 1+ n 1 + n 1+ 2
n2 +3 ⎢ n 2 ⎥ 1 n 2 1 n
.an = =⎢
⎣ ⎥
⎦ = 2 · n = · 1/2n .
2n2 + 1 1 1 2 1 2n2
2n2 1 + 1 + 1+
2n2 2n2 2n2
1 We are using the following result: If lim f (x) = L and f (n) = un , n ∈ N, then lim un = L.
x→+∞
1.2. Solved Exercises 47
n
1 1 1
Therefore, as lim = lim = lim = 0, we can conclude that
n 2n 2
(e3 )0
. lim an =0· = 0.
e0
Thus, we get n
n2 + 3
. lim · earctan(n) = 0.
2n2 + 1
3n sin 23n + 1 3n
3. a) Let an = = · sin 23n + 1 . We know that
23n + 1 23n + 1
. |sin(x)| ≤ 1, ∀x ∈ R;
% 3n %
% %
therefore, sin 2 + 1 ≤ 1, ∀n ∈ N, that is, the sequence with general term sin 23n + 1 is a bounded
3n
sequence. We prove that the sequence with general term 3n is a null sequence.
2 +1
n
3
3n 3n
8
. lim = lim n = lim n = 0.
23n + 1 8 +1 1
1+
8
We conclude that the given sequence is a null sequence because it is the product of a null sequence and a
bounded sequence (see Theorem 1.1.6).
1
b) Let an = · cos(n + 1) · log(n). We know that −1 ≤ cos(x) ≤ 1, ∀x ∈ R. Then
n
. − 1 ≤ cos(n + 1) ≤ 1, ∀n ∈ N.
Therefore,
√ log(n) 1 log(n) √
.
n
− log
n =− ≤ · cos(n + 1) · log(n) ≤ = log n n , ∀n ∈ N.
n n n
√
Because we know that lim n n = 1, we have
√
n
. lim log n = 0.
It follows from the Squeeze Theorem that
. lim an = 0.
n2 + 3
c) Let an = √ · cos n3 + 2 . For all n ∈ N, we have
n n3 + 2
% %%
%
.0 ≤ %cos n3 + 2 % ≤ 1.
n2 + 3
Let us consider the sequence of general term √ . By dividing the numerator and denominator of
n n3 + 2
the fraction that defines this sequence by the highest power of n, we obtain
n2 + 3 n2 + 3 (n2 + 3)2 1 6 9
+ 3 + 5
n 5/2 n 5/2 n 5 n n n
. lim √ = lim √ = lim = lim = 0.
n n3 + 2 n3 + 2 n3 + 2 2
1+ 3
n5/2 n3/2 n3 n
We conclude that the sequence (an ) is a null sequence because it is the product of a null sequence and a
bounded sequence (see Theorem 1.1.6).
48 Sequences of Real Numbers
1 √
n n n! n!
d) Let an = n! = . Let bn = n . It is evident that bn > 0, ∀n ∈ N.
n nn n
(n + 1)!
n
bn+1 (n + 1)n+1 (n + 1)! nn (n + 1) nn n 1
. lim = lim = lim = lim = lim = .
bn n! (n + 1)n+1 n! (n + 1)n+1 n+1 e
nn
1
By Theorem 1.1.11 we can conclude that lim an = .
e
√ n4
n4
and bn = n2 e−n . It is evident that bn > 0, ∀n ∈ N.
n
e) Let an = n2 e−n − 4
n +1
(n + 1)2 2
bn+1 2 n
. lim = lim en+1 = lim (n + 1) e = 1 lim n + 1 1
= .
bn n2 en+1 n2 e n e
en
√
n 1
By Theorem 1.1.11 we can conclude that lim n2 e−n = . Using Theorem 1.1.14 we obtain the following
e
result:
n4
n4 1 1 1
. lim = lim
4
n +1 n4 = lim n4 = e .
n4+1 1
1+ 4
n4 n
1 1
We conclude that lim an =− = 0.
e e
n
n sin(n) 1 n
f) Let an = √ = ·√ · sin(n). We know that
2n 5n3 + 1 2 5n3 + 1
. |sin(n)| ≤ 1, ∀n ∈ N;
n
therefore, the sequence sin(n) is a bounded sequence. We show that the sequence √ is a
5n3 + 1
null sequence.
n 1 1
√ √
n n 3/2 n n
. lim √ = lim √ = lim = lim = 0.
5n3 + 1 5n3 + 1 5n3 + 1 1
5 +
n3/2 n3 n3
n
1
Since lim = 0, we have
2
n
1 n
. lim ·√ = 0.
2 5n3 + 1
We conclude that the given sequence is a null sequence because it is the product of a null sequence by a
bounded sequence (see Theorem 1.1.6).
1.2. Solved Exercises 49
g) It is not possible to directly apply Theorem 1.1.5, as the indeterminate form ∞ − ∞ occurs. Dividing and
multiplying the general term of the sequence by its conjugate, we obtain
√ √
n2 + 2n − n n2 + 2n + n n2 + 2n − n2
lim n2 + 2n − n = lim √ = lim √
2
n + 2n + n n2 + 2n + n
2n
.
2n n 2 2
= lim √ = lim √ = lim = lim = 1.
n2 + 2n + n n2 + 2n + n 2
n + 2n 2
+1 1 + + 1
n n2 n
(n + 1)n+2 n 1
h) Let an = − · sin . Then
(n + 2)n+1 3 n
(n + 1)n+2 n 1 n + 1 n+2 n 1
lim n+1
− · sin = lim · (n + 2) − · sin
(n + 2) 3 n n+2 3 n
n+2
1 1 n 1
= lim 1− · (n + 2) · sin − · sin
n+2 n 3 n
n+2 n+2
. 1 1 1 1 n 1
= lim 1− · n · sin + 1− · 2 · sin − · sin
n+2 n n+2 n 3 n
⎛ ⎞
1 1
n+2 sin n+2 sin
⎜ 1 n 1 1 1 n ⎟
⎜
= lim ⎝ 1 − · + 1− · 2 · sin − · ⎟.
n+2 1 n+2 n 3 1 ⎠
n n
n+2
1 1 1
By Theorem 1.1.14, lim 1 − = e−1 . Since lim = 0, we conclude that lim sin = 0
n+2 n n
1
sin
n 1
and lim = 1. Using Theorem 1.1.5, we arrive at the following result: lim an = e−1 − .
1 3
n
i) By dividing both the numerator and denominator of the fraction that defines the general term of the sequence
by the power of the largest base and exponent, we obtain, by Theorem 1.1.10,
n n−1
3n − 5 3 1
−
3 −5
n
5n 5 5
. lim = lim n = lim = 0.
5n + 3 5 +3 3
1+ n
5n 5
4. In this exercise, we apply the Squeeze Theorem. When dealing with sequences defined by summations, a useful
technique is to first identify the largest and smallest terms in the summation. Then, these terms can be used to
bound the sequence.
a) The given sequence has a general term an which is defined as the sum from k = 1 to k = n − 1 of
sin2 (n) sin2 (n)
2 2
. As we show below, the largest term is 2 (which corresponds to k = 1), and the smallest
n + 3k n +3
2
sin (n)
is 2 (corresponding to k = n − 1). Clearly,
n + 3(n − 1)2
50 Sequences of Real Numbers
.n
2
+ 3k2 > n2 , ∀ k, n ∈ N
and
.n
2
+ 3k2 ≤ n2 + 3(n − 1)2 , ∀ k, n ∈ N such that k ≤ n − 1.
Therefore, for all n and 1 ≤ k ≤ n − 1, we obtain
1 1 1
n2 < n2 + 3k2 ≤ n2 + 3(n − 1)2 ⇒ > 2 ≥ 2
n2 n + 3k2 n + 3(n − 1)2
.
sin2 (n) sin2 (n) sin2 (n)
⇒ > 2 ≥ 2 .
n 2 n + 3k 2 n + 3(n − 1)2
sin2 (n)
n−1
n−1 1 1
⇔ · sin2 (n) ≤ < − 2 · sin2 (n), ∀n ∈ N.
n2 + 3(n − 1) 2
k=1
2
n + 3k 2 n n
n−1 n−1
Let bn = = . Dividing the numerator and denominator of the fraction that
n2 + 3(n − 1)2 4n2 − 6n + 3
2
defines the sequence by n , we get
1 1 1 1
− 2 − 2
n−1 n n n n
. lim = lim = lim = 0.
4n2 − 6n + 3 4n2 − 6n + 3 6 3
4 − +
n2 n n2
1 1
Let cn = − 2 . It is evident that lim cn = 0.
n n
As we know that | sin2 (n)| ≤ 1, ∀n ∈ N, the sequence sin2 (n) is bounded. Since the product of a null
sequence by a bounded sequence is a null sequence, we can assert that the sequences of general terms
n−1
. · sin2 (n)
n2 + 3(n − 1)2
and
1 1
. − 2 · sin2 (n)
n n
are null sequences. Finally, as the two limits are equal, the Squeeze Theorem allows us to conclude that
. lim an = 0.
5n
b) The general term an of the sequence is defined as the sum from k = 1 to k = n of √ . The largest
n4 + k
5n 5n
term is √ (which corresponds to k = 1), and the smallest is √ (corresponding to k = n), as
n4 + 1 n4 + n
shown below. We have
. n4 + k > n , ∀ k, n ∈ N,
2
1.2. Solved Exercises 51
and
. n4 + k ≤ n4 + n, ∀ k, n ∈ N such that k ≤ n.
5n
n
5n 5n
n· √ ≤ √ <n· , ∀n ∈ N
n4 +n k=1
n4 + k n2
.
5n2
n
5n 5n2
⇔ √ ≤ √ < = 5, ∀n ∈ N.
n4 +n k=1
n4 +k n2
5n2 n4
Let bn = √ = 5· . Dividing the numerator and denominator of the radicand of this
n4 + n n4 + n
4
sequence by n , we obtain &
' 1
. lim bn = 5 · lim '
( = 5.
1
1+
n3
Finally, as the two limits are equal, the Squeeze Theorem allows us to conclude that
. lim an = 5.
√
3
2n
c) The general term an of the sequence is defined as the sum from k = 1 to k = n of √3
. The largest
n 4+k
√
3
√
3
2n 2n
term is √
3
(which corresponds to k = 1), and the smallest is √ 3
(corresponding to k = n),
4
n +1 4
n +n
as shown below. We have √ √
3 3
. n4 + k > n4 , ∀ k, n ∈ N,
and √ √
3 3
. n4 + k ≤ n4 + n, ∀ k, n ∈ N such that k ≤ n.
Therefore, for all n and 1 ≤ k ≤ n, we obtain
√
3
√
3
√
3 1 1 1
n4 < n4 + k ≤ n4 + n ⇒ √
3
> √
3
≥ √
3
n4 n4 + k n4 + n
. √
3
√
3
√
3
2n 2n 2n
⇒ √
3
> √
3
≥ √
3
.
n4 n4 + k n4 + n
Because (an ) is defined as the sum of n terms, we obtain
√ √ √
3
2n
n 3 3
2n 2n
n· √
3
≤ <n· √ √
3 3
, ∀n ∈ N
n4 + n k=1 n 4+k n4
. √ √ √
3
2n4 n 3
2n
3
2n4 √
3
⇔ √
3
≤ √3
< √3
= 2, ∀n ∈ N.
n4 + n k=1 n 4+k n 4
52 Sequences of Real Numbers
Dividing the numerator and denominator of the fraction that defines the sequence on the left side of the
inequality by n4/3 , we have
√
3
√ 2n4
√ √ √
3
2n4 3
n4
3
2 3
2 √
3
. lim √ = lim √ = lim n = lim = 2.
3
n4 + n
3
n4 + n 1+ √ 1
√ 3 1+ √
3 n4 3
n
n4
Finally, as the two limits are equal, the Squeeze Theorem allows us to conclude that
√
3
. lim an = 2.
1
5. Let an = n + − n2 + 1. We can simplify this expression as follows:
2n
√ √
1 1 (n − n2 + 1)(n + n2 + 1) 1 n2 − (n2 + 1) 1 1
n+ − n2 + 1 = + √ = + √ = − √
2n 2n n+ n +12 2n n + n2 + 1 2n n + n2 + 1
. √ √ √ √
n + n2 + 1 − 2n n2 + 1 − n ( n2 + 1 − n)( n2 + 1 + n) 1
= √ = √ = √ = √ .
2n(n + n2 + 1) 2n(n + n2 + 1) 2n(n + n2 + 1)2 2n(n + n2 + 1)2
6. a) We will use the fact that the cosine function is bounded. This means that
. |cos(n)| ≤ 1, ∀n ∈ N,
In addition, √ √ √ √
√ √ 2n + 1 − 2n 2n + 1 + 2n
lim 2n + 1 − 2n = lim √ √
2n + 1 + 2n
.
2n + 1 − 2n 1
= lim √ √ = lim √ √ = 0.
2n + 1 + 2n 2n + 1 + 2n
As a result of the above, we can conclude that (an ) is a null sequence since it is the product of a bounded
sequence and a null sequence (see Theorem 1.1.6).
The sequence (an ) has three sublimits, namely, −1, 0, and 1, because it has subsequences that converge
π
to these real numbers. The sequence cn = arctan(n) has limit . Therefore, we can conclude that the
2
nπ π π
sublimits of the sequence bn = sin · arctan(n) are − , 0, and .
2 2 2
1.2. Solved Exercises 53
n (−1)n n
7. a) Let un = 1+2 . The even-indexed subsequence of (un ) is given by
2k (−1)2k 2k 2k 2k
.u2k = 1+2 = 1+2 , k ∈ N.
√
Let us consider the sequence an = n
1 + 2n . By Theorem 1.1.11, as 1 + 2n > 0, ∀n ∈ N, we can calculate
the limit of (an ) by
1 + 2n+1 1
1 + n+1
1+ 2n+1 2n+1 2
. lim = lim = lim = 2.
1 + 2n 1 + 2n 1 1
+
2n+1 2 2n+1
The limit of the sequence (an ) is 2; therefore, all its subsequences have this limit. In particular, the even-
indexed subsequence has limit 2. However, this subsequence is equal to the sequence (u2k ). We can affirm
that lim u2k = 2.
b) The odd-indexed subsequence of (un ) is given by
2k+1 (−1)2k+1 (2k+1) 2k+1 −(2k+1) 1
, k ∈ N,
2k+1
.u2k+1 = 1+2 = 1+2 = 1+
22k+1
8. a) We will use the Principle of Mathematical Induction to demonstrate that the sequence (un ) satisfies
.0 < un < 2, ∀n ∈ N.
.un+1 − un > 0, ∀n ∈ N.
because in item a), we have already established that un > 0 and 2 − un > 0. Therefore, the sequence is
increasing.
c) Since the sequence is bounded (as shown in item a)) and increasing (as shown in item b)), we can conclude
that the sequence (un ) is convergent because by Theorem 1.1.12, every bounded monotonic sequence is
convergent.
d) Let b ∈ R be the limit of the sequence. Since every subsequence of a convergent sequence is convergent to
the same limit, we have
. lim un+1 = b.
Hence,
√ √
.b = lim un+1 = lim 2 un = 2 b.
√
Therefore, b satisfies equation b = 2 b, which means that b2 − 2b = 0. Thus, b(b − 2) = 0, and we can
exclude the solution b = 0 because by item b), we have
√
.un ≥ u1 = 2, ∀n ∈ N.
√ x √ √
9. a) Let f be the function defined by f (x) = 2 = ex·log( 2) . Since 2 > 1, f is continuous and increasing
on R.
√ √
For n = 1, the formula is trivial: 2 ≤ a1 = 2 < 2.
√
Induction hypothesis: 2 ≤ an < 2.
√
Induction thesis: 2 ≤ an+1 < 2.
Proof: Assuming that the property is valid for n (induction hypothesis) and using the fact that f is an
increasing function, we have
√ √ √2 √
. 2 ≤ an < 2 ⇒ 2 = f 2 ≤ f (an ) = an+1 < f (2) = 2.
which implies that the property is valid for the order n + 1. Therefore, the Principle of Mathematical
Induction is satisfied, and we conclude that
√
. 2 ≤ an < 2, ∀n ∈ N.
1.2. Solved Exercises 55
√ √ √2
. 2 = a1 < a2 = 2 .
Proof: Assuming that the property is valid for n ∈ N (induction hypothesis), the validity of the property for
n + 1 is a direct consequence of the fact that f is increasing:
xn + x2n
Since by item a), xn > 0 for all n ∈ N, we have xn + x2n > 0, which implies that − < 0. Therefore,
2 + xn
xn+1 − xn < 0.
c) Since for all n ∈ N we have xn > 0 (item a)) and (xn ) is a decreasing sequence (item b)), we have
0 < xn ≤ x1 , that is, 0 < xn ≤ a, for all n ∈ N. In other words, the sequence of general term xn is
bounded. Because every monotone and bounded sequence is convergent, (xn ) is also convergent.
Let us assume that lim xn = b. Note that b ≥ 0 because xn > 0 for all n ∈ N.
xn
If the sequence converges to b, we also have lim xn+1 = b. Furthermore, the sequence is convergent,
2 + xn
since it is the quotient of two convergent sequences where the denominator never becomes zero and has a
limit different from zero.
56 Sequences of Real Numbers
Thus, we have
xn lim xn b
lim xn+1 = lim ⇔ lim xn+1 = ⇔b=
2 + xn lim(2 + xn ) 2+b
. b 2b + b2 − b b2 + b
⇔ b− =0⇔ =0⇔ = 0 ⇔ b2 + b = 0 (b = −2)
2+b 2+b 2+b
⇔ b(b + 1) = 0 ⇔ b = 0 ∨ b = −1.
11. a) We want to show that xn+1 − xn ≥ 0 for any n ∈ N0 . Now, if n = 0, we have x1 − x0 = a > 0. If n ≥ 1,
then
.xn+1 − xn = xn + x2n−1 − xn = x2n−1 ≥ 0.
Proof: Using the given formula xn+1 = x2n + x1 , we can see that x2n and x1 are both greater than zero
n n
since xn is positive by the induction hypothesis. Therefore, the sum of these two quantities, which is xn+1 ,
is also greater than zero.
c) Let us analyze the difference xn+1 − xn to determine whether the sequence is increasing or decreasing:
xn 1 xn 1 −x2n + 2
.xn+1 − xn = + − xn = − + = .
2 xn 2 xn 2xn
√
By item b), xn > 2, ∀n ∈ N; therefore, −x2n + 2 < 0, ∀n ∈ N. Then xn+1 − xn < 0, ∀n ∈ N, thus
proving that the sequence is decreasing.
first term is the maximum of the set of its terms; therefore, x1 = 2 ≥ xn ,
d) If a sequence is decreasing, its √
∀n ∈ N, and (xn ) is bounded: 2 < xn ≤ 2, ∀n ∈ N. Thus, we can conclude that (xn ) is convergent as it
is monotonic and bounded.
e) Let a = lim xn . All subsequences of (xn ) have limit a and
xn 1 a 1 a2 + 2
.a = lim xn+1 = lim + = + = .
2 xn 2 a 2a
a2 + 2 √ √ √
Solving the equation a = , we get a = − 2 or a = 2. As xn > 2, ∀n ∈ N, we conclude that
√ 2a
a = 2.
√
13. a) We will show by induction that xn − 3 > 0, ∀n ∈ N.
√
If n = 1, then x1 = 3 > 3, and the proposition is satisfied.
√
Induction hypothesis: xn − 3 > 0.
√
Induction thesis: xn+1 − 3 > 0.
√
√ x2n + 3 √ (xn − 3)2 √
Proof: By definition, xn+1 − 3 = − 3 = . We know that xn > 3; therefore,
√ 2 xn 2xn
xn+1 − 3 > 0.
√
Then, using the Principle of Induction, we proved that xn − 3 > 0, ∀n ∈ N.
b) We intend to show that the sequence is decreasing, that is, xn+1 − xn ≤ 0, ∀n ∈ N:
x2n + 3 −x2n + 3
.xn+1 − xn = − xn = .
2 xn 2xn
√
By item a), xn > 3, ∀n ∈ N; therefore, −x2n + 3 < 0, ∀n ∈ N. Then xn+1 − xn < 0, ∀n ∈ N, that is to
say, the sequence is decreasing.
c) If a sequence is decreasing, its first term is the maximum
√ of the set of terms of the sequence; therefore,
x1 = 3 ≥ xn , ∀n ∈ N. We have (xn ) bounded: 3 < xn ≤ 3, ∀n ∈ N. We can conclude that (xn ) is
convergent because it is monotonic and bounded.
d) Let a = lim xn . All subsequences of (xn ) have limit a and
x2n + 3 a2 + 3
.a = lim xn+1 = lim = .
2 xn 2a
a2 + 3 √ √ √
Solving the equation a = , we get a = − 3 or a = 3. As xn > 3, ∀n ∈ N, we conclude that
√ 2a
a = 3.
58 Sequences of Real Numbers
25n 25
14. a) In order to prove by definition that lim = , we must prove that the proposition
3n + 1 3
% %
% 25n 25 %%
.∀ε > 0 ∃ p ∈ N : n > p ⇒ %% − <ε
3n + 1 3 %
is true, that is, we need to show that for each positive real ε there is at least one order after which, for every
natural n, % %
% 25n 25 %%
. %% − < ε.
3n + 1 3 %
Let ε > 0. Solving the inequality for n, we obtain
% % % %
% 25n 25 %% % 75n − 75n − 25 % 25 25 − 3ε
. %% − % < ε ⇔ %% %<ε⇔
% <ε⇔n> .
3n + 1 3 3(3n + 1) 3(3n + 1) 9ε
25 − 3ε 25 − 3ε
If we choose p ∈ N such that p ≥ , then, as n > p, we get n > , that is,
9ε 9ε
% %
% 25n 25 %%
. %% − < ε.
3n + 1 3 %
Suppose, for instance, that we want to calculate the smallest order p from which
% %
% 25n 25 %%
. %% − < 0.2.
3n + 1 3 %
25 − 3ε
Substituting ε by 0.2 in the expression , we get the value 13.5; therefore, as p ∈ N, we can choose
9ε
p = Int(13.5) = 13. Thus, if n > 13, the desired result follows. This case is illustrated in Fig. 1.16.
√
n
b) In order to prove by definition that lim √ = 1, we should prove that the proposition
n+1
% √ %
% n %
.∀ε > 0 ∃ p ∈ N : n > p ⇒ %% √ − 1%% < ε
n+1
1.2. Solved Exercises 59
is true, that is, we need to show that for each positive real ε there is at least one order after which, for every
natural n, % √ %
% n %
. %% √ − 1%% < ε.
n+1
Let ε > 0. Solving the inequality for n, we obtain
% √ % %√ √ %
% n % % n − n − 1% √ 1−ε
. %% √ − 1%% < ε ⇔ %% √ %<ε⇔ √ 1 <ε⇔ n> .
n+1 n+1 % n+1 ε
If 1 − ε ≤ 0, the last inequality is true regardless of the choice of p ∈ N. If 1 − ε > 0, then the previous
inequalities are still equivalent to
1−ε 2
.n > .
ε
1−ε 2 1−ε 2
If we choose p ∈ N such that p ≥ , then, as n > p, it follows that n > , that is,
ε ε
% √ %
% n %
. %% √ − 1%% < ε.
n+1
√
n
Figure 1.17: If ε = 0.2, then −ε < √
n+1
− 1 < ε if n > 16
Suppose, for example, that we want to calculate the smallest order p from which
% √ %
% n %
. %% √ − 1%% < 0.2.
n+1
1−ε 2
Substituting ε by 0.2 in the expression , we get the value 16; therefore, we can choose p = 16.
ε
Thus, if n > 16, the desired result follows. This case is illustrated in Fig. 1.17.
√
2 n + 5−n
c) We intend to prove, using the definition, that lim √ = 2. For this, we must prove that the
n+1
proposition % √ %
% 2 n + 5−n %
.∀ε > 0 ∃ p ∈ N : n > p ⇒ %% √ − 2%% < ε
n+1
60 Sequences of Real Numbers
is true, that is, we need to show that for each positive real ε there is at least one order after which, for every
natural n, % √ %
% 2 n + 5−n %
. %% √ − 2%% < ε.
n+1
Let ε > 0.
% √ % % √ √ % % −n %
% 2 n + 5−n % % 2 n + 5−n − 2 n − 2 % % % −n
. %% √ − 2%% < ε ⇔ %% √ % < ε ⇔ % 5√ − 2 % < ε ⇔ 2√− 5 < ε.
n+1 n+1 % % n+1 % n+1
√
2 n+5−n
Figure 1.18: If ε = 0.45, then −ε < √
n+1
− 2 < ε if n > 11
However,
2 − 5−n 2
. √ < √ , ∀n ∈ N;
n+1 n+1
therefore, if
2
.√ < ε,
n+1
then
2 − 5−n
. √ < ε.
n+1
Solving the inequality for n, we obtain
2 √ 2−ε
.√ <ε⇔ n> .
n+1 ε
If 2 − ε ≤ 0, the last inequality is true regardless of the choice of p ∈ N. If 2 − ε > 0, then the previous
inequalities are still equivalent to
2−ε 2
.n > .
ε
2−ε 2 2−ε 2
If we choose p ∈ N such that p ≥ , then, as n > p, it results in n > , that is,
ε ε
% √ %
% 2 n + 5−n %
. %% √ − 2%% < ε.
n+1
1.2. Solved Exercises 61
is true, that is, we need to show that for each positive real ε there is at least one order after which, for every
natural n, % %
% sin(n) %
. %% 1 + − 1%% < ε.
n
Let ε > 0. % % % %
% % % sin(n) %
. %%
1+
sin(n)
− 1%% < ε ⇔ %% % < ε ⇔ |sin(n)| < ε.
n n % n
sin(n)
Figure 1.19: If ε = 0.05, then −ε < 1 + n
− 1 < ε if n > 20
But
|sin(n)| 1
. ≤ , ∀n ∈ N;
n n
therefore, if
1
. < ε,
n
we have
|sin(n)|
. < ε.
n
62 Sequences of Real Numbers
In other words, we must show that for each positive real L, there is at least one order after which, for every
natural n, n3 > L.
√
Let L > 0. √ Solving the inequality for n, we have n3 > L ⇔ n > 3 L. Therefore, if we choose p ∈ N such
3 3
that p > L, we have proven that lim n = +∞.
√
In Fig. 1.20 you can see that by choosing L = 1000, then from n > p = 3 1000 = 10, all terms of the
sequence are greater than L.
n2
.∀L ∈ R+ ∃ p ∈ N : n > p ⇒ >L
n+1
is true, that is, we must show that for each positive real number L, there is at least one order after which,
for every natural number n,
n2
. > L.
n+1
1.2. Solved Exercises 63
n2
Figure 1.21: If L = 10, then n+1
> L if n > 10
15. Let us denote Nn as the number of segments, Ln as the length of each segment, and ln as the length of the
curve in the nth iteration, that is, after repeating n times the process of constructing the curve. We will prove by
induction that Nn = 4n . If n = 1, we have four segments by construction, so N1 = 41 = 4. Suppose (induction
hypothesis) that Nn = 4n . Then Nn+1 = Nn × 4 since each segment originates four new segments. Substituting
the expression of Nn , we get Nn+1 = 4n × 4 = 4n+1 .
1
In each iteration, we divide each segment by 3. During the first iteration, the length of each segment is L1 = .
n n+1 3
1 1
Assuming (induction hypothesis) that Ln = , we can prove that Ln+1 = as follows:
3 3
1 n n+1
Ln 1
.Ln+1 = = 3 = .
3 3 3
n 4 n
The length of the curve in the nth iteration is ln = Nn × Ln . We know that Nn = 4n , so ln = 4n 13 = 3 .
The length L of the Koch curve is determined by the limit of ln when n → +∞. Then, by Theorem 1.1.10,
n
4
.L = lim = +∞.
3
64 Sequences of Real Numbers
2. By definition, prove that the following sequences 6. Let (un ) and (vn ) be two null sequences such that
(un ) are null sequences, that is, that lim un = 0: vn = 0, ∀n ∈ N. We say that (un ) is of higher
un
order than (vn ) if lim = 0. Order the following
1 1 vn
a) un = c) un = √ null sequences:
n n
1 1 1 1 1 1 1 1
1 1 . , √ , , , , , , .
b) un = 2 d) un = n 2n 10 n 2n en n! log(n) n3 nn
n 2
3. Evaluate the limits of the following sequences and 7. Calculate the limit of the sequences with general
provide an argument for the calculations: terms:
2
1−n n + 3 2n 5n + 5 n
a) a) c)
5n + 3 n+1 2n + 1
n2 + 2 n
b) n+5 3 n
3n + 1 b) d) 1 − 2
2n + 1 n
n2 + 3n n2 − 1
c) − 8. Evaluate, if it exists:
n+2 n
3n 1
d) a) lim n
(n + 1)!
4n3 + 1 2n
−n3 + 2 1
e) b) lim n n(n + 1) · · · 2n
4n3 − 7 n
4. Find the limits of the following sequences and justify n!
the calculations: 9. Determine p ∈ R such that lim n
= 3.
(p n)n
√
n
a) 10. Find the limits of the following sequences and pro-
4n + 1
√ vide a clear explanation of the reasoning behind each
n calculation:
b) 1 √
2
− n n
a) cos(x) , x ∈ R
c) n2 + 1 − n2 − 1
1
√ √ 1 b) cos2 (n) sin
d) n+1− n n+ n
2
n (n − 1) (n − 2) (n − 3)
1 1 1 c)
e) √ +√ + ··· + √ (n + 1) (n + 2) (n + 3)
2
n +1 2
n +2 2
n +n
2 n
5. We say that the sequence (un ) grows faster than the d) n n!
un n
sequence (vn ) if → +∞.
vn 1 1 1
e) √ +√ + ··· + √
a) Prove that nn grows faster than n!. n2 + 1 n2 + 2 n2 + 2n + 1
1.3. Proposed Exercises 65
n
1 n+1 15. Consider the sequence of real numbers defined re-
1−
n
f)
n n cursively
g) n
(n + 1)! − n! ⎧
⎪
⎨ x0 = 1
1 1 1
h) √ + √ + ··· + √ .
⎪
1 2
n n+1 2n ⎩xn+1 = xn + , ∀n ∈ N0 .
2 xn
1 1 1
i) + + ··· +
n2 (n + 1)2 (2 n)2
n n n a) Show by induction that
j) √ +√ + ··· + √
n4 + 1 n4 + 2 n4 + n √
. 2 ≤ xn ≤ 2, ∀n ∈ N.
11. For each item, provide, when possible, examples
of sequences u, v, and w such that un → +∞, b) Show that xn ≥ xn+1 , ∀n ≥ 1.
vn → −∞, and wn → 0.
c) Show that the sequence is convergent and cal-
a) un + vn → 1 culate its limit.
If the terms of the sequence .(an ) are all positive, it might seem at first
glance that .(Sn ) is not convergent. In fact, assuming that the sum of an
infinite number of positive terms is a real number is not an intuitive concept.
In this case, intuition fails precisely because we intend to generalize to infinity
a concept, that of sum, which we have intuitive for a finite number of terms.
It is common for intuition to deceive us in cases of “passing” from finite to
infinite.
Zeno of Elea (490–435
It is true that we will not always be able to find the limit of the sequence
B.C.) Greek philosopher .(Sn ), so, by this process, we will not be able to find the sum of an infinite
and disciple of Par- number of terms if it exists. Nevertheless, we are interested in knowing how
menides. Most of his the sequence .(an ) should be so that a real number, the sum of all its terms,
teaching records have
come to us through
is associated with this sequence.
secondary sources, notably
through Aristotle. Zeno’s One of the historical examples illustrating these difficulties in dealing with
great fame came from the concept of an infinite sum of terms is one of the famous Paradoxes of
his paradoxes. These Zeno, which we summarize as follows:
paradoxes aimed to defend
his master Parmenides’ A walker wants to move from one place to another at a constant speed
theory that reality was (Fig. 2.1). In T minutes, he covers half of the distance. He still has the
unique, immutable, and
immobile so that change,
other half to go. In .T /2 minutes, he will cover half of what remains (.1/4
movement, time, and of the total distance). Moreover, now he has .1/4 of the distance left. He
plurality would be nothing will cover half of the rest (.1/8 of the total) in .T /4 minutes. This argument
more than illusions. Par- may be repeated indefinitely. The walker spends
menides’ theory presented
conclusions that contra- T T T T
dicted what the senses T+
. + + + + ...
2 4 8 16
conveyed; Zeno’s goal was
not to present evidence minutes on his journey.
that reinforced the theory
Given that it is the sum of infinite positive terms, we are inclined to assume
itself but to show that
opposing theories also that the sum is infinite. The walker will not be able to reach his goal because
led to contradictions. it will take an infinite amount of time. This reasoning contradicts intuition
(Source of image: Jan and practical life, which is called a paradox.
de Bisschop, Paradig-
mata graphices variorum This is due to the lack of understanding of the concept of an infinite sum
artificum, Rijksmuseum.) of terms (more precisely, of series), which came to be fully understood only
in the 19th century. In this case, the walker will reach his goal in time 2T ,
as intuition tells us.
S1 = a1
S2 = a1 + a 2
S3 = a1 + a 2 + a 3
. .. .
.
Sn = a1 + a2 + a3 + · · · + an
..
.
n
. lim Sn = lim ai
i=1
exists and is finite. If this limit does not exist or is not finite, the series is
divergent.
In the case of convergence, the sum of the series is the value S of the limit
of (Sn ), that is,
∞
.S = lim Sn = an .
n=1
∞
Note: The identification of a series with the symbol n=1 an is a misuse
of language because it is the identification of a series with its sum when it
exists. However, despite being a common misuse of language, it has been
found to be a valuable and harmless convention.
70 Numerical Series
n
n
1 − rn
Sn =
. ai = a1 ri−1 = a1 · .
i=1 i=1
1−r
We know that (Sn ) is convergent if and only if |r| < 1, so the geometric
series is convergent if and only if the absolute value of the ratio of the
geometric sequence that generated it is less than 1. In this case
∞
a1
. an = .
n=1
1−r
The series in Zeno’s Paradox (see Sect. 2.1) is geometric with a1 = T and
r = 1/2; it is convergent with the sum 2T .
If r = 1, the general term is constant; that is,
∞
∞
. an = a1 ,
n=1 n=1
∞
∞ n
3n 1 3
. (−1)n = −
n=1
e2n+1 n=1
e e2
3 3
is geometric with ratio r = − 2 and first term − 3 . Since −1 < r < 1,
e e
the series converges and its sum is
− e33 −3
. = .
1 − (− e32 ) e (e2 + 3)
1
After the first term, we have a geometric series with ratio r = 2 and first
10
25
term 3 . Therefore,
10
25
23 3 1151
2.325 =
. + 10 1 = .
10 1 − 102 495
∞
1
Example 2.2.4 Consider the series √ ; let us study the sequence of
n=1
n
its partial sums and the corresponding limit:
S1 = 1
1
S2 = 1 + √
2
1 1
S3 = 1 + √ + √
. 2 3
..
.
1 1 1
Sn = 1 + √ + √ + · · · + √
2 3 n
..
.
As
1 1 1 1 1 1 1 n √
1 + √ + √ + ··· + √ ≥ √ + √ + √ + ··· + √ = √ = n
.
2 3 n n n n n n
√
and lim n = +∞, we conclude by Theorem 1.1.2 that the sequence (Sn )
has limit +∞. The series under study is divergent.
∞
1
Example 2.2.5 Consider the series , commonly known as the har-
n=1
n
monic series. By mathematical induction, we can prove that the subse-
quence (S2n ) of the partial sums of this series satisfies the inequality
n
S 2n ≥ 1 +
. , ∀n ∈ N.
2
For n = 1, the proposition is true:
1 1
S2 = 1 +
. ≥1+ .
2 2
72 Numerical Series
Assuming that the proposition is valid for n, we prove that it is also valid
for n + 1.
1 1 n 1 n+1
S2n+1 = S2n +
. + · · · + n+1 ≥ 1 + + 2n n+1 = 1 + .
2n +1 2 2 2 2
Thus, the desired inequality is proved using the method of induction. By
Theorem 1.1.2, we can conclude that the sequence (S2n ) has limit +∞,
implying that the sequence of partial sums of the harmonic series does not
have a finite limit. Therefore, the harmonic series is divergent.
Figure 2.2 shows the sequence of partial sums of the harmonic series and its
general term. It illustrates how the sum of an infinite number of increasingly
smaller positive quantities can have a value as large as desired.
1 1 1
. = − ,
n(n + 1) n n+1
we can write the sequence of partial sums as follows (see Fig. 2.3):
1
S1 = 1 −
2
1 1 1 1
S2 = 1 − + − = 1 −
2 2 3 3
Figure 2.3: The sequences 1 1 1 1
. S3 = 1 − + − =1−
1 1
an= n(n+1) and Sn= 1− n+1
3 3 4 4
..
.
1
Sn = 1 −
n+1
..
.
As lim Sn = 1, the series is convergent and its sum is 1:
∞
1
. = 1.
n=1
n(n + 1)
is the sequence
∞
1
Example 2.2.8 The general term of the series can be ex-
n=1
n2 + 3n
1 1 1
pressed as − . The sequence of partial sums can be con-
3 n n+3
structed using this expression:
1 1
S1 = 1−
3 4
1 1 1 1 1 1 1 1 1
S2 = 1− + − = 1− + −
3 4 3 2 5 3 4 2 5
1 1 1 1 1 1 1
S3 = 1− + − + −
3 4 2 5 3 3 6
1 1 1 1 1 1
= 1− + − + −
3 4 2 5 3 6
.
1 1 1 1 1 1 1 1 1
S4 = 1− + − + − + −
3 4 2 5 3 6 3 4 7
1 1 1 1 1 1 1 1
= 1− + − + − + −
3 4 2 5 3 6 4 7
1 1 1 1 1 1
= 1+ − + − −
3 2 5 3 6 7
1 1 1 1 1 1 1 1 1
S5 = 1+ − + − − + −
3 2 5 3 6 7 3 5 8
74 Numerical Series
1 1 1 1 1 1 1 1
= 1+ − + − − + −
3 2 5 3 6 7 5 8
1 1 1 1 1 1
= 1+ + − − −
3 2 3 6 7 8
. ..
.
1 1 1 1 1 1
Sn = 1+ + − − −
3 2 3 n+1 n+2 n+3
..
.
1 1 1 11
As lim Sn = 1+ + = , the series is convergent and
3 2 3 18
∞
1 11
.
2 + 3n
= .
n=1
n 18
The last three examples are particular cases of a type of series called tele-
scopic series. These are series whose general term an can be expressed in
the form αn − αn+k , with k ∈ N:
∞
. (αn − αn+k ), (2.1)
n=1
allowing us to determine the expression of the sequence of partial sums by
canceling consecutive terms and, consequently, in case the sequence con-
verges to easily calculate the sum of the series. These series are convergent
if and only if lim vn , where vn = αn+1 + · · · + αn+k , exists and is finite.
In the particular case of existing, finite, lim αn , we have:
∞
k
. (αn − αn+k ) = αi − k a, (2.2)
n=1 i=1
where a = lim αn . In fact, the sequence of partial sums is given by
n
n
n
= α1 + · · · + αk + αk+1 + · · · + αn
−(αk+1 + · · · + αn + αn+1 + · · · + αn+k )
. = α1 + · · · + αk − (αn+1 + · · · + αn+k )
k
k
= αi − αi+n
i=1 i=1
2.2. Definition of Series: Convergence – General Properties 75
If (αn ) is convergent, then lim αi+n exists and lim αn = lim αi+n from
which it is concluded that
k
k
k
. lim Sn = αi − lim αi+n = αi − k a.
i=1 i=1 i=1
n
Proof: Since the series is convergent, the sequence Sn = ai is also
i=1
convergent. This is true for Sn−1 as well, and we have lim Sn = lim Sn−1 .
Then
∞
n n
Example 2.2.9 The series is divergent because lim = 1.
n=1
n + 1 n + 1
∞
1
Example 2.2.10 Consider the series √ . Although
n=1
n
1
. lim √ = 0,
n
Proof:
a) Let (Sn∗ ) and (Sn∗∗ ) be the sequences of partial sums of the series an
and bn , respectively. Because they are convergent with sums A and B,
we have
∗
. lim Sn = A and lim Sn∗∗ = B.
Let (Sn ) be the sequence of partial sums of the series (an + bn ), that is,
n
n
n
Sn =
. (ai + bi ) = ai + bi = Sn∗ + Sn∗∗ .
i=1 i=1 i=1
Then
Thus,
. lim Sn = lim λSn∗ = λ lim Sn∗ = λA,
that is, the series λan is convergent and has sum λA.
2.2. Definition of Series: Convergence – General Properties 77
Notes:
1. From the proof of item a), it is evident that the given series may
diverge, and yet the series (an + bn ) can still converge. The proof
also indicates that if the sequences of partial sums have infinite limits
of the same sign, that is, both series are divergent, and the sequence
of partial sums will be divergent. The same thing will happen if one
of the series is convergent and the other divergent. If (Sn∗ ) and (Sn∗∗ )
have infinite limits with opposite signs, the series (an + bn ) may
be convergent or divergent because an indeterminate form appears in
the calculation of the limit.
2. Based on the proof of part b), it can be concluded that if the value of
λ is not equal to 0, then the series λan will converge if and only if
the series an also converges. However, if the value of λ equals 0,
then the series λan will converge since all its terms are zero.
1
is also telescopic with αn = and k = 3. In this case too, lim αn = 0;
n+3
1 1 1 37
therefore, the series is convergent and its sum is + + = .
4 5 6 60
According to the previous theorem, the given series is also convergent and
∞
1
n=1
n(n + 3)(n + 6)
∞ ∞
1 1 1 1 1 1
.
= − − −
18 n=1 n n+3 18 n=1 n+3 n+6
1 11 1 37 73
= · − · = .
18 6 18 60 1080
∞ ∞ ∞ ∞
1
n+1 n+1
5n+1 1 5 5 1
.
n+2
= · = and
n=1
8 n=1
8 8 8 n=1 8 n=1
n(n + 1)
5 52
are convergent. The first is geometric with ratio r = and first term 3 .
8 8
The second is telescopic, and as we saw in Example 2.2.6, it converges and
has a sum of 1. We can conclude that the series
∞
5n+1 1
. +
n=1
8n+2 n(n + 1)
25 217
is convergent and its sum is +1= .
192 192
Proof: As
m
n
.|an+1 + · · · + am | = ai − ai = |Sm − Sn |,
i=1 i=1
2.2. Definition of Series: Convergence – General Properties 79
∞
our aim is to prove that the series an converges if and only if
n=1
1 1
. = + ··· +
n+1 n+n
1 1 1 1
≥ + ··· + =n· = .
n+n n+n 2n 2
1
Therefore, the condition of the theorem is not satisfied for δ < , which
2
means that the harmonic series is divergent.
∞
Definition 2.2.3 The remainder of order p of the series an is the
n=1
series ∞
∞
r =
. p an+p = an .
n=1 n=p+1
80 Numerical Series
According to the previous corollary, the remainder of any order will also be
convergent if a series is convergent. The sum of the remainder of order p
of a convergent series gives us the error made by taking the partial sum Sp
as the approximate value of the sum of the series. This error is given by
∞
∞
p ∞
. an − Sp = an − an = an+p = rp .
n=1 n=1 n=1 n=1
∞
∞
Then, the series bn is convergent and bn = an .
n=1 n=1
∞
n
The series bn is convergent if and only if Sn = bi is convergent. But
n=1 i=1
n
k1
k2
kn
kn
Sn =
. bi = ai + ai + · · · + ai = a i = Sk n ,
i=1 i=1 i=k1 +1 i=kn−1 +1 i=1
that is, (Sn ) is a subsequence of (Sn ); therefore, it is convergent and has
the same limit:
∞
∞
. bn = lim Sn = lim Sn = an .
n=1 n=1
∞
Note: If the series an is convergent, then the theorem establishes that
n=1
the following “associative property” is valid:
However, this property is not valid if the series is divergent. For instance,
∞
the series (−1)n is divergent because its general term does not tend to
n=1
zero. Nevertheless, (−1 + 1) + (−1 + 1) + · · · = 0.
82 Numerical Series
Gottfried Wilhelm Leib- Assuming that the first term of the series is positive, we can write
niz (1646–1716) was a
∞
German philosopher and
mathematician. He ob-
. (−1)n−1 an , an > 0, ∀n ∈ N.
tained a doctorate in Law n=1
when he was twenty years
old. He pursued a career
in Law and International
Theorem 2.3.1 (Leibniz’s Test) If (an ) is a decreasing sequence of pos-
Politics as an advisor to itive terms and lim an = 0, then the series (−1)n−1 an is convergent.
kings and princes. During
his countless travels,
Leibniz befriended the Proof: Consider the sequence (Sn ) of the partial sums of the series:
greatest intellectuals of
his time. He developed .Sn = a1 − a2 + a3 − · · · + (−1)n−1 an .
differential and integral
calculus simultaneously We will study the subsequences of even order terms and odd order terms.
and independently of the Let k ∈ N be arbitrary.
English mathematician
Isaac Newton and estab-
lished the foundations of
S2k = a1 − a2 + · · · + a2k−1 − a2k
.
dynamics. Leibniz was one S2k+1 = a1 − a2 + · · · + a2k−1 − a2k + a2k+1 .
of the last intellectuals to
master almost all of the We claim that the subsequence (S2k ) is increasing. This is because, as (an )
knowledge of his time.
is decreasing, we get:
(Source of image: Oil
portrait by Christoph
S2k+2 − S2k = a1 − a2 + · · · + a2k−1 − a2k + a2k+1 − a2k+2 −
Bernhard Francke (1665–
1729), Herzog Anton . −(a1 − a2 + · · · + a2k−1 − a2k )
Ulrich-Museum.)
= a2k+1 − a2k+2 ≥ 0.
As the subsequences of even and odd terms have the same limit, (Sn )
∞
converges. Therefore, the series (−1)n−1 an is convergent.
n=1
is convergent. As this series differs from the initial series in one term, we
can affirm that the series under study is convergent.
84 Numerical Series
∞
π
Example 2.3.2 The series cos(nπ) sin is alternating because
n=2
n
π
. cos(nπ) = (−1)n and an = sin > 0, ∀n > 1
n
π π
(note that 0 < ≤ , ∀n > 1). We can write the series in the form
n 2
∞ π
n
(−1) sin . Let us check the conditions of Leibniz’s test:
n=2
n
π
(i) lim sin = 0.
n
π
(ii) We have already seen that sin > 0, ∀n > 1.
n
π π
(iii) As we also justified earlier, 0 < ≤ , ∀n > 1; the sequence
π n 2
π
is decreasing and the sine function is increasing on 0, , so
n π 2
sin is decreasing for n ≥ 2.
n
We can conclude that the alternating series is convergent.
∞
1
Example 2.3.3 Consider the series (−1)nα
, α ∈ R.
n=1
n
If α ≤ 0, the series diverges because its general term does not tend to zero.
1
If α > 0, the series is convergent because an = α is a decreasing sequence
n
with positive terms and lim an = 0.
Not all alternating series meet the conditions of Leibniz’s test. In this
situation, the test is not applicable, and it is necessary to look for another
approach.
∞ ∞
(−1)n (−1)n 1 (−1)n
. √ 1+ √ = (−1)n √ + .
n=1
n n n=1
n n
1 (−1)n
Although an = √ + is a null sequence and an > 0, ∀n > 1, we
n n
cannot apply Leibniz’s test because (an ) is not decreasing. However, we can
2.3. Alternating Series 85
show that the given series is divergent because it is the sum of a convergent
1 1
series, the series (−1)n √ , with a divergent series, the series
n n
(see Note 1 on page 77).
0 ≤ (−1)n (S − Sn ) ≤ an+1 , ∀n ∈ N.
.
Proof: We know from the proof of the previous theorem that (S2k ) is
an increasing subsequence of (Sn ), and it has the same limit, S, as the
subsequence (S2k+1 ). Similarly, (S2k+1 ) is decreasing. This leads to the
following inequalities:
Therefore:
0 ≤ S2k−1 − S ≤ a2k ,
.
0 ≤ S − S2k ≤ a2k+1 .
This implies that
0 ≤ (−1)n (S − Sn ) ≤ an+1 .
.
|S − Sn | ≤ an+1 , ∀n ∈ N.
.
86 Numerical Series
Note: According to the previous corollary, it is evident that under the con-
ditions of Leibniz’s test, the absolute value of the error made in using a
partial sum as an estimation for the sum of an alternating series is, in abso-
lute value, always less than the absolute value of the first of the disregarded
terms.
∞
Proof: By Theorem 2.2.3, the series |an | is convergent if and only if
n=1
Since
|an+1 + · · · + am | ≤ |an+1 | + · · · + |am |
.
and
| |an+1 | + · · · + |am | | = |an+1 | + · · · + |am |,
.
we have
∞
that is, the series an is convergent.
n=1
Note: It is worth noting that the converse of this theorem is not true. This
means that the series an can be convergent even if the series of modules,
|an |, is not. For instance, consider the harmonic series, which diverges,
and the alternating harmonic series, which converges. The harmonic series
is the series of modules of the alternating harmonic series.
∞
1
Example 2.4.1 Consider the alternating harmonic series, (−1)n−1
,
n=1
n
which we know to be conditionally convergent. Let us rearrange its terms
such that each positive term is followed by two negative terms. We will
obtain the following series:
1 1 1 1 1 1 1 1
1−
. − + − − + − − + ···
2 4 3 6 8 5 10 12
For this series, we have the following partial sums:
S1 = 1
1
S2 = 1 −
2
1 1 1 1 1 1 1
S3 = 1 − − = − = 1− = S2
. 2 4 2 4 2 2 2
1 1 1
S4 = 1 − − +
2 4 3
1 1 1 1
S5 = 1 − − + −
2 4 3 6
2.4. Absolute Convergence 89
1 1 1 1 1 1 1 1 1 1
S6 = 1 − − + − − = 1− − + − −
2 4 3 6 8 2 4 3 6 8
1 1 1 1 1 1 1 1 1
= − + − = 1− + − = S4
2 4 6 8 2 2 3 4 2
..
.
1 1 1 1 1 1 1 1 1 1
. S9 = S6 + − − = 1− + − + − −
5 10 12 2 2 3 4 5 10 12
1 1 1 1 1 1 1 1 1 1 1 1
= 1− + − + − = 1− + − + −
2 2 3 4 10 12 2 2 3 4 5 6
1
= S6
2
..
.
n
1
where Sn = (−1)i−1 .
i=1
i
1
Using induction, we can demonstrate that S3n = S2n , which implies that
2
if we designate by S the limit of the sequence (Sn ), then we have
1
. lim S3n = S.
2
1 1 1
As S3n+1 = S3n + and S3n+2 = S3n + − , we have
2n + 1 2n + 1 4n + 2
1
. lim S3n+1 = lim S3n + lim
2n + 1
and
1 1
. lim S3n+2 = lim S3n + lim − lim ,
2n + 1 4n + 2
that is,
1
. lim S3n+1 = lim S3n+2 = lim S3n = S.
2
Therefore, we conclude that lim Sn = 12 S. In other words, the series ob-
tained by this rearrangement of the terms of the alternating harmonic series
is convergent, and its sum equals half the original series’ sum.
90 Numerical Series
Note that the sequences .(Sn ) and .(Tn ) are increasing, with positive terms.
Suppose that the series . 2n a2n is convergent. By Theorem 2.5.1, the
sequence .(Tn ) is bounded. Then
= a1 + Tn−1 .
2.5. Series of Nonnegative Terms 91
≥ a1 + a2 + 2 a4 + 4 a8 + · · · + 2n−1 a2n
.
1
= a1 + (2 a2 + 4 a4 + 8 a8 + · · · + 2n a2n )
2
1
= a1 + Tn .
2
This implies that the sequence .(Tn ) is also bounded. Hence, the series
. 2n a2n is convergent.
∞
1
Example 2.5.1 Consider the series . , .p ∈ R, usually referred to as
n=1
np
the p-series.
If .p ≤ 0, the series is divergent because its general term does not tend to
zero.
1
If .p > 0, we can apply Theorem 2.5.2 because the sequence .an = p is
n
decreasing with positive terms. The series
∞
∞ ∞ n
1 2n 1
. 2n p = n =
n=1
(2n ) n=1
(2p ) n=1
2p−1
1 1
is geometric, with ratio . p−1 , and is convergent if and only if . p−1 < 1,
2 2
that is, if and only if .p > 1.
∞
1
Conclusion: The p-series . p
converges if and only if .p > 1.
n=1
n
92 Numerical Series
∞
1
Example 2.5.2 Consider the series . α, where .α ∈ R. The
n=2
n log(n)
function
1
f (x) =
. α
x log(x)
is positive and continuous on .[2, +∞[. We can find the derivative of f to
determine whether it is increasing or decreasing. We have
α α−1
log(x) + α log(x)
f (x) = 0 ⇔− 2α =0
x2 log(x)
α−1
. log(x) log(x) + α
⇔− 2α =0
x2 log(x)
⇔ log(x) + α = 0.
If .α ≥ 0, this equation has no root in .[2, +∞[, which implies that .f (x) < 0,
.∀x ∈ [2, +∞[ .
If .α < 0, then .f (x) < 0, .∀x > e−α , and this implies that .f (x) < 0,
−α
.∀x ∈ ]e , +∞[ .
1
Then, the sequence .an = α is decreasing from the order q if
n log(n)
−α
.q ∈ N is such that .q ≥ max{2, e }. The series
∞
∞
∞
2n 1 1 1
. α = α = α α
n=q
2n log(2n ) n=q
nα log(2) log(2) n=q
n
⎧
⎪ 1
⎨ , if n = 2k
2k
.an =
⎪
⎩0, if n = 2k .
The series
∞
∞ ∞
2n
. 2n a2n = = 1
n=1 n=1
2n n=1
n
.Sn − a1 = a2 + a3 + a4 + · · · + an ≤ f (x) dx.
1
If we consider the area of the circumscribed rectangles (see Fig. 2.5), we Figure 2.4: Inscribed rectan-
have the inequality gles in the graph of f
n
.Sn−1 = a1 + a2 + a3 + · · · + an−1 ≥ f (x) dx.
1
From the two previous inequalities, we can conclude that
n
.Sn − a1 ≤ f (x) dx ≤ Sn−1 .
1
n Figure 2.5: Circumscribed
If the integral is divergent, . lim f (x) dx = +∞ since f is positive. rectangles in the graph of f
n→+∞ 1
Then, by the inequality on the right, the limit of the sequence of the partial
94 Numerical Series
sums of the series is also .+∞; that is, the series diverges.
n If the integral
converges, then there exists and it is finite . lim f (x) dx. Conse-
n→+∞ 1
n
quently, the sequence .Sn = ai is bounded. As the terms of the series
i=1
are positive, it is convergent by Theorem 2.5.1. n
If the series is convergent, by the inequality on the right, . f (x) dx is
1
bounded, so the improper integral is convergent.
n If the series is divergent,
based on the inequality on the left, . lim f (x) dx = +∞.
n→+∞ 1
Note: When we use the Integral Test, it is not necessary for the series or
the integral to start at .n = 1. Additionally, the function f does not need
to decrease on the interval .[1, +∞[, but only on some interval of the form
.[N, +∞[, .N ∈ N.
∞
en
Example 2.5.3 Consider the series . . The real-valued function
n=1
1 + e2n
ex
f of a real variable defined by .f (x) = is continuous and positive
1 + e2x
on .[1, +∞[. Let us study the monotonicity of f :
∞
log(n)
Example 2.5.4 Consider the series . . The real-valued function
n=2
n
log(x)
f of a real variable defined by .f (x) = is continuous and positive on
x
2.5. Series of Nonnegative Terms 95
1
· x − log(x) 1 − log(x)
.f (x) =
x = < 0, ∀x ∈ ]e, +∞[.
x 2 x2
Thus, the function f decreases on .]e, +∞[. Now, we compute the integral:
t
log(x) 1
. lim dx = lim (log(t))2 − (log(2))2 = +∞.
t→+∞ 1 x t→+∞ 2
∞
log(n)
According to the Integral Test, the series . diverges since the
n
+∞
n=2
log(x)
improper integral . dx is also divergent.
1 x
Proof: Let .(Sn ) and .(Sn ) be the sequences of partial sums of series . an
and . bn , that is,
n
n
Sn =
. ai and Sn = bi .
i=1 i=1
Note: Omitting a finite number of terms does not change the convergence
or divergence of the series, as we have seen. Therefore, the previous theorem
remains valid if there exists .p ∈ N such that .an ≤ bn , .∀n ≥ p.
96 Numerical Series
∞
1
Example 2.5.5 Consider the series . . We know that
n=1
n!
1 1 1
0<
. = ≤ n−1 , ∀n ∈ N.
n! n(n − 1)(n − 2) . . . 2 2
∞
1
Also, the series .
n−1
is convergent because it is geometric with ratio
n=1
2
1
. . By the General Comparison Test, we can conclude that the given series
2
is convergent.
∞
cos(n)
Example 2.5.6 The series . is not of positive terms. However,
n=1
n2
we can observe that the following inequality holds:
cos(n) 1
.0 <
n2 ≤ n2 , ∀n ∈ N. (2.4)
∞
1
The series . is a p-series with .p = 2 > 1 and is, therefore, convergent.
n=1
n2
By applying the General Comparison Test and taking into
account inequality
∞
cos(n)
(2.4), we can conclude that the series .
n2 is convergent. This
n=1
∞
cos(n)
means that the series . is absolutely convergent.
n=1
n2
∞
sin(n)
Example 2.5.7 Consider the series . (−1)n n. Let us analyze
n=1
log(10)
the series of modules:
∞
∞
| sin(n)|
sin(n)
. (−1)n n = n.
log(10) log(10)
n=1 n=1
We notice that the following inequality holds for every natural number n:
| sin(n)| 1
0<
. n ≤ n.
log(10) log(10)
2.5. Series of Nonnegative Terms 97
∞
1 1
The series . n is geometric with ratio .r = . Since
n=1
log(10) log(10)
1
10 > e ⇒ log(10) > 1 ⇒ 0 <
. < 1,
log(10)
we have .|r| < 1. Therefore, the geometric series converges. By applying
∞
| sin(n)|
the General Comparison Test, the series . n is convergent. This
n=1
log(10)
implies that the given series is absolutely convergent.
an
Proof: Suppose .lim = k. By Definition 1.1.7,
bn
an
.∀δ > 0 ∃ p ∈ N : n > p ⇒ − k < δ.
bn
k
Let .δ = . From a certain order, we have
2
an k k an k k an 3
.
bn − k < 2 ⇔ − 2 < bn − k < 2 ⇔ 2 < bn < 2 k,
98 Numerical Series
which implies
3 k
a <
. n k bn and bn < a n .
2 2
From these inequalities and using Corollary 1, we get the intended result.
an
Proof: Let .lim = 0. By Definition 1.1.7,
bn
an
.∀δ > 0 ∃ p ∈ N : n > p ⇒
bn < δ.
since .an ≥ 0 and .bn > 0. Consequently, .0 ≤ an < δbn , and from Corollary 1
of Theorem 2.5.4, the result follows.
an
Proof: Let .lim = +∞. By Definition 1.1.6,
bn
an
∀δ > 0 ∃ p ∈ N : n > p ⇒
. > δ.
bn
2.5. Series of Nonnegative Terms 99
Corollary 5 Let . an and . bn be two series such that .an > 0 and
b > 0, .∀n ∈ N. If there exists .p ∈ N such that
. n
an+1 bn+1
. ≤ , ∀n ≥ p,
an bn
then:
an
This means that the sequence . is decreasing after order p. Therefore,
bn
ap an
there exists a constant k (we can take .k = ) such that . ≤ k. In other
bp bn
words:
a ≤ k bn , ∀n ≥ p.
. n
∞
1 + (−1)n
Example 2.5.8 The terms of the series . are nonnegative.
n=1
n2
As
1 + (−1)n 2
.0≤ ≤ 2 , ∀n ∈ N,
n2 n
∞
1
and the series . 2
is convergent, Corollary 1 of Theorem 2.5.4 allows us
n=1
n
to conclude that the given series is convergent.
100 Numerical Series
∞
2n2 + n
Example 2.5.9 Consider the series of positive terms . . Since
n=1
3n5 + 3
2n2 + n
3n5 + 3 2n5 + n4 2
. lim = lim = ,
1 5
3n + 3 3
n3
∞
2n2 + n
by Corollary 2 of Theorem 2.5.4, the series . is convergent be-
n=1
3n5 + 3
∞
1
cause . 3
converges.
n=1
n
∞
log(n)
Example 2.5.10 The series . is of nonnegative terms. Since
n=1
n3
∞
1
.
2
is convergent and
n=1
n
log(n)
n3 log(n)
. lim = lim = 0,
1 n
n2
by Corollary 3 of Theorem 2.5.4, the original series is also convergent.
∞
arctan(n) + n
Example 2.5.12 The terms of the series . √ are positive.
n=1
n n+1
∞
1
Let us consider the series . , which is divergent because it is a p-
n=1
n1/2
1
series with .p = . The limit
2
arctan(n) + n arctan(n)
√ 1/2 +1
n n+1 (arctan(n) + n) n n
. lim = lim √ = lim =1
1 n n+1 1
1 + 3/2
n1/2 n
is finite and different from zero. Therefore, by Corollary 2 of the General
∞
arctan(n) + n
Comparison Test, the series . √ diverges.
n=1
n n+1
Proof:
a) The series . bn = rn is geometric with ratio r and is convergent since
.0 < r < 1. Moreover, we have
an+1 rn+1
. ≤ n = r, ∀n ≥ p,
an r
which implies by Corollary 5 of the General Comparison Test that the series
. an is convergent.
b) The series . bn = 1 is divergent. Additionally, we have
an+1 bn+1
. ≥1= , ∀n ≥ p.
an bn
By Corollary 5 of the General Comparison Test, we can conclude that the
series . an is divergent.
102 Numerical Series
an+1
However, if .lim = 1 and the convergence is for values greater than 1,
an
an+1
that is, there is an order .p ∈ N from which . ≥ 1, then, by the Ratio
an
∞
Test, the series . an diverges.
n=1
∞
k n n!
Example 2.5.13 Let .k > 0. The series . is of positive terms. As
n=1
nn
k n+1 (n + 1)!
n
(n + 1)n+1 k n+1 (n + 1)! nn n 1
. lim = lim = lim k · =k· ,
k n n! k n n! (n + 1)n+1 n+1 e
nn
k k
we can apply D’Alembert’s test: If . < 1, the series is convergent; if . > 1,
e e
the series is divergent.
n
k n+1
If = 1, D’Alembert’s test is inconclusive. However, as .
. is an
e n
n
n
increasing sequence with limit e, . is a decreasing sequence with
n+1
n
1 n
limit . , which implies that .e · is decreasing with limit 1, that is,
e n+1
an+1
. tends to 1 from above. Then, if .k = e, the series is divergent.
an
∞
2
(n + 1)!
Example 2.5.14 The series . (−1)n is alternating. Let us
n=1
(2n)! 5n
∞ 2
(n + 1)!
start by studying the series of modules, . , using D’Alembert’s
n=1
(2n)! 5n
test. Let .an be the general term of the series of modules:
2
(n + 2)!
2
an+1 2(n + 1) ! 5n+1 (n + 2)! (2n)! 5n
lim = lim 2 = lim 2
an (n + 1)! 2(n + 1) ! 5n+1 (n + 1)!
.
(2n)! 5n
2
(n + 2)! (2n)!
= lim 2
(2n + 2)(2n + 1)(2n)! 5 (n + 1)!
104 Numerical Series
∞
(n!)2 + n!
Example 2.5.15 The series . is of positive terms and
n=1
(4n)! + n4
(n!)2 + n! 2(n!)2
0<
.
4
< , ∀n ∈ N.
(4n)! + n (4n)!
∞
2(n!)2
Let us study the series . by D’Alembert’s test:
n=1
(4n)!
2
2 (n + 1)!
(4n + 4)! (n + 1)2
. lim = lim = 0 < 1.
2(n!)2 (4n + 4)(4n + 3)(4n + 2)(4n + 1)
(4n)!
∞
2(n!)2
Therefore, the series . converges, so by the General Comparison
n=1
(4n)!
Test, the given series converges.
1 1 1 1 1 1 1 1 1
. + · + 2 · + 2 · 2 + 3 · 2 + ··· ,
2 2 3 2 3 2 3 2 3
1 1 1 1 1 1 1
that is, .a1 = , .a2 = · , .a3 = 2 · , .a4 = 2 · 2 , . . . ; in general,
2 2 3 2 3 2 3
⎧
⎪
⎪
⎪
⎪
1 1
⎨ 2 n2 · 3 n2 ,
⎪ if n is even
.an =
⎪
⎪ 1 1
⎪
⎪ n+1 · n−1 , if n is odd.
⎪
⎩2 2
3 2
2.5. Series of Nonnegative Terms 105
Then
⎧
⎪
⎪ 1 1
⎪
⎪ n+2 ·
n n
⎪
⎪
n
22 · 32 1
⎪
⎪ 2 2
3 2
= 2−1 = ,
⎪
⎪
= n+2 if n is even
⎪ 1 1 n
⎪ 2 2 · 32 2
an+1 ⎨ n2 · n2
. = 2 3
an ⎪
⎪ 1 1
⎪
⎪ n+1 ·
⎪
n+1 n−1
⎪
⎪2
n+1
2 2 ·3 2 1
⎪ −1
2 2
⎪ 3 = n+1
⎪ n+1 = 3 = , if n is odd.
⎪
⎪ 1 1 2 2 ·3 2 3
⎩ n+1 · n−1
2 2
2 3
an+1
Note that we cannot use D’Alembert’s test, as .lim does not exist.
an
However, it follows from the Ratio Test that the series converges because
an+1 1
. ≤ < 1, ∀n ∈ N.
an 2
∞
(n!)2
Example 2.5.17 Consider the series . (−1)n
. Let .an be the gen-
n=1
2n nn
eral term of the series of modules. Using D’Alembert’s test, we get
2
(n + 1)!
2
an+1 2n+1 (n + 1)n+1 (n + 1)! 2n nn
lim = lim = lim n+1
.
an (n!)2 2 (n + 1)n+1 (n!)2
n
2 n n
n
n+1 n
= lim = +∞.
2 n+1
The limit value is greater than 1, which means the series of modules is
divergent. Since we used D’Alembert’s test, the given series is also divergent.
106 Numerical Series
Proof:
√
a) If . n an ≤ r, ∀n > p, then .an ≤ rn < 1, .∀n ≥ p. The series . rn is
convergent because it is a geometric series with ratio r, where .0 < r < 1.
Therefore, the series . an is convergent.
√
b) If . kn akn ≥ 1, ∀kn > p, then .akn ≥ 1, .∀kn > p, so it does not tend to
zero. As a result, .(an ) is not a null sequence, which implies that the series
. an is divergent.
√
Proof: Let .a = lim n an .
a) Let r be such that .a < r < 1. We can affirm that
√
.∃ p ∈ N : n > p ⇒ an < r,
n
√
Note: If .lim n an = 1, the test is inconclusive, as there exist both conver-
gent and divergent series in this situation. For instance, the harmonic series
∞
1
. is divergent and
n
n=1
n 1 1
. lim = lim √ = 1,
n n
n
∞
1
and the series . 2
is convergent and
n
n=1
1 1
= lim √
n
. lim = 1.
n2 n
n2
∞
n 2
n+1
Example 2.5.18 The series . is of positive terms. As
n
n=1
n2 n
n n+1 n+1
. lim = lim = e > 1,
n n
by Cauchy’s Root Test, the series is divergent.
∞
1
Example 2.5.19 Let us consider the series . .
(3 + (−1)n )n
⎧ n=1
⎪
⎪
⎪
⎪ 1
1 ⎨ , if n is even
n 4
. =
(3 + (−1)n )n ⎪
⎪ 1
⎪
⎪
⎩ 2 , if n is odd.
√ 1
Then . n an ≤ < 1, .∀n ∈ N, and by the Root Test, the series is convergent.
2
∞
2n
∞ 2n
log(n) log(n)
Example 2.5.20 The series . = is of
n=1
nn n=2
nn
positive terms. We will study this series using the Cauchy’s Root Test:
2n 2
n log(n) log(n)
. lim = lim =0
nn n
108 Numerical Series
2
log(x) 2 log(x)
. lim = lim = 0.
x→+∞ (x) x→+∞ x
⎧
⎨(1 − √
n
n)n , if n is odd
a =
. n
⎩n2 e−n , if n is even.
⎧ √
⎨ n |(1 − n n)n |, if n is odd
n
|an | = √
⎩ n n2 e−n , if n is even
.
⎧√
⎨ n n − 1, if n is odd
=
⎩e−1 √n
n2 , if n is even.
√ √
Since .lim( n n−1) = 0 and .lim e−1 n2 = e−1 , we obtain .lim n |an | = e−1 .
n
∞
As .lim n |an | < 1, the series . |an | is convergent. Therefore, the series
n=1
∞
. an is absolutely convergent.
n=1
Note: The Cauchy’s Root Test is more general than D’Alembert’s test.
This means that if Cauchy’s Root Test is inconclusive about a series, then
D’Alembert’s test will also be inconclusive. In fact, by Theorem 1.1.11
an+1 √
.lim
an .= a ⇒ lim
n a
n = a. This theorem is significant because if .a = 1,
then D’Alembert’s test is inconclusive, and the same happens for Cauchy’s
Root Test. However, it is essential to note that the reciprocal is not true.
It may be possible to draw conclusions through Cauchy’s Root Test, even if
D’Alembert’s test fails to do so.
2.5. Series of Nonnegative Terms 109
∞
2−n−(−1) . Using Cauchy’s
n
Example 2.5.22 Let us consider the series .
n=1
Root Test,
(−1)n 1
2−n−(−1) = lim 2−1 2−
n n
. lim n = < 1,
2
2−(n+1)−(−1)
n+1
2, if n is even
= 2−n−1−(−1)
n+1 n
+n+(−1)
.
−n−(−1)n
= −3
2 2 , if n is odd.
1 an 1
Proof: If . k ∈ R, .lim · − = k is equivalent to
bn an+1 bn+1
1 an 1
∀δ > 0 ∃p ∈ N : ∀n > p, ·
. − − k < δ.
bn an+1 bn+1
But
1 an 1 1 an 1
.
bn · an+1 − bn+1 − k < δ ⇔ k − δ < bn · an+1 − bn+1 < k + δ.
110 Numerical Series
k
a) Let .k ∈ R+ and .δ = . There exists an order .n0 ∈ N from which we
2
have
k 1 an 1 k 1 an 1
k− < · − ⇔ < · −
2 bn an+1 bn+1 2 bn an+1 bn+1
2 1 an 1 2 1 an 1
. ⇔ 1< · − ⇔ an+1 < an+1 · −
k bn an+1 bn+1 k bn an+1 bn+1
2 an an+1
⇔ an+1 < − .
k bn bn+1
n+1
n+1
2 ai−1 ai
ai < −
i=n0 +1 i=n0 +1
k bi−1 bi
n+1
2 an0 an +1 an +1 an +2 an an+1
. ⇔ ai < − 0 + 0 − 0 + ··· + −
i=n0 +1
k bn0 bn0 +1 bn0 +1 bn0 +2 bn bn+1
n+1
2 an0 an+1 2 an0
⇔ ai < − < .
i=n0 +1
k bn0 bn+1 k bn0
n+1
2 an0
0 < Sn+1 =
. ai = Sn0 + an0 +1 + · · · + an+1 ≤ Sn0 + ,
i=1
k bn0
1 an 1 an bn an+1 bn+1
. · − <0⇔ < ⇔ > .
bn an+1 bn+1 an+1 bn+1 an bn
1 an 1
. · − < 0,
bn an+1 bn+1
∞
1 1
Proof: In Kummer’s test, consider .bn = . The series . is divergent,
n n=1
n
and we have
1 an 1 an
lim · − = lim n · −n−1
bn an+1 bn+1 an+1
.
an
= lim n − 1 − 1 = a − 1. Joseph L. Raabe (1801–
an+1 1859), was a Swiss mathe-
matician and physicist. He
This demonstrates the corollary. was a professor at the
Zurich Polytechnic Insti-
tute. His name is mainly
Note: Often, cases that are inconclusive by D’Alembert’s test can be solved
associated with the con-
by Raabe’s test. vergence test for a series
of positive terms that ex-
tends D’Alembert’s Test.
He also studied various as-
Example 2.5.23 Consider the series pects of planetary move-
ments. (Source of im-
∞ ∞
1 × 3 × · · · × (2n − 1) 1 age: Lithographie by Carl
. · = an . Friedrich Irminger (1813–
n=1
2 × 4 × · · · × 2n n n=1
1863).)
We have
an+1 n(2n + 1)
. lim = lim = 1,
an (n + 1)(2n + 2)
112 Numerical Series
an (n + 1)(2n + 2)
lim n −1 = lim n −1
an+1 n(2n + 1)
.
(n + 1)(2n + 2) − n(2n + 1) 3n + 2 3
= lim = lim = > 1,
2n + 1 2n + 1 2
Raabe’s test shows that the series is convergent.
2.6. Products of Series 113
obtaining
∞
∞
∞ ∞
. an bk = an · B = B · an = B × A.
n=1 k=1 n=1 n=1
However, one may ask if it is possible to form a series whose terms are
products of the form .an bk , in some order, so that the sum of this series is
.A × B. The answer to this question is given in the following theorem:
c = cφ(i,j) = ai × bj .
. n
∞
Definition 2.6.1 The Cauchy product of two convergent series, . an
∞ ∞
n=1
n
and . bn , is the series . ak bn−k+1 .
n=1 n=1 k=1
114 Numerical Series
∞
xn
Example 2.6.1 Consider the series . , .x ∈ R. As
n=0
n!
n+1
x
(n + 1)! |x|
. lim n = lim = 0, ∀x ∈ R,
x n +1
n!
n=0
n! k!(n − k)!
k=0
∞
(x + y)n
= .
n=0
n!
Note: The Cauchy product of two series that are not absolutely convergent
may lead to a divergent series.
∞
(−1)n
Example 2.6.2 The series .√ is conditionally convergent. Cal-
n=0
n+1
culating the Cauchy product of the series with itself, we get
2.6. Products of Series 115
∞
n
(−1)k (−1)n−k
·
n=0 k=0 (k + 1) (n − k + 1)
∞
n
(−1)n
.= √ √
n=0 k=0
k+1 n−k+1
∞
n ∞
1
= (−1) n
√ √ = (−1)n an ,
n=0 k=0
k + 1 n − k + 1 n=0
∞
(−1)n
Example 2.6.3 The series . is conditionally convergent, and the
n=1
n
∞
(−1)n
series . is absolutely convergent. By Mertens’s Theorem, the
n=1
n2
Cauchy product of the series, which is alternating, is convergent:
∞
∞ ∞
n
(−1)n (−1)n (−1)k (−1)n−k+1
× = ·
n=1
n n=1
n2 n=1 k=1
k (n − k + 1)2
.
∞
n
1
n+1
= (−1) .
n=1
k(n − k + 1)2
k=1
116 Numerical Series
∞
∞
log(n7 + 1)
1 c)
a) n2
n2 + 3n + 2 n=1
n=1
∞
∞ 1 √
d) n+ − n
b) arctan(n + 3) − arctan(n) n
n=1
∞ √
n=1
√
∞ n+1− n
1 1 e) √ √
c) n
− n+1 n=1
n+1+ n
n=0
e e
∞
∞ cos(n)
1 1 f)
d) − n=1
n2 + 4
n=1
arctan(n) arctan(n + 1)
∞
1
∞
g)
1 1 n3 + 10 cos(n)
e) − n=1
22n−1 23n−2
n=1 ∞
1 − (−1)n
∞
h) √
3n + 7n n
f) n=1
n=0
3n · 7n
5. Investigate the convergence of the following series:
∞
1 2 3
g) · n · n+1 ∞
log(n)
n=0
4n−1 5 6 a) (−1)n
n=1
n
∞
log( n+2 ) ∞
h) n
2
n=2
log(n) log(n + 2) b) (−1)n+1
n=1
en + e−n
2. Show that the series ∞
1
c) (−1)n
∞
n=1
log(3n)
1 1
. + n ∞ 1
(n + 1)2 π sin n
n=2 n
d) (−1)
n=1
n
is convergent and find its sum, knowing that ∞
∞ 3n
1 π2 e) (−1)n
2
= . e2n+1
n 6 n=1
n=1
∞
sin nπ
3. Write the following repeating decimals as fractions: f) √ 2
n=1
n+1
a) 0, 5 c) 1, 345
∞
(−1)n+1
b) 0, 34 d) 0, 324101 6. Consider the series :
n=1
n
4. Test the convergence of the following series using a
comparison test: a) Study it for convergence.
b) Indicate a partial sum Sn that approximates
∞
2 n2 − 1 the sum of the series with an error less than
a) 1
n=1
3 n5 + 2 n + 1 1000
.
2.7. Solved Exercises 117
c) Indicate an upper bound of the error commit- 9. Test for convergence or divergence the following se-
ted when S5 is taken as the sum of the series. ries. If convergence occurs, indicate whether it is
conditional or absolute:
7. Use the Ratio Test or D’Alembert’s Test to deter- ∞
(−1)n
mine the convergence or divergence of the following a)
log(n n)
series: n=2
∞
∞
nn n2 2 n
a) b) √ + 1+
π n n! n=1 n5 + 1 n
n=1
∞
∞
1
n 2n c) log 1 − 2
b) n
n=0
4 n3 + 1 n=2
∞
∞ sin(n) + 2n
(n + 1)! − n! d)
c) n + 5n
n=1
n! (n + 1)! n=1
∞
∞ en+1
2 · 4 · 6 · · · · · (2n + 2) e) (−1)n
d) nn
n=1
1 · 4 · 7 · · · · · (3n + 1) n=1
1
∞
∞ (−1)n +
(n + 1)! n
e) √ f)
n=1
nn 3n + 2 n=1
n
∞ ∞
3n + n! 1
f) g) log 1 + n
n! + nn n=1
2
n=1
∞
∞
n! √ n
g) h) 1− n
n
n=1
(2 n)! + 2n n=1
∞
∞
n! + 3n (−1)n
h) i)
((n + 1)!)2 n=1
1 − (−1)n n2
n=0
(−1)n + 4 3 2n
∞
8. Use the Root Test or Cauchy’s Root Test to deter- j) +
n=1
n + 4n n+2
mine the convergence or divergence of the following
∞
series: 1 n
k) 1−
∞
n2 n=1
n
2
a) 1− ∞
n 1
n=3 l) 3
∞
n=2 n log(n)
2
b) e−n ∞ n+2 1
n=0 m) dx
∞ n x2
√
n n
n=1
c) 2−1 ∞
(−1)n
n=1 n) √ √
n+1+ 3n+1
∞ n n=1
π 1 ∞
d) cos + 1
n=1
6 n o) √ √
nn+ 1+ n
∞ n n=1
1 n ∞
e) + (−1)n 1
n=0
2 4n + 1 p) (−1)n 1 − cos
n
∞ 2 n n=1
n −2 ∞
f) 1
3n2 q) (−1)n 2n sin n
n=1 3
n=1
118 Numerical Series
∞
√ ∞
n! n 1 1
r) (−1)n √ n) arcsin
n=1
nn n + 1 n=1
2n + 5 n
∞
√ 1 11. Determine whether the following series are conver-
s) ( n + 1) sin
n=1
n2 gent or divergent. If convergence occurs, indicate if
∞
it is conditional or absolute:
1
t)
∞
n=2
log(n!) 1 1
a) sin n
Hint: Note that n! < nn n=1
n 2
∞ √
3
n2
10. Assess the convergence or divergence of the given b) √
series. If the series converges, specify whether the n=1
n n + 2n2
convergence is absolute or conditional: ∞
cos(πn)
∞
c) √
n cos(n π) n=2 n2 − 1
a) √
n3 + 1 ∞
n=0 (−1)n
∞
d) √
n! n=2
n log(n)
b)
3 × 5 × · · · × (2 n + 1) ∞
n=1 4 + sin(n)
∞
e) √
1 + cos2 (n) n=1
3
n+1
c) √
n ∞
n=1 n2n sin(n3 )
∞
f)
3n (n + 1)! n=1
(3n2 + 5)n
d)
(n + 1)n ∞
n=0 n−1
∞ g) (−1)n arcsin
(−1)n n2 + 1
n=3
e) sin
n ∞
n=1 4 × 7 × · · · × (3n + 1)
∞ √ h)
n n=1
8 × 11 × · · · × (3n + 5)
f)
n 2 + cos(n)
∞
n=1 arctan(n3 ) 2n (2n)!
∞ i) √ +
3n n + n2 3n (2n + 1)!
n=1
g)
2n + nn
n=1 ∞ sin
√1
n
∞
1 + n(−1)n j) √
h) n=1
n+ n
1 + 2n3
n=1
∞
1 1
∞
2n k) n sin − (n + 2) sin
i) n=1
n n+2
n! + 1
n=1
∞
∞ 2n + 3
n! 1 l)
j) − n n=1
(n + 1)!
(2n)! 2
n=0
∞
∞
arctan(n + 1) − arctan(n)
n (2n)! m)
k) n=1
n2
n=1
4n (n!)2
∞
log(n)
∞
sin(n) n)
l) 1
n 3 log(n) n=1 n sin
n=2 n
∞ ∞
1 n n n+1
m) (−1)n 1 − o) log(2) log
n=1
2n n=1
n
2.7. Solved Exercises 119
∞
1 a) Show that Leibniz’s test cannot be applied.
p) p , p ∈ R
n=3 n log(n) log log(n) b) Show that the series is divergent.
24. The Sierpinski triangle is a geometric figure con- to the sides of the triangle and to the circles that
structed as follows: A square is divided into four precede and follow it in the sequence. Find the sum
equal squares, and the upper right square is removed. of the areas of the circles.
Each of the remaining three squares is divided in the
same way, and the upper right square of each is re-
moved.
SOLUTIONS
∞
1
1. a) Let an be the general term of the series . Since n2 + 3n + 2 = (n + 1)(n + 2), we get
n=1
n2 + 3n + 2
1 A B
an = = +
n2 + 3n + 2 n+1 n+2
A(n + 2) + B(n + 1)
. =
(n + 1)(n + 2)
(A + B)n + 2A + B
= ,
(n + 1)(n + 2)
from which
A+B =0 A=1
. ⇔
2A + B = 1 B = −1.
Therefore,
1 1
.an = − .
n+1 n+2
1
The series is telescopic with αn = and k = 1, using the notation of formula (2.1) on page 74. In
n+1
this case, lim αn = 0, so we can conclude that the series converges and, using formula (2.2) on the page
mentioned above, its sum is
∞
1 1 1
. − = α1 = .
n=1
n + 1 n + 2 2
Alternatively, we can write the sequence of partial sums and determine its limit as follows:
1 1
S1 = −
2 3
1 1 1 1 1 1
S2 = − + − = −
2 3 3 4 2 4
1 1 1 1 1 1
.
S3 = − + − = −
2 4 4 5 2 5
..
.
1 1
Sn = −
2 n+2
.
.
.
1 1
As lim Sn = , the series is convergent and its sum is :
2 2
∞
1 1
. = .
n=1
n2 + 3n + 2 2
122 Numerical Series
∞
+∞
b) Let us consider the series arctan(n + 3) − arctan(n) = − arctan(n) − arctan(n + 3) . It is
n=1 n=1
telescopic with αn = arctan(n) and k = 3, using the notation of formula (2.1) on page 74. In this case,
π
lim αn = , therefore, we can conclude that the series converges, and, using the formula (2.2), its sum is
2
∞
π 5π
. − (arctan(n) − arctan(n + 3)) = − α1 + α2 + α3 − 3 · = − arctan(2) − arctan(3).
n=1
2 4
An alternative path to solve this problem is to write the sequence of partial sums and find its limit:
π
S1 = arctan(4) − arctan(1) = arctan(4) −
4
π
S2 = arctan(4) − + arctan(5) − arctan(2)
4
π
S3 = arctan(4) − + arctan(5) − arctan(2) + arctan(6) − arctan(3)
4
π
S4 = arctan(4) − + arctan(5) − arctan(2) + arctan(6) − arctan(3) + arctan(7) − arctan(4)
4
π
= − + arctan(5) − arctan(2) + arctan(6) − arctan(3) + arctan(7)
4
π
. S5 = − + arctan(5) − arctan(2) + arctan(6) − arctan(3) + arctan(7) + arctan(8) − arctan(5)
4
π
= − − arctan(2) + arctan(6) − arctan(3) + arctan(7) + arctan(8)
4
π
= − − arctan(2) − arctan(3) + arctan(6) + arctan(7) + arctan(8)
4
.
.
.
π
Sn = − − arctan(2) − arctan(3) + arctan(n + 1) + arctan(n + 2) + arctan(n + 3)
4
.
..
π 3π 5π
As lim Sn = − − arctan(2) − arctan(3) + = − arctan(2) − arctan(3), the series is convergent
4 2 4
and
+∞
5π
. arctan(n + 3) − arctan(n) = − arctan(2) − arctan(3).
n=1
4
∞
1 1 1
c) The series n
− n+1
is telescopic, with αn = n and k = 1, using the notation of formula (2.1)
n=0
e e e
1
on page 74. As lim n = 0, the series is convergent. The sum of the series is
e
∞
1 1 1 1
. − = 0 − lim n = 1.
n=0
en en+1 e e
1 1
Another process to find the sum is to write the general term in the form 1− , highlighting the fact
en e
1 1
that this series is geometric with ratio r = . As 0 < < 1, the series converges and:
e e
∞ ∞ ∞
1 1 1 1 1 1 1 1
. − = 1 − = 1 − = 1 − = 1.
n=0
en en+1 n=0
en e e n=0 en e 1− 1
e
2.7. Solved Exercises 123
∞
1 1 1
d) The series − is telescopic, with αn = and k = 1, using the
n=1
arctan(n) arctan(n + 1) arctan(n)
2
notation of formula (2.1) on page 74. In this case, lim αn = . Therefore, we can conclude that the series
π
converges and, using formula (2.2), has sum
∞
1 1 1 2 4 2 2
. − = − = − = .
n=1
arctan(n) arctan(n + 1) arctan(1) π π π π
Alternatively, we can write the sequence of partial sums and find its limit:
1 1 4 1
S1 = − = −
arctan(1) arctan(2) π arctan(2)
4 1 1 1 4 1
S2 = − + − = −
π arctan(2) arctan(2) arctan(3) π arctan(3)
4 1 1 1 4 1
S3 = − + − = −
. π arctan(3) arctan(3) arctan(4) π arctan(4)
.
..
4 1
Sn = −
π arctan(n + 1)
.
.
.
As
4 1 4 2 2
. lim Sn = lim − = − = ,
π arctan(n + 1) π π π
the series is convergent, and its sum is
∞
1 1 2
. − = .
n=1
arctan(n) arctan(n + 1) π
∞
∞
1 1
e) The following two series and are geometric series. To see this, observe that:
n=1
22n−1 n=1
23n−2
∞
∞
∞ ∞
1 1 1 1
. = = 2 = 2
n=1
22n−1 n=1
22n · 2−1 n=1
22n
n=1
4n
and
∞
∞
∞ ∞
1 1 1 1
. = = 4 = 4 .
n=1
23n−2 n=1
23n · 2−2 n=1
23n
n=1
8n
1 1
The first series has ratio r = , and the second has ratio r = . The series are convergent as in both cases
4 8
|r| < 1. The sum of the series can be calculated as follows:
∞
1 ∞
1
1 2 1 4
. =2· 4
= and =4· 8
= .
n=1
22n−1 1− 1
4
3 n=1
23n−2 1− 1
8
7
∞
1 1 2 4 2
By Theorem 2.2.2 the series − is convergent, and its sum is equal to − = .
n=1
22n−1 23n−2 3 7 21
124 Numerical Series
∞ ∞ ∞ ∞
3n + 7n 1 1 1 1
f) The series n · 7n
can be written in the form n
+ n
. Both n
and n
are
n=0
3 n=0
7 3 n=0
7 n=0
3
1 1
geometric series with ratios r = and r = , respectively.
7 3
The series are convergent as in both cases |r| < 1. Using the formula for the sum of a convergent geometric
series, we get
∞ ∞
1 1 7 1 1 3
. = = and = = .
n=0
7n 1 − 1
7
6 n=0
3n 1 − 1
3
2
∞
3n + 7n
By applying Theorem 2.2.2, we can conclude that the given series is convergent, and its sum
n=0
3n · 7n
7 3 8
is equal to + = .
6 2 3
∞
1 2 3 1
g) The series n−1
· n · n+1 is geometric with ratio r = because
n=0
4 5 6 120
1 2 3 6·4 4
. · · = n n n = .
4n−1 5n 6n+1 4 ·5 ·6 ·6 120n
Thus, we can calculate the sum as follows:
∞
∞
1 2 3 1 1 480
. · · = 4 =4· = .
n=0
4n−1 5n 6n+1 n=0
120 n 1 − 1
120
119
1
Thus the series is telescopic, with αn = and k = 2, using the notation of formula (2.1) on page 74.
log(n)
1
The series is convergent because lim = 0, and
log(n)
∞
1 1 1 1
. − = + .
n=2
log(n) log(n + 2) log(2) log(3)
∞
∞
1 1
2. By reindexing the series, we can write = . Additionally,
n=2
(n + 1)2 n=3
n2
∞ ∞
1 1 1
.
2
=1+ + 2
.
n=1
n 4 n=3
n
∞ ∞
1 π2 1 π2 1 π2 5
Knowing that 2
= , then 2
= −1− = − , that is,
n=1
n 6 n=3
n 6 4 6 4
∞
1 π2 5
.
2
= − .
n=2
(n + 1) 6 4
2.7. Solved Exercises 125
∞
1 1
The series n
is convergent because it is geometric with ratio . By Theorem 2.2.2, the sum of these two
n=2
π π
series is convergent and
∞
∞
∞ 1
1 1 1 1 π2 5 2 π2 5 1
. + n = + = − + π 1 = − + :
n=2
(n + 1)2 π n=2
(n + 1) 2
n=2
π n 6 4 1 − π
6 4 π(π − 1)
3. a) The rational number 0, 5 = 0, 55555 . . . can be written in the form of a fraction using geometric series. By
writing
5 5 5 5
.0, 5 = 0, 5 + 0, 05 + . . . = + 2 + 3 + 4 + ···
10 10 10 10
1 5
it is clear that we have a geometric series with ratio r = and first term . Therefore,
10 10
5
10 5
.0, 5 = = .
1− 1
10
9
b) As in the previous item, we can use geometric series to write the rational number 0, 34 = 0, 343434 . . . in
the form of a fraction. By writing
34 34 34 34
.0, 34 = 0, 34 + 0, 0034 + . . . = + 4 + 6 + 8 + ···
102 10 10 10
1 34
it is evident that we have a geometric series with ratio r = and first term . Therefore,
102 102
34
102 34
.0, 34 = = .
1− 1
102
99
∞
2 n2 − 1
4. a) The series is of positive terms. To study this series, we will compare it with the series
n=1
3 n5 + 2 n + 1
∞
1
, which is convergent because it is a p-series with p = 3:
n=1
n3
2 n2 − 1
3 n5 + 2 n + 1 (2 n2 − 1) n3 2 n5 − n3 2
. lim = lim = lim = .
1 5
3n + 2n + 1 3 n5 + 2 n + 1 3
n3
2
Since ∈ R+ , according to Corollary 2 of the General Comparison Test, we can conclude that the given
3
∞
2 n2 − 1
series 5 + 2n + 1
is convergent.
n=1
3 n
∞
2 π
1 1
b) The series sin is of positive terms because ∈ ]0, 1] ⊂ 0, , ∀n ∈ N. To determine its
n=1
n n 2
∞
1
convergence, we compare it with the series 2
, which we know is convergent as it is a p-series with
n=1
n
p = 2. We calculate the limit
2 ⎡ ⎤2
1 1
sin sin
n ⎢ n ⎥
. lim ⎢
= lim ⎣ ⎥ = 1,
1 1 ⎦
n2 n
which belongs to R+ . Therefore, by Corollary 2 of the General Comparison Test, the given series is conver-
gent.
∞
log(n7 + 1)
c) The series is a series of positive terms. We can analyze this series by comparing it with
n=1
n2
∞
1 3
the series 3/2
, which is convergent as it is a p-series with p = :
n=1
n 2
log(n7 + 1)
n2 n3/2 log(n7 + 1) log(n7 + 1)
. lim = lim = lim = 0.
1 n 2 n1/2
n 3/2
∞
1
Since the series 3/2
is convergent, the fact that this limit is zero allows us to conclude, by Corollary 3
n=1
n
∞
log(n7 + 1)
of the General Comparison Test, that the series is also convergent.
n=1
n2
log(n7 + 1)
Note: For calculating the limit lim , we used L’Hôpital’s Rule applied to the calculation of
n1/2
log(x7 + 1) ∞
lim , where the indeterminate form of type ∞
arises:
x→+∞ x1/2
2.7. Solved Exercises 127
7x6
√
log(x7 + 1) x7 +1 14x6 x
. lim = lim = lim = 0.
x→+∞ (x1/2 ) x→+∞ 1 x→+∞ x7 + 1
√
2 x
is finite
and different from zero; therefore, by Corollary 2 of the General Comparison Test, the series
∞
1 √
n + − n is convergent.
n=1
n
∞ √ √
n+1− n
e) The series √ √ is of positive terms. Considering that
n=1
n+1+ n
√ √ √ √ √ √
n+1− n n+1− n n+1+ n 1
.√ √ = √ √ 2 = √ √ 2 ,
n+1+ n n+1+ n n+1+ n
∞
1
we can write the series in the form √ √ 2 . We will analyze this series by comparing it with
n=1 n+1+ n
the harmonic series:
1
√ √ 2
n+1+ n n 1
. lim = lim √ √ 2 = .
1 n+1+ n 4
n
1
Since ∈ R+ , we can conclude, by Corollary 2 of the General Comparison Test, that the given series
√ 4 √
∞
n+1− n
√ √ is divergent.
n=1
n+1+ n
128 Numerical Series
∞
cos(n)
f) The series 2+4
is not of positive terms. However, by bounding the modulus of the general term, we
n=1
n
can obtain the following inequality:
! !
! cos(n) ! 1 1
.0 < !! 2 !≤ ≤ 2 , ∀n ∈ N. (2.5)
n + 4! n2 + 4 n
∞
1
It is worth noting that the series 2
is convergent since it is a p-series with p = 2. Using the inequality
n=1
n
∞ ! !
! cos(n) !
(2.5), we can apply the General Comparison Test to show that the series ! !
! n2 + 4 ! converges. Therefore,
n=1
∞
cos(n)
the original series 2+4
is absolutely convergent.
n=1
n
g) Recall that −1 ≤ cos(n) ≤ 1, ∀n ∈ N. Using this fact, we obtain n3 + 10 cos(n) ≥ n3 − 10, ∀n ∈ N. Thus,
we have that
.n
3
+ 10 cos(n) ≥ n3 − 10 > 0, ∀n ≥ 3.
1 1
.0 < ≤ 3 , ∀n ≥ 3. (2.6)
n3 + 10 cos(n) n − 10
∞
1
The series is of positive terms, and we can study its convergence by comparing it with the
n=3
n3 − 10
∞
1
series 3
, which we know to be convergent. Since
n=3
n
1
n3 − 10 n3
. lim = lim 3 =1
1 n − 10
n 3
∞
1
is a positive real number, the series 3 − 10
is also convergent. Using inequality (2.6) and the General
n=3
n
∞
1
Comparison Test, we can conclude that the series 3 + 10 cos(n)
is convergent. This series differs from
n=3
n
∞
1
the given series only in a finite number of terms; therefore, the series converges.
n=1
n3 + 10 cos(n)
∞
1 − (−1)n
h) Consider the series √ , where an is its general term. We can express an as follows:
n=1
n
⎧
⎪
⎨0, if n is even
.an = 2
⎪
⎩√ , if n is odd.
n
2.7. Solved Exercises 129
∞ ∞
1 − (−1)n 2
Hence, √ = √ . This is a series of positive terms. Let us now consider the series
n=1
n n=1
2n −1
∞
1 1
√ , which is divergent because it is a p-series with p = . The limit
n=1
n 2
2
√ √
2n − 1 2 n 4n √
. lim = lim √ = lim = 2
1 2n − 1 2n − 1
√
n
belongs to R+ ; therefore, by Corollary 2 of the General Comparison Test, the given series is divergent.
∞
∞
log(n) log(n)
5. a) The series (−1)n = (−1)n is alternating. To begin, let us study the series of modules
n=1
n n=2
n
∞
log(n)
, comparing it with the harmonic series. We can see that
n=2
n
log(n)
n
. lim = lim log(n) = +∞.
1
n
This result allows us to conclude, by Corollary 3 of the General Comparison Test, that the series of modules
is divergent. Since the initial series is alternating, let us check if we can apply Leibniz’s test:
log(n) log(x) ∞
(i) lim = 0 (notice that lim is an indeterminate form of type that can be solved
n x→+∞ x ∞
(log(x)) 1
using L’Hôpital’s Rule: lim = lim = 0).
x→+∞ (x) x→+∞ x
log(n)
(ii) > 0, ∀n > 1.
n
log(x) 1 − log(x)
(iii) The function f (x) = is decreasing on ]e, +∞[. Indeed, f (x) = < 0 ⇔ x > e,
x x2
log(n)
which implies that is decreasing for n ≥ 3.
n
Therefore, we can state that the alternating series is convergent. As the series of modules is divergent, the
given series is conditionally convergent, taking into account Definition 2.4.1.
∞ ∞
log(n) log(n)
Note: The series and are both divergent since they differ only in a finite number of
n=2
n n=3
n
terms.
∞
2 2
b) The series (−1)n+1n + e−n
is alternating, given that n > 0, ∀n ∈ N. To begin with, let us
n=1
e e + e−n
study the series of modules:
∞ ! ! ∞
! 2 ! 2
!(−1)n+1 != .
. ! e n + e −n ! e n + e−n
n=1 n=1
130 Numerical Series
We have
2 2
.0 < < n , ∀n ∈ N. (2.7)
en + e−n e
∞ ∞ n
1 1 1 1
The series n
= is geometric with ratio r = . It is convergent because |r| = < 1. By
n=1
e n=1
e e e
∞
2
Theorem 2.2.2, the series is also convergent. Using the General Comparison Test and considering
n=1
en
∞
2
inequality (2.7), the series n + e−n
is convergent. As this is the series of modules of the given series,
n=1
e
by Definition 2.4.1, we can conclude that the initial series is absolutely convergent.
∞
1 1
c) The series (−1)n is alternating because an = > 0, ∀n ∈ N. To study the convergence
n=1
log(3n) log(3n)
∞
1
of the series, let us first consider the series of modules, . We can compare this series with the
n=1
log(3n)
harmonic series: 1
log(3n) n
. lim = lim = +∞.
1 log(3n)
n
∞
1
It follows by Corollary 4 of the General Comparison Test that the series is divergent. Since the
n=1
log(3n)
initial series is alternating, we can check the conditions of Leibniz’s test to see if it converges:
1
(i) lim = 0.
log(3n)
1
(ii) > 0, ∀n ≥ 1.
log(3n)
1
(iii) The sequence log(3n) is increasing; therefore, is a decreasing sequence.
log(3n)
According to Leibniz’s test, the alternating series is convergent. However, as the series of its modules is
divergent, the given series is conditionally convergent, considering Definition 2.4.1.
∞ 1 1
sin sin 1 π
d) The series (−1)n n
is alternating, because an = n
> 0, ∀n ≥ 1 (recall that 0 < < ,
n n n 2
n=1
∞ 1 ∞
sin 1
∀n ≥ 1). The series of modules is n
. Consider the convergent series 2
, which is a p-series
n=1
n n=1
n
with p = 2. The limit
1
sin n 1
sin
n n
. lim = lim =1
1 1
n2 n
is finite and different from zero; therefore, by Corollary 2 of the General Comparison Test, the series of
modules is convergent; thus, considering Definition 2.4.1, we can conclude the given series is absolutely
convergent.
2.7. Solved Exercises 131
∞
3n 3n
e) The series (−1)n 2n+1
is alternating because 2n+1 > 0, ∀n ∈ N. To analyze the absolute convergence
n=1
e e
of the series, we consider the series of modules:
∞ ! !
∞ ∞ n ∞
!
!(−1)n 3
n !
!= 3n 1 3 1 3 n
. ! 2n+1 ! 2n+1
= 2
= 2
.
n=1
e n=1
e n=1
e e e n=1 e
∞
3 n 3
The series 2
is geometric with ratio r = 2 . Since |r| < 1, the series is convergent. We know by
n=1
e e
Theorem 2.2.2 that the product of a constant by a convergent series is convergent; therefore, the series of
modules is convergent. Then, the given series is absolutely convergent, again considering Definition 2.4.1.
f) Considering that ⎧
⎨0, if n is even
nπ
. sin =
2 ⎩(−1) n−1
2 , if n is odd,
∞ nπ
sin 2
we can write the series √ as
n=1
n+1
∞ ∞
(−1)n−1 1 (−1)n−1
. √ = √ √ .
n=1 2n 2 n=1 n
∞
1
This series is alternating, and its series of modules is √ , which is divergent. We will now apply
n=1
n
Leibniz’s test to determine the convergence of our given series:
1
(i) lim √ = 0.
n
1
(ii) √ > 0, ∀n ∈ N.
n
1 1 1
(iii) √ − √ > 0, ∀n ∈ N, and therefore, the sequence √ is decreasing.
n n+1 n
Thus, the alternating series is convergent. However, since the series of absolute values is divergent, we know
the given series is conditionally convergent.
∞
∞
1 1
6. a) The series (−1)n+1 is alternating. The series of modules is the harmonic series, which is
n=1
n n=1
n
divergent.
To determine the convergence of the alternating series, let us apply Leibniz’s test:
1
(i) lim = 0.
n
1
(ii) > 0, ∀n ∈ N.
n
1 1 1
(iii) − = > 0, ∀n ∈ N, and therefore, the sequence is decreasing.
n n+1 n(n + 1)
Thus, the alternating series is convergent. As its series of modules is divergent, the given series is conditionally
convergent.
132 Numerical Series
b) From the Corollary of Theorem 2.3.2, we know that |S − Sn | ≤ an+1 . It is sufficient to require that
1
an+1 < , that is, n > 999. Therefore, the intended sum is S1000 .
1000
c) By the corollary referred to in the previous item, |S − S5 | ≤ a6 = 1
6
, and this value is an upper bound of
the error.
∞
nn
7. a) The series is of positive terms. Let an be the general term of the series. We can find the limit
n=1
π n n!
of the ratio of successive terms as follows:
(n + 1)n+1
an+1 π n+1 (n + 1)! (n + 1)n+1 π n n!
lim = lim n
= lim n n+1
an n n π (n + 1)!
.
π n n!
(n + 1)(n + 1)n n! 1 n+1 n e
= lim n
= lim = .
n π (n + 1) n! π n π
Since this value is less than 1, D’Alembert’s test guarantees that the series is convergent.
∞
n 2n
b) The series 3+1
is of positive terms. Let an be the general term of the series. We will take the
n=0
4 n
limit of the ratio of successive terms of the series:
(n + 1)2n+1
an+1 4(n + 1)3 + 1 (n + 1) 2n+1 (4n3 + 1)
lim = lim n
= lim
an n2 n 2n 4(n + 1)3 + 1
.
4n3 + 1
2(n + 1)(4n3 + 1) 2 (n + 1) (4n3 + 1)
= lim = lim = 2.
3
n 4(n + 1) + 1 n (4n3 + 12n2 + 12n + 5)
As this limit has a value greater than 1, the series is divergent by D’Alembert’s test.
c) Noting that
∞ ∞ ∞
(n + 1)! − n! (n + 1)n! − n! n
. = = ,
n=1
n! (n + 1)! n=1
n! (n + 1)! n=1
(n + 1)!
let an be the general term of the series. As
n+1
an+1 (n + 2)! (n + 1)(n + 1)! n+1
. lim = lim = lim = lim = 0 < 1,
an n n(n + 2)! n(n + 2)
(n + 1)!
it follows from D’Alembert’s test that the series is convergent.
∞
2 · 4 · 6 · · · · · (2n + 2)
d) The series is of positive terms. Let an be the general term of the series. We
n=1
1 · 4 · 7 · · · · · (3n + 1)
have
2 · 4 · 6 · · · · · (2n + 2)(2n + 4)
an+1 1 · 4 · 7 · · · · · (3n + 1)(3n + 4) 2n + 4 2
. lim = lim = lim = .
an 2 · 4 · 6 · · · · · (2n + 2) 3n + 4 3
1 · 4 · 7 · · · · · (3n + 1)
2.7. Solved Exercises 133
As this value is less than 1, we can conclude by D’Alembert’s test that the series is convergent.
∞
(n + 1)!
e) The series √ is of positive terms. Let an be the general term of the series. Evaluating
n=1
nn 3n + 2
an+1
lim , we obtain
n→∞ an
(n + 2)!
√ √
an+1 (n + 1)n+1 3n + 5 (n + 2)! nn 3n + 2
lim = lim = lim √
an (n + 1)! (n + 1)! (n + 1)n+1 3n + 5
. n
√
n 3n + 2
√ n
(n + 2) nn 3n + 2 n + 2 3n + 2 n 1
= lim n+1
√ = lim = .
(n + 1) 3n + 5 n + 1 3n + 5 n + 1 e
Since this value is less than 1, the series is convergent according to D’Alembert’s test.
∞
3n + n!
f) The series is of positive terms. It is easy to prove by induction that 3n < n!, ∀n ≥ 7, which
n=1
n! + nn
implies that
3n + n! 2 n!
. ≤ n , ∀n ≥ 7.
n! + nn n
∞
2 n!
Let an be the general term of the series . D’Alembert’s test confirms that the series converges
n=1
nn
since
2 (n + 1)!
n
an+1 (n + 1)n+1 (n + 1)! nn n 1
. lim = lim = lim = lim = < 1.
an 2 n! n! (n + 1)n+1 n+1 e
nn
∞
3n + n!
By the General Comparison Test, we can conclude that the series also converges.
n=1
n! + nn
Note: Two series that differ only in a finite number of terms are both convergent or both divergent.
∞
n!
g) The series is of positive terms. We can establish the inequality
n=1
(2 n)! + 2n
n! n!
. ≤ , ∀n ∈ N.
(2 n)! + 2n (2 n)!
∞
n!
Let us investigate the convergence of the series . If an denotes the general term of this series,
n=1
(2 n)!
then
(n + 1)!
an+1 2 (n + 1) ! (n + 1)! (2 n)! n+1
. lim = lim = lim = lim = 0.
an n! n! 2 (n + 1) ! (2n + 2)(2n + 1)
(2 n)!
By D’Alembert’s test, the series is convergent since this limit is less than 1. Applying the General Comparison
∞
n!
Test, the original series is convergent.
n=1
(2 n)! + 2n
134 Numerical Series
∞
∞ ∞
n! + 3n n!
h) The series 2 is of positive terms. Let us consider the series an = 2 and
n=0 (n + 1)! n=0 n=0 (n + 1)!
∞
∞
3n
bn = 2 . If both series are convergent, it follows from Theorem 2.2.2 that the given series
n=0 n=0(n + 1)!
a
is convergent. Let us begin by finding the limit of n+1 a
:
n
(n + 1)!
2 2
an+1 (n + 2)! (n + 1)! (n + 1)! n+1
. lim = lim
n!
= lim 2
= lim = 0 < 1.
an n! (n + 2)! (n + 2)2
2
(n + 1)!
∞
As the limit is less than 1, we can use D’Alembert’s test to conclude that the series an is convergent.
n=0
∞
Similarly, the series bn is convergent:
n=0
3(n + 1)
2 2
bn+1 (n + 2)! (n + 1) (n + 1)! n+1
. lim = lim
3n
= lim 2
= lim = 0 < 1.
bn n (n + 2)! n(n + 2)2
2
(n + 1)!
Since both series are convergent, we can conclude that the given series is also convergent as it is the sum
of two convergent series.
∞
n2
2
8. a) Consider the series of positive terms 1− and let an represent its general term. Since
n=3
n
2
√ 2 n 2 n 1
= e−2 = 2 < 1,
n
. lim n
an = lim 1− = lim 1 −
n n e
∞
∞ n
2
2 1
b) The series e−n = is of positive terms. We have
n=0 n=0
e
n2 n
n 1 1
. lim = lim = 0.
e e
As this value is less than 1, Cauchy’s Root Test ensures the convergence of the series.
∞
√
n n
c) The series 2−1 is of positive terms. We find that:
n=1
√
n n √
n
. lim
n
2−1 = lim 2 − 1 = 0,
√
since lim n 2 = 1. As this value is less than 1, Cauchy’s Root Test allows us to conclude that the series is
convergent.
2.7. Solved Exercises 135
∞
n
π 1 π 1 π
d) The series cos + is of positive terms because 0 < + < , ∀n ∈ N. Applying
n=1
6 n 6 n 2
Cauchy’s Root Test, we conclude that the series is convergent. In fact:
n √
n π 1 π 1 π 3
. lim cos + = lim cos + = cos = < 1.
6 n 6 n 6 2
∞ n
1 n
e) The series + (−1)n is of positive terms. The sequence of general term
n=0
2 4n + 1
n
1 n 1 n
.
n
+ (−1)n = + (−1)n
2 4n + 1 2 4n + 1
3 1
has two sublimits: and . Since the upper limit is less than 1, we can apply Corollary 1 of the Root Test
4 4
to conclude that the series is convergent.
∞ 2
n ∞ 2 n ∞ 2 n
n −2 1 n −2 n −2
f) We can write + =− . The series is of positive terms
n=1
3n2 3 n=2 3n 2
n=2
3n 2
and differs from the given series in only one term. We can state that the two series are either both convergent
or both divergent. Using Cauchy’s Root Test, we can show that the series converges because
n
n n2 − 2 n2 − 2 1
. lim = lim = < 1.
3n2 3n2 3
∞
(−1)n 1
9. a) The series n)
is alternating because an = n)
> 0, ∀n ≥ 2; its general term can be written
n=2
log(n log(n
∞ ! ! ∞
(−1)n ! (−1)n ! 1
in the form . Let us start by studying the series of modules, ! ! = ,
n log(n) ! n log(n) ! n log(n)
n=2 n=2
1
applying the Cauchy’s Condensation Test. The sequence is obviously decreasing. The
n log(n) n∈N\{1}
series
∞ ∞ ∞
2n 1 1 1
.
n n
= =
n=2
2 log(2 ) n=2
n log(2) log(2) n=2 n
is divergent as it is the product of a constant by the harmonic series. Then, the series of modules is divergent
by Cauchy’s Condensation Test. As a result, if convergent, the series under study is conditionally convergent.
Since the series is alternating and satisfies the following conditions:
1
(i) >0
n log(n)
1
(ii) lim =0
n log(n)
1
(iii) As mentioned earlier, the sequence is decreasing
n log(n) n∈N\{1}
it follows from the Leibniz’s test that the alternating series is convergent. From the study of the series of
modules, we can state that the series is conditionally convergent, considering Definition 2.4.1.
136 Numerical Series
n n
n2 2 n2 2 n4
. lim √ + 1+ = lim √ + lim 1 + = lim + e2 = e2 .
n5 +1 n n5 + 1 n n5 + 1
As the general term of the series does not converge to zero, the series is divergent (refer to Theorem 2.2.1).
1 1
c) First, we note that 0 < 1 − 2
< 1, ∀n ≥ 2, which implies that log 1 − 2 < 0, ∀n ≥ 2. Thus, the
n n
series is of negative terms. Next, we consider the series of modules
∞ ! ! ∞
! !
!log 1 − 1 ! = 1
− log 1 − 2 .
. ! n 2 ! n
n=2 n=2
∞
1
We can compare this series with the series 2
, which is convergent because it is a p-series with p = 2.
n=2
n
The limit
1
− log 1 − 2 2
n 1 1 n
. lim = − lim n2 log 1 − 2 = − lim log 1 − 2 = − log(e−1 ) = 1
1 n n
n2
is finite and different from zero. Therefore, the series of modules is convergent by Corollary 2 of the
General Comparison Test. Then, by Theorem 2.4.1, the given series is absolutely convergent, considering
Definition 2.4.1.
∞
sin(n) + 2n
d) The series is of positive terms because −1 ≤ sin(n) ≤ 1 and 2n > 1, ∀n ∈ N. The
n=1
n + 5n
previous inequalities allow us to obtain the following estimate:
n n
sin(n) + 2n 1 + 2n 1 2n 1 2
.0 < n
< n
= n + n = + , ∀n ∈ N.
n+5 5 5 5 5 5
∞ n
∞ n
1 1 2
The series is convergent because it is geometric with ratio r = , and the series is
n=1
5 5 n=1
5
2
convergent because it is geometric with ratio r = . Then, the series
5
∞ n n
1 2
. +
n=1
5 5
is convergent by Theorem 2.2.2. By the General Comparison Test, the original series is convergent. As it is
a series of positive terms, it is absolutely convergent.
∞
en+1 en+1
e) The series (−1)nn
is alternating because an = > 0, ∀n ∈ N. To determine the convergence
n=1
n nn
∞
en+1
of the series, we first examine the series of modules, , using D’Alembert’s test:
n=1
nn
2.7. Solved Exercises 137
en+2
n
an+1 (n + 1)n+1 en+2 nn e n 1
. lim = lim = lim = lim = 0 × = 0.
an en+1 en+1 (n + 1)n+1 n+1 n+1 e
nn
As this value is less than 1, the series of modules is convergent. Therefore, the given series is absolutely
convergent by Theorem 2.4.1, considering Definition 2.4.1.
1 1
∞ (−1)n +
∞ (−1)n +
f) To study the convergence of the series n = n , we will start by rewriting its
n=1
n n=2
n
general term:
1
(−1)n + n
.
n = (−1)n · n + (−1) .
n n2
n + (−1)n
Let an = . It is evident that an > 0, ∀n ∈ N \ {1}, and therefore the series under study is
n2
∞
n + (−1)n
alternating. Let us study the series of modules, , by comparison with the harmonic series.
n=1
n2
The limit
n + (−1)n
n2 n(n + (−1)n ) n + (−1)n
. lim = lim = lim =1
1 n 2 n
n
belongs to R+ . As the harmonic series is divergent, the series of modules is also divergent, by Corollary 2
of the General Comparison Test.
1
∞
(−1)n ∞
1 (−1)n + (−1)n 1
Let us consider the series and . Taking into account that n = + 2
n n 2 n n n
n=1 n=1
and that the two previous series are convergent (the first one is the alternating harmonic series, and the
1
∞ (−1)n +
second one is a p-series with p = 2), the series n is convergent by Theorem 2.2.2. From
n=1
n
the study of the series of modules, we can affirm that the series is conditionally convergent, according to
Definition 2.4.1.
n + (−1)n
Note: The Leibniz’s Test is not applicable because the sequence is not monotonic (see
n2
Fig. 2.6).
1 1
g) Bearing in mind that 1 + n > 1, ∀n ≥ 1, we conclude that log 1 + n > 0, ∀n ≥ 1. Thus, the
2 2
∞ ∞
1 1
series log 1 + n is of positive terms. Let us compare this series with the series n
, which is
n=1
2 n=1
2
1
convergent as it is a geometric series with ratio r = . The limit
2
1
log 1 + n n
2 1 1 2
. lim = lim 2n log 1 + n = lim log 1 + n = log(e) = 1
1 2 2
2n
138 Numerical Series
n + (−1)n n + (−1)n
Figure 2.6: The sequences (−1)n and
n2 n2
is finite and different from zero; therefore, the series is convergent by Corollary 2 of the General Comparison
Test. It is absolutely convergent since it is a series of positive terms.
∞
∞
√ n √ n
h) The series 1− n
n = (−1)n n n − 1 is an alternating series. Let us begin by studying the
n=1 n=1
series of modules using Cauchy’s Root Test:
" √ n √
. lim
n n
n − 1 = lim n n − 1 = 0
√
since lim n n = 1. As this value is less than 1, the series of modules is convergent. It follows by Theo-
rem 2.4.1 that the given series is convergent, and by Definition 2.4.1, it is absolutely convergent.
∞
(−1)n
i) To determine the convergence of the series , we first rewrite its general term as
n=1
1 − (−1)n n2
(−1)n 1 −1
. = = 2 .
1 − (−1)n n2 (−1)n − n2 n + (−1)n+1
∞
∞ ∞
(−1)n −1 1
. = =− .
n=1
1 − (−1) n
n 2
n=1
n 2 + (−1) n+1
n=1
n 2 + (−1)n+1
Note that the general term of this last series is positive, whatever n ∈ N. We can compare the series with
∞
1
the series , which we know is convergent because it is a p-series with p = 2. The limit
n=1
n2
1
n2 + (−1)n+1 n2 1
. lim = lim 2 = lim =1
1 n + (−1)n+1 (−1)n+1
1 +
n2 n2
belongs to R+ . Therefore, by Corollary 2 of the General Comparison Test, the series is convergent, and the
convergence is absolute.
2.7. Solved Exercises 139
∞
∞ ∞ ∞ 2n ∞
(−1)n + 4 3
j) Consider the series an = and b n = . The series an is of
n=1 n=1
n + 4n n=1 n=1
n+2 n=1
positive terms; in fact, the expression for an is as follows:
⎧
⎪ 5
⎪
⎨ if n is even
n + 4n
.an =
⎪
⎪ 3
⎩ if n is odd.
n + 4n
∞
5 5
Consequently, an ≤ , ∀n ∈ N. The series is a convergent geometric series because the ratio is
4n n=1
4n
∞
1 5
r = . Since an ≤ n , ∀n ∈ N, the General Comparison Test implies that an is convergent.
4 4 n=1
∞
We can use the Cauchy’s Root Test to study the series of positive terms, bn :
n=1
2n 2
n 3 3
. lim = lim = 0 < 1.
n+2 n+2
∞
This implies that the series bn is convergent.
n=1
∞
∞
Since both an and bn are convergent, we can conclude that the given series is convergent by
n=1 n=1
Theorem 2.2.2. Since it is a series of positive terms, it is absolutely convergent.
∞
1 n
k) The series 1− has a general term that is not a null sequence. Specifically,
n=1
n
n
1
. lim 1− = e−1 .
n
Theorem 2.2.1 can be applied to establish that this series is divergent.
∞
1
l) We will use the Cauchy Condensation Test to determine the convergence of the series 3 .
n log(n)
n=2
1
It is clear that the sequence 3 is decreasing with positive terms. We can apply the Cauchy
n log(n)
Condensation Test by considering the series
∞
∞
∞
2n 1 1 1
. 3 = 3 = 3 3
,
n=2 2n log(2n ) n=2 n3 log(2) log(2) n=2
n
∞
1
which is convergent as it is the product of a constant by a convergent p-series, . Therefore, by the
n=2
n3
∞
1
Cauchy Condensation Test, the series 3 is convergent. Furthermore, since it is a series of
n=2 n log(n)
positive terms, it is absolutely convergent.
140 Numerical Series
1
m) If > 0 for all x = 0, we can say that the integral is positive, which implies that the series consists of
x2
positive terms. As
# $
n+2 1 1 n+2 1 1
. dx = − = − ,
n x2 x n n n+2
we can write the series as
∞
n+2 ∞
1 1 1
. dx = − .
n=1 n x2 n=1
n n+2
1
This series is telescopic with αn = and k = 2, using the notation of formula (2.1) on page 74. The
n
∞
1 1
telescopic series converges if (αn ) is a convergent sequence. As αn → 0, the series − is
n=1
n n+2
convergent. Furthermore, it is absolutely convergent since it consists of positive terms.
∞
∞
(−1)n 1
n) The series √ √ is alternating. The series of modules is the series √ √ ,
n=1
n+1+ 3n+1 n=1
n+1+ 3n+1
∞
1 1
which we can compare with the series 1/2
, which is divergent because it is a p-series with p = . As
n=1
n 2
the value of the limit
1
√ √ √
n+1+ 3n+1 n
. lim = lim √ √ =1
1 n+1+ 3n+1
n1/2
belongs to R+ , Corollary 2 of the General Comparison Test guarantees that the series of modules is divergent.
Let us study the alternating series by applying the Leibniz’s test:
1
(i) lim √ √ = 0.
n+1+ 3n+1
1
(ii) √ √ > 0, ∀n ∈ N.
n+1+ 3n+1
1 √ √
(iii) The sequence √ √ is decreasing, because n + 1 + 3 n + 1 is obviously increas-
n+1+ 3n+1
ing.
The test conditions are satisfied; therefore, the series is convergent. As the series of modules is divergent,
we conclude that the alternating series is conditionally convergent.
∞
1
o) The series √ √ is of positive terms. Let us study this series by comparing it with the
n=1
n n+1+ n
∞
1 3
series , which we know to be convergent, as it is a p-series with p = . The limit
n=1
n3/2 2
1
√ √
n n+1+ n n3/2 1
. lim = lim √ √ =
1 n n+1+ n 2
n3/2
belongs to R+ . It follows from Corollary 2 of the General Comparison Test that the given series is convergent.
It is absolutely convergent since it is a series of positive terms.
2.7. Solved Exercises 141
∞
1 1
p) The series (−1)n 1 − cosis alternating, because 1 − cos > 0, ∀n ≥ 1. The series of
n=1
n n
∞
1
modules is the series 1 − cos . As the value of the limit
n=1
n
1
1 − cos
n 1
. lim =
1 2
n2
∞
1
belongs to R+ . As the series 2
is convergent because it is a p-series with p = 2, then by Corollary 2 of
n=1
n
the General Comparison Test, the series of modules is convergent. Therefore, by Theorem 2.4.1, the given
series is convergent, and it is absolutely convergent (see Definition 2.4.1).
∞
1 1
q) The series (−1)n 2n sin is alternating, because an = 2n sin > 0, ∀n ≥ 1 (note that
n=1
3n 3n
∞
1 π 1
0 < n
< , ∀n ≥ 1). The series of modules is 2n sin n . Considering that if x > 0, then
3 2 n=1
3
sin(x) ≤ x, we get
n n
1 1 2
n
.2 sin ≤ 2n = , ∀n ∈ N.
3n 3 3
∞ n
2 2
The series is geometric with ratio r = ; therefore, it is convergent. According to the General
n=1
3 3
Comparison Test, the series of modules is convergent, so the given series is absolutely convergent.
∞
√ √
n! n n! n
r) The series (−1)n√ is alternating because √ > 0, ∀n ∈ N. Let us study the series
n=1
nn n + 1 nn n + 1
∞ √ √
n! n n! n
of modules, √ , using D’Alembert’s test. Taking a n = √ :
n=1
nn n + 1 nn n + 1
√
(n + 1)! n + 1
√
an+1 (n + 1)n+1 n + 2 (n + 1)! nn (n + 1)
lim = lim √ = lim
an n! n n! (n + 1)n+1 n(n + 2)
. √
nn n + 1
n
(n + 1)2 nn n+1 n 1
= lim = lim = .
(n + 1)n+1 n(n + 2) n(n + 2) n + 1 e
Since this value is less than 1, the series is convergent. As it is the series of modules of the initial series, we
∞ √
n! n
can affirm that the series (−1)n n √ is absolutely convergent.
n=1
n n+1
∞
√ 1 1 π
s) The series ( n + 1) sin 2
is of positive terms because 0 < 2 < . Let us consider the series
n=1
n n 2
142 Numerical Series
∞
1 3
, which we know to be convergent because it is a p-series with p = . The limit
n=1
n3/2 2
√ 1 1
( n + 1) sin sin √
n2 n2 n+1
. lim = lim · √ =1
1 1 n
n3/2 n2
belongs to R+ , and therefore, by Corollary 2 of the General Comparison Test, the series is convergent. As
it is of positive terms, it is absolutely convergent.
∞
1
t) The series is of positive terms. Knowing that n! < nn and that the logarithmic function is
n=2
log(n!)
increasing, we have
1 1 1
.0 < = < , ∀n ≥ 2.
log(nn ) n log(n) log(n!)
∞
1
We proved in exercise 9a) that the series is divergent. By the General Comparison Test, the
n=2
n log(n)
∞
1
series is divergent.
n=2
log(n!)
∞ ∞
n cos(nπ) n cos(nπ) n
10. a) The series √ = √ is alternating because cos(nπ) = (−1)n and √ > 0,
n 3+1 n 3+1 n 3+1
n=0 n=1
∞
n
∀n ∈ N. First, we will examine the series of modules, √ , by comparing it with the series
n=1 n3 + 1
∞
1 1
1/2
, which is divergent because it is a p-series with p = . The limit
n=1
n 2
n √
√
n3 + 1 n3
. lim = lim √ =1
1 3
n +1
n1/2
is finite and different from zero. Therefore, by Corollary 2 of the General Comparison Test, the series
∞
n
√ is divergent.
n 3+1
n=1
Next, we will apply Leibniz’s test to the alternating series:
n
(i) lim √ = 0.
n3 + 1
n
(ii) √ > 0, ∀n ∈ N.
n3 + 1
(iii) To verify that it is a decreasing sequence, consider the real function of real variable
x −x3 + 2
f (x) = √ . The derivative of f is f (x) = √ , and therefore, f (x) < 0,
√ x3 + 1 2(x3 + 1) x3 + 1
∀x > 3 2. The sequence then decreases for n ≥ 2.
Since the test conditions are satisfied, we can conclude that the alternating series is convergent. As the
series of modules is divergent, the given series is said to be conditionally convergent.
2.7. Solved Exercises 143
∞
n!
b) The series is of positive terms. Let us study it using D’Alembert’s test. Let
n=1
3 × 5 × · · · × (2 n + 1)
an be the general term of the series:
(n + 1)!
an+1 3 × 5 × · · · × (2 n + 1)(2 n + 3) n+1 1
. lim = lim = lim = .
an n! 2n + 3 2
3 × 5 × · · · × (2 n + 1)
Since this limit is less than 1, it follows from D’Alembert’s test that the series is convergent. The series is
absolutely convergent, as it is of positive terms.
∞
1 + cos2 (n)
c) The series √ is of positive terms. We have the following inequality
n=1
n
1 1 + cos2 (n)
.√ ≤ √ , ∀n ∈ N.
n n
∞
1 1
Since the series √ is divergent as it is a p-series with p = , the General Comparison Test allows us
n=1
n 2
to conclude that the given series is divergent.
∞
3n (n + 1)!
d) The series is of positive terms. Let us study it using D’Alembert’s test. Let an be the
n=0
(n + 1)n
general term of the series:
3n+1 (n + 2)!
an+1 (n + 2)n+1 n+1 n 3
. lim = lim n = 3 lim = .
an 3 (n + 1)! n+2 e
(n + 1)n
This limit value is greater than 1; therefore, by D’Alembert’s test, the series is divergent.
∞
(−1)n
e) We know that the function sin(x) is odd; therefore, the series sin is alternating and can be
n=1
n
∞ ∞
1 1
written in the form (−1)n sin . Let us begin by studying the series of modules, sin by
n=1
n n=1
n
comparing it with the harmonic series. The limit
1
sin
n
. lim =1
1
n
belongs to R+ . It follows from Corollary 2 of the General Comparison Test that the series of modules is
divergent.
Let us study the alternating series by applying Leibniz’s test:
1
(i) lim sin = 0.
n
1
(ii) sin > 0, ∀n ∈ N.
n
144 Numerical Series
1 π
(iii) The sequence is decreasing, the sine function is increasing on [0, 2
], and therefore, the sequence
n
1
sin is decreasing.
n
We conclude that the alternating series is convergent. Since the series of modules is divergent, the alternating
series is conditionally convergent.
∞
√
n
f) The series is of positive terms. The limit
n=1
n2 + cos(n)
√
n
n2 + cos(n) n2
. lim = lim 2 =1
1 n + cos(n)
n3/2
∞
1 3
is finite and different from zero. The series 3/2
is convergent, since it is a p-series with p = .
n=1
n 2
The General Comparison Test ensures that the original series is convergent. As it is of positive terms, it
converges absolutely.
∞
3n 3n 3n
g) The series is of positive terms, and the inequality ≤ n holds in N. Let us study
n=1
2n + nn 2n + nn n
∞
3n
the series n
using the Cauchy’s Root Test. Let an be the general term of the series:
n=1
n
√ 3
. lim n an = lim = 0.
n
Since this limit is less than 1, the series is convergent. Using the General Comparison Test, we can affirm
that the series under study is convergent and it is absolutely convergent since it is of positive terms.
∞
1 + n(−1)n n + (−1)n
h) The series 3
is alternating, since we can write its general term as (−1)n . Let
n=1
1 + 2n 1 + 2n3
∞
n + (−1)n
us start by studying the series of modules, . The limit
n=1
1 + 2n3
n + (−1)n
1 + 2n3 n3 + (−1n )n2 1
. lim = lim =
1 2n3 + 1 2
n2
∞
1
is finite and different from zero. Since the series 2
is a p-series with p = 2, it converges. Then,
n=1
n
∞
n + (−1)n
Corollary 2 of the General Comparison Test allows us to conclude that the series is con-
n=1
1 + 2n3
vergent. Therefore, by Theorem 2.4.1, the alternating series is convergent, and it converges absolutely (see
Definition 2.4.1).
2.7. Solved Exercises 145
∞
2n 2n 2n
i) The series is of positive terms, and the inequality ≤ holds in N. Let us study the
n=1
n! + 1 n! + 1 n!
∞
2n
series using D’Alembert’s test. Let an be the general term of the series:
n=1
n!
2n+1
an+1 (n + 1)! 2
. lim = lim = lim = 0.
an 2n n+1
n!
This limit has a value less than 1; therefore, according to D’Alembert’s test, the series is convergent. By
the General Comparison Test, the initial series is convergent and absolutely convergent as it is of positive
terms.
∞
∞
n! 1
j) Let us consider the series and n
. The first one, which is of positive terms, can be studied
n=0
(2n)! n=0
2
using D’Alembert’s test. Let an be the general term of the series:
(n + 1)!
an+1 (2n + 2)! (n + 1)!(2n)! n+1
. lim = lim = lim = lim = 0.
an n! n!(2n + 2)! (2n + 2)(2n + 1)
(2n)!
This limit is less than 1; therefore, by D’Alembert’s test, the series is convergent. As it is of positive terms,
it is absolutely convergent.
1
The second series is geometric with ratio r = ; therefore, it is convergent. Again, as it is of positive terms,
2
it converges absolutely.
∞ ∞
n! 1
The original series is the sum of the series and − n , so it is convergent and absolutely
n=0
(2n)! n=0
2
convergent (consider the inequality: |a − b| ≤ |a| + |b|, ∀a, b ∈ R).
∞
n (2n)!
k) The series n (n!)2
is of positive terms. Let an be the general term of the series,
n=1
4
(n + 1)(2n + 2)!
2
an+1 4n+1 (n + 1)! (n + 1)(2n + 2)! 4n (n!)2
lim = lim = lim 2
an n(2n)! n(2n)! 4n+1 (n + 1)!
4n (n!)2
. (n + 1)(2n + 2)(2n + 1) 2(n + 1)2 (2n + 1)
= lim 2
= lim
4 n(n + 1) 4 n(n + 1)2
2n + 1
= lim = 1,
2n
2n + 1
and therefore, D’Alembert’s test is inconclusive. The sequence tends to 1 by values greater than
2n
1, from which we conclude, by the Ratio Test, that the series is divergent.
146 Numerical Series
Note: We can also, alternatively, study the series using Raabe’s test (a common procedure when D’Alembert’s
test is inconclusive). In this case,
an 2n −n 1
. lim n − 1 = lim n − 1 = lim =− .
an+1 2n + 1 2n + 1 2
∞
1
l) The modulus of the general term of the given series is bounded by the general term of the series ,
n=2
n3 log(n)
since ! !
! sin(n) ! 1
. !! !≤ , ∀n ∈ N \ {1}.
n3 log(n) ! n3 log(n)
∞
1
As the series is convergent as it is a p-series with p = 3 and
n=1
n3
1
n3 log(n) 1
. lim = lim = 0,
1 log(n)
n3
∞
1
Corollary 3 of the General Comparison Test allows us to conclude that the series is convergent.
n3 log(n) n=2
Consequently, the series of modules is convergent, making the given series absolutely convergent.
∞
n
1
m) The general term of the series (−1)n 1− is not a null sequence. In fact,
n=1
2n
n 2n 1
2
1 1
. lim 1− = lim 1− = e−1/2 ,
2n 2n
n
1
which implies that the sequence of general term (−1)n 1− does not have a limit. From Theo-
2n
rem 2.2.1, it follows that the series is divergent.
∞
∞
1 1 1
n) The series · arcsin is of positive terms. Consider the series 2
, which we know to be
n=1
2n + 5 n n=1
n
convergent as it is a p-series with p = 2. The limit
1 1 1 1
· arcsin arcsin
2n + 5 n 2n + 5 n 1
. lim = lim · =
1 1 1 2
n2 n n
∞
1 1
belongs to R+ . By Corollary 2 of the General Comparison Test, the series · arcsin is
n=1
2n + 5 n
convergent. As it is of positive terms, it is absolutely convergent.
2.7. Solved Exercises 147
∞
1 1 1 π
11. a) The series sin n consists of positive terms because 0 < n < , ∀n ∈ N. Consider the series
n=1
n 2 2 2
∞
1 1
. This series is convergent since it is a geometric series with r = . To show that the first series is
n=1
2n 2
also convergent, we can use Corollary 3 of the General Comparison Test. Specifically, we can take the limit
1 1 1
sin n sin n
n 2 1 2
. lim = lim · = 0 × 1 = 0.
1 n 1
2n 2n
Since the limit value equals zero, we can conclude that the series under consideration is convergent. More-
over, as it is of positive terms, it is absolutely convergent.
∞ √ ∞
3
n2 1
b) The series √ 2
is of positive terms. Consider the series 4/3
, which is convergent as it is
n=1
n n + 2n n=1
n
4
a p-series with p = . The limit
3
√3
n2
√
n n + 2n2 n2 1
. lim = lim √ =
1 n n + 2n2 2
n4/3
is finite √
and different from zero; therefore, by Corollary 2 of the General Comparison Test, the series
∞ 3
n2
√ is convergent. As it is a series of positive terms, it is absolutely convergent.
n=1
n n + 2n2
∞
cos(πn) 1
c) The series √ is alternating because cos(nπ) = (−1)n and √ > 0. We will study the
n=2 n2 − 1 n2 − 1
∞
1
series of modules, √ , by comparing it with the harmonic series. The limit
n 2−1
n=2
1 √
√
n2 − 1 n2
. lim = lim √ =1
1 n −1
2
is finite and different from zero; therefore, by Corollary 2 of the General Comparison Test, the series
+∞
1
√ is divergent.
n 2−1
n=2
Let us apply Leibniz’s test to study the alternating series:
1
(i) lim √ = 0.
n2 − 1
1
(ii) √ > 0, ∀n ∈ N \ {1}.
n2 − 1
√ 1
(iii) The sequence ( n2 − 1) is increasing so √ is decreasing for n ≥ 2.
n2 − 1
148 Numerical Series
The conditions for applying the test are satisfied; therefore, the alternating series is convergent. As the
series of modules is divergent, the given series is conditionally convergent.
∞
(−1)n 1
d) The series √ is alternating because an = √ > 0, ∀n ≥ 2. Let us start by studying
n=2
n log(n) n log(n)
the series of modules
∞ !
!
! (−1)n ! +∞ 1
!√ ! √
. ! n log(n) ! = n log(n)
.
n=2 n=2
∞
1 3
and the fact that the series 3/4
is divergent because it is a p-series with p = , by Corollary 4 of the
n=2
n 4
General Comparison Test, the series of modules is divergent. Consequently, if convergent, the original series
will be conditionally convergent. Since the series is alternating, let us check if we are in the conditions of
Leibniz’s test:
1
(i) √ > 0.
n log(n)
1
(ii) lim √ = 0.
n log(n)
√ 1
(iii) Since ( n log(n)) is clearly increasing, the sequence √ is decreasing.
n log(n) n∈N\{1}
As we are in the conditions of Leibniz’s test, we can conclude that the alternating series is convergent. From
the study on the series of modules, we can affirm that the series is conditionally convergent.
∞
4 + sin(n) 3 4 + sin(n)
e) The series √
3
is of positive terms. We can observe that 0 < √ 3
≤ √3
, ∀n ∈ N.
n=1
n + 1 n + 1 n+1
∞
1 1
As the series √
3
is a p-series with p = , it is divergent. Since the limit
n=1
n 3
3
√ √
3
n+1 33n
. lim = lim √
3
=3
1 n+1
√
3
n
is finite and different from zero, Corollary 2 of the General Comparison Test allows us to conclude that the
∞ ∞
3 4 + sin(n)
series √3
is divergent. Hence, the General Comparison Test ensures that the series √
3
n=1
n + 1 n=1
n+1
is divergent.
∞
n2n sin(n3 )
f) We begin by analyzing the series of modules, as the series is not of positive terms. The
n=1
(3n2 + 5)n
inequality ! 2n !
! n sin(n3 ) ! n2n
.0 ≤ !! 2 n
!≤
!
(3n + 5) (3n + 5)n
2
2.7. Solved Exercises 149
holds true in N.
We evaluate the limit
n n2n n2 1
. lim = lim = ,
(3n2 + 5)n 3n2 + 5 3
∞
n2n
which is less than 1. Consequently, by Cauchy’s Root Test, the series 2 + 5)n
is convergent. By
n=1
(3n
∞ ! !
! n2n sin(n3 ) !
the General Comparison Test, the same happens to the series ! !
! (3n2 + 5)n !. The original series is
n=1
absolutely convergent since the series of modules is convergent, considering Definition 2.4.1.
n−1 n−1
g) Let an = arcsin 2
. Since 0 < 2 < 1, ∀n > 2, we have that an > 0, and therefore, the series
n +1 n +1
∞
+∞
(−1)n an is alternating. Let us start by studying the series of modules an . The limit
n=3 n=3
n−1
arcsin
n2 + 1
. lim =1
n−1
n2 + 1
∞
n−1
belongs to R+ , and the series 2+1
is divergent (compare it with the harmonic series). We can then
n=3
n
conclude that the series of modules is divergent. Consequently, if convergent, the series under study is
conditionally convergent. Since the series is alternating, let us check if we are in the conditions of Leibniz’s
test:
n−1
(i) arcsin > 0, ∀n > 2.
n2 + 1
n−1
(ii) lim arcsin 2
= 0.
n +1
x−1
(iii) Let f be the real function of a real variable defined by f (x) = arcsin . This function is
x2 + 1
decreasing on [3, +∞[, because
−x2 + 2x + 1
.f (x) = √ < 0, ∀x ∈ [3, +∞[.
(x2 + 1) x4 + x2 + 2x
is decreasing.
As we are in the conditions of Leibniz’s test, we can conclude that the alternating series is convergent. The
study done on the series of modules shows that the series is conditionally convergent.
∞
4 × 7 × · · · × (3n + 1)
h) The series is of positive terms. Let us use D’Alembert’s test to study it. If an
n=1
8 × 11 × · · · × (3n + 5)
is the general term of the series, then:
150 Numerical Series
4 × 7 × · · · × (3n + 1)(3n + 4)
an+1 8 × 11 × · · · × (3n + 5)(3n + 8) 3n + 4
. lim = lim = lim = 1.
an 4 × 7 × · · · × (3n + 1) 3n + 8
8 × 11 × · · · × (3n + 5)
Since this limit is 1, we cannot draw any conclusions using D’Alembert’s test. Therefore, let us use Raabe’s
test, which is a common procedure when D’Alembert’s test is inconclusive. In this case,
an 3n + 8 4n 4
. lim n − 1 = lim n − 1 = lim = .
an+1 3n + 4 3n + 4 3
As this value is greater than 1, we can conclude that the series is convergent. Moreover, since it is a series
of positive terms, it converges absolutely.
∞ ∞
arctan(n3 ) 2n (2n)!
i) Consider the two series √ 2
and n (2n + 1)!
. The first one is of positive terms, and we
n=1
n + n n=0
3
∞
1
compare it with the p-series 2
, which we know to be convergent. Indeed, the limit
n=1
n
arctan(n3 )
√
n + n2 n2 π
. lim = lim √ arctan(n3 ) =
1 n + n2 2
n2
∞
arctan(n3 )
is finite and different from zero, so, by Corollary 2 of the General Comparison Test, the series √
n=1
n + n2
is convergent.
The second series is also of positive terms. We can study it using D’Alembert’s test. Let an be the general
term of the series:
This limit is less than 1; therefore, by D’Alembert’s test, the second series is convergent.
The original series is the sum of the two series studied, so it is convergent. Since it is of positive terms, the
series is absolutely convergent.
∞ sin
√1
n 1 π
j) The series √ consists of positive terms, since 0 < √ < , ∀n ∈ N. To check its convergence,
n=1
n+ n n 2
∞
1
let us compare it with the series , which we know to be convergent because it is a p-series with
n=1
n3/2
2.7. Solved Exercises 151
3
p= . Taking the limit, we have
2
sin √1 sin √1
n n
√ √ sin √1n
n+ n n+ n n
. lim = lim = lim · √ = 1.
1 1 1 n+ n
√ √
n3/2 n n n
Since the limit is finite and different from zero, by Corollary 2 of the General Comparison Test, the series
∞ sin √1
n
√ is convergent. Moreover, as a series of positive terms, it converges absolutely.
n=1
n + n
∞
1 1
k) The series n sin − (n + 2) sin is telescopic. We can write the sequence of partial
n=1
n n+2
sums:
1
S1 = sin(1) − 3 sin
3
1 1 1
S2 = sin(1) − 3 sin +2 sin −4 sin
3 2 4
1 1 1 1 1
S3 = sin(1) − 3 sin +2 sin −4 sin + 3 sin − 5 sin
3 2 4 3 5
1 1 1
= sin(1) + 2 sin −4 sin −5 sin
2 4 5
.
1 1 1 1 1
S4 = sin(1) + 2 sin −4 sin −5 sin + 4 sin − 6 sin
2 4 5 4 6
1 1 1
= sin(1) + 2 sin −5 sin −6 sin
2 5 6
.
..
1 1 1
Sn = sin(1) + 2 sin − (n + 1) sin − (n + 2) sin
2 n+1 n+2
.
..
As
1
sin
1 n
. lim n sin = lim =1
n 1
n
we have
1
. lim Sn = sin(1) + 2 sin − 2.
2
The series is convergent and
∞
1 1 1
. n sin − (n + 2) sin = sin(1) + 2 sin − 2.
n=1
n n + 2 2
152 Numerical Series
∞
2n + 3
l) The series is of positive terms. We can establish the following inequality:
n=1
(n + 1)!
2n + 3 2n
. ≤2 , ∀n ≥ 2.
(n + 1)! (n + 1)!
∞
2n
To determine the convergence of the series , we can apply D’Alembert’s test. Let an be the
n=1
(n + 1)!
general term of the series:
2n+1
an+1 (n + 2)! 2
. lim = lim = lim = 0.
an 2n n+2
(n + 1)!
This limit is less than 1, indicating that the series is convergent according to D’Alembert’s test. Using the
General Comparison Test, we can affirm that the series under study is convergent and, hence, absolutely
convergent.
∞
arctan(n + 1) − arctan(n)
m) The series is of positive terms because arctan(x) is an increasing function
n=1
n2
∞
1
on R. Let us compare this series with the series , which we know is convergent as it is a p-series
n=1
n2
with p = 2. By Corollary 3 of the General Comparison Test, given the value of the limit
arctan(n + 1) − arctan(n)
n2
. lim = lim (arctan(n + 1) − arctan(n)) = 0,
1
n2
we can conclude, as we are comparing with a convergent series, that the series under study is convergent.
As it is a series of positive terms, it converges absolutely.
∞
log(n)
n) The general term of the series 1 is not a null sequence. In fact,
n=1
n sin n
2.7. Solved Exercises 153
1
log(n) n
. lim = lim log(n) · = +∞.
1 1
n sin sin
n n
Therefore, by Theorem 2.2.1, the series diverges.
∞
n n+1 n+1
o) Consider the series log(2) log . It consists of positive terms because > 1, ∀n ∈ N.
n=1
n n
∞
n
Let us analyze the series log(2) . It converges because it is a geometric series with 0 < r = log(2) < 1.
n=1
The limit
n n+1
log(2) log
n n+1
. lim n = lim log =0
log(2) n
allows us to use Corollary 3 of the General Comparison Test to conclude that the series under study is
convergent because we are comparing it with a convergent series. As the series is of positive terms, it is
absolutely convergent.
p
p) Given that n ≥ 3 implies log(n) > 1, then log log(n) > 0 so n log(n) log log(n) > 0, ∀p ∈ R.
Therefore, the series is of positive terms.
Let us assume that p ≤ 0. If n > ee (for example, if n ≥ 33 = 27), then log log(n) > 1. Thus,
−p
log log(n) ≥ 1, with equality holding for p = 0. Consequently, we can write:
−p
1 log log(n) 1
. p = ≥ .
n log(n) log log(n) n log(n) n log(n)
+∞
1
Since the series is divergent as seen in Example 2.5.2, we can apply the General Com-
n=27
n log(n)
+∞
1
parison Test to deduce that the series p is divergent. Therefore, the series
n=27
n log(n) log log(n)
+∞
1
p is also divergent.
n=3
n log(n) log log(n)
Now, suppose that p > 0. With the application of the Integral Test in mind, let us consider the function
1
.f (x) = p
x log(x) log log(x)
p
and let g(x) = x log(x) log log(x) . We know that g(x) > 0, ∀x ∈ [3, +∞[. Additionally, since the
logarithm function is continuous and increasing, g is continuous and increasing on [3, +∞[, so f is positive,
+∞
continuous, and decreasing on [3, +∞[. We can now study the improper integral f (x) dx.
3
If p = 1,
# $y
y 1 1 −p+1
p dx = log log(x)
3 x log(x) log log(x) −p + 1 3
.
1 −p+1 −p+1
= log log(y) − log log(3) .
−p + 1
154 Numerical Series
If p = 1,
y 1 y
. dx = log log log(x) = log log log(y) − log log log(3) .
3 x log(x) log log(x) 3
∞
(n + 1)n
12. a) The series is of positive terms. To analyze this series, we will use D’Alembert’s test. Let an
n=1
3n n!
be the general term of the series:
(n + 2)n+1
an+1 3n+1 (n + 1)! 1 n+2 n+2 n e
. lim = lim n = lim = .
an (n + 1) 3 n+1 n+1 3
3n n!
Since this limit is less than 1, according to D’Alembert’s test, the series converges.
∞
(n + 1)n
b) In the previous item, we proved that the series is convergent, which implies, by Theorem 2.2.1,
n=1
3n n!
(n + 1)n
that its general term is a null sequence. Therefore, lim = 0.
3n n!
∞
13. We will study the convergence of the series of modules, |an |. We will use the Root Test or one of its corollaries,
n=1
to do this. We have
⎧
⎪ n2 n
⎪
⎪ n n
⎪
⎪
n
= , if n = 3k, k ∈ N
⎪
⎪ n+1 n+1
⎪
⎪
⎨ 1
⎪
1
.
n
|an | = n
= √ , if n = 3k + 1, k ∈ N0
⎪ n!
⎪
n
n!
⎪
⎪
⎪
⎪ !! !
⎪
⎪ (−1)n !! 1
⎪
⎪ n !
if n = 3k + 2, k ∈ N0 .
⎩ ! (n + 1)n ! = n + 1 ,
1 1
The sublimits of n
|an | are
and 0; therefore, lim n |an | = . Since this value is less than 1, we can affirm, by
e e
∞
∞
Corollary 1 of the Root Test, that the series |an | is convergent. Consequently, the series an is absolutely
n=1 n=1
convergent.
2.7. Solved Exercises 155
∞
14. One initial approach is to analyze the general term of the series an . Since lim a2n = lim a2n+1 = +∞, then
n=1
lim an = +∞, and by Theorem 2.2.1, the series diverges.
Another approach: ⎧
⎪
⎪
n+2
an+1 ⎨ , if n is even
n
. =
an ⎪
⎪ n+1
⎩ , if n is odd, n > 1.
n−1
an+1
We find that > 1, ∀n ∈ N \ {1}, and therefore by the Ratio Test, the series diverges.
an
∞
15. If (an ) is a sequence of positive terms such that the series nan is convergent, then the sequence (an ) is
n=1
bounded. We can prove this by reductio ad absurdum: If (an ) was
not bounded, then lim (nan ) would not be
∞
zero, and in this case, the series nan would not be convergent. It follows from the fact that (an ) is bounded
n=1
that
a2n an
. lim = lim = 0.
nan n
∞
By Corollary 3 of the General Comparison Test, the series a2n is convergent.
n=1
∞
16. As the series an converges, lim an = 0. By hypothesis, an > 0, ∀n ∈ N, so log(1 + an ) > 0. Thus, we have
n=1
two series of positive terms. The limit
log(1 + an )
. lim =1
an
∞
is finite and different from zero; therefore, by Corollary 2 of the General Comparison Test, the series log(1+an )
n=1
also converges.
log(1 + x) 0
Note: The limit lim is an indeterminate form of type . Applying L’Hôpital’s Rule, we get
x→0 x 0
log(1 + x) 1
. lim = lim =1
x→0 (x) x→0 1+x
log(1 + x) log(1 + an )
therefore, lim = 1, which implies that lim = 1.
x→0 x an
∞
17. a) The series (−1)n−1 an is alternating. Leibniz’s test cannot be applied because the sequence (an ) does
n=1
not decrease. In fact, if n is even,
1 1 n2 − n − 1
.an+1 − an = − 2 = 2 > 0.
n+1 n n (n + 1)
156 Numerical Series
∞
b) Let us prove that the series (−1)n−1 an is divergent. If it was convergent, then, by Theorem 2.2.4, any
n=1
∞
series obtained from it by association of its terms would be convergent. Consider the series bn , where
n=1
(bn ) is the sequence defined by
1 1 (2n)2 − 2n + 1 4n2 − 2n + 1
.bn = a2n−1 − a2n = − = = .
2n − 1 (2n)2 (2n)2 (2n − 1) 4n2 (2n − 1)
The limit
4n2 − 2n + 1
4n2 (2n − 1) 4n2 − 2n + 1 1
. lim = lim =
1 4n(2n − 1) 2
n
is finite and different from zero; therefore, based on Corollary 2 of the General Comparison Test, the series
∞ ∞
bn diverges because the harmonic series is divergent. However, we obtained the series bn by
n=1 n=1
association of the terms of the original series. Hence, the original series must be divergent.
∞
18. a) Suppose that the series bn is convergent. Then, its general term is a null sequence, which implies
n=1
that all its subsequences have limit zero. Therefore, we have lim b2n−1 = 0, or equivalently, lim an = 0.
∞
Conversely, let lim an = 0 and (Sn ) be the sequence of partial sums of the series bn . We can see that
n=1
S2n = 0 and S2n−1 = an for all n ∈ N. Since lim an = 0, it follows that lim Sn = 0. By definition, if the
sequence of partial sums of a series has a finite limit, the series converges. As a result, it can be inferred
∞
that the series bn converges.
n=1
∞
∞
∗ denote the general terms of the sequences of partial sums of the series
b) Let Sn and Sn |bn | and |an |,
n=1 n=1
respectively. We can observe that
∞
Reciprocally, if the series ∗ exists and is finite. This implies that
an converges absolutely, then lim Sn
n=1
∞
∗ + S∗
lim S2n+1 = lim(Sn ∗
n+1 ) = lim 2Sn = lim S2n , and we can state that the series bn converges
n=1
absolutely.
∞
19. If the series an is convergent, then the sequence of its partial sums, (Sn ), is also convergent. This implies
n=1
that all its subsequences are convergent and have the same limit. Since the sequence (an ) is decreasing, we have
and as lim (S2n − Sn ) = 0 and an ≥ 0, we can conclude that lim n a2n = 0, which implies that lim 2 n a2n = 0.
Similarly, we have
As lim (S2n+1 − Sn+1 ) = 0 and (an ) is a sequence of nonnegative terms, we conclude that lim n a2n+1 = 0.
This also implies that lim (2n + 1) a2n+1 = 0 because lim an = 0 since, by hypothesis, the series is convergent.
We have shown that the subsequence of even order terms of (n an ) has the same limit as its subsequence of odd
order terms, which is zero. Hence lim n an = 0.
a2 + b 2
20. We can use the inequality ab ≤ , ∀ a, b ∈ R, along with the fact that the sequence (an ) is of positive
2
terms, to establish the following expression:
1
1 √ n2p
+ an 1 1
.0 ≤ an ≤ = + an .
np 2 2 n2p
∞
∞
1 1
By hypothesis, the series an is convergent, and if p >
, then 2p > 1, which implies that the p-series 2p
n=1
2 n=1
n
is convergent. The sum of these two series is convergent, and given the previous inequality, we can state by the
∞ √
an
General Comparison Test the convergence of the series .
n=1
np
∞
1 1
21. Consider the series , which is divergent as it is the harmonic series. Let bn = . Then
n=2
n−1 n−1
1
bn+1 n n−1 1 an+1
. = = =1− ≤ , ∀n ≥ 2.
bn 1 n n an
n−1
∞
By Corollary 5 of the General Comparison Test, we can determine that the series an is divergent.
n=1
158 Numerical Series
g
22. The functions g and g are, by hypothesis, continuous and positive on [p, +∞[, so is continuous and positive.
g
2
g (x) g (x)g(x) − g (x) g
In addition, = 2 < 0, so is decreasing on [p, +∞[. We are in the conditions of
g(x) g(x) g
the Integral’s Test. But
y g (x)
. dx = log g(y) − log g(p) ;
p g(x)
g (x)
+∞
therefore, the improper integral dx is convergent if and only if there exists and is finite lim log g(x) .
p g(x)
x→+∞
By the continuity of the logarithm function, lim log g(x) exists finite if and only if there exists finite and dif-
x→+∞
ferent from zero the lim g(x). But, by hypothesis, g is positive and increasing (g is positive), so it cannot
x→+∞
have a limit 0 when x tends to +∞.
+∞
g (n)
Conclusion: The series is convergent if and only if lim g(x) exists and is finite.
n=p
g(n) x→+∞
23. We can prove by induction that (vn ) is a sequence of positive terms. Applying D’Alembert’s test, we find that
∞
the series vn is convergent because
n=2
vn+1 π
. lim = lim sin = 0 < 1.
vn n
Furthermore,
arctan(vn )
. lim = 1,
vn
∞
which implies by Corollary 2 of the General Comparison Test that the series arctan(vn ) is convergent.
n=2
24. Let us look at the numbers of the construction of the Sierpinski triangle in a table.
1 0 1 1 0
2 2
1 1 1
3 1 1×
2 2 2
2 2
1 1 1
32 3 3×
22 22 22
2 2
1 1 1
33 32 32 ×
23 23 23
. . . . .
. . . . .
. . . . .
2.7. Solved Exercises 159
So the total area of the squares removed is the sum of the geometric series
∞
2 ∞
1 1 3 n
.A = 3n = = 1.
n=0
2n+1 4 n=0 4
Since the initial square has area 1, we conclude that the area of the Sierpinski triangle is zero.
∞
25. The sum of the areas of the circles is given by the series πr12 + 2
3πrn , where r1 is the radius of the largest
n=2
circle, r2 is the radius of the circles tangent to it, etc.
π
We need to find the expression of rn as a function of r1 . In the figure, the angle ∠BAC measures radians and
3
π r1 1 r1 1 r2
cos = , and therefore, = . Similarly, = . Then |AB| = 2r1 and |DB| = 2r2 . As a
3 |AB| 2 |AB| 2 |DB|
1
result, we have the equality 2r1 = r1 + r2 + 2r2 , from which we conclude that r2 = r1 . Similarly, we can show
n−1 3
1 1 1
that r3 = r2 = r1 and, more generally, rn = r1 , n ≥ 2.
3 9 3
The series now takes the form
∞ ∞ 2 ∞
1 1 11 2
2 2
.πr1 + 3πrn = πr12 + 3π n−1
r 2
1 = πr 2
1 1+3 n−1
= πr1 .
n=2 n=2
3 n=2
9 8
1 π |BC| 1
What is the value of r1 ? Considering that |BC| = , |AB| = 2r1 , and sin = , we have r1 = √ .
2 3 |AB| 2 3
Finally, the sum of the areas of the circles is
∞
2 2 11π
.πr1 + 3πrn = .
n=2
96
∞
26. The area of the colored region can be written as A = (An − Bn ), where An is the area of the circle and Bn
n=1
is the area of the square inscribed in that circle. Let rn be the radius of the nth circle and ln be the length of the
side of the inscribed square.
160 Numerical Series
What is the relationship between r1 and rn ? We know that 2r2 = l1 , so 4r22 = l12 . As l12 = 2r12 , we have that
1 1
r22 = r12 . By induction, it is easy to show that rn
2
= n−1 r12 . Thus,
2 2
∞
∞
1
.A = (π − 2) rn
2
= (π − 2) r 2 = 2(π − 2) r12 .
n=1 n=1
2n−1 1
1 1 1 1 ∞
a) + + + + ··· 5. Show that if an = A ∈ R, then
3 8 15 24
n=0
1 1 1
b) + + + ···
1×2×3 2×3×4 3×4×5 ∞
1 1 1 . (an−1 + an + an+1 ) = 3A − a1 − 2a0 .
c) + + +· · · n=1
1×2×3×4 2×3×4×5 3×4×5×6
6. Test for convergence or divergence the following se-
2. Find the sum of the series: ries. If convergence occurs, indicate whether it is
∞
conditional or absolute:
2n + 1
a) (−1)n ∞
n(n + 1) (−1)n+1
n=1
a)
∞ n
1 a n=1
b) tan n , a ∈ R \ {kπ, k ∈ Z}. ∞
2n 2 (−1)n−1
n=1 b)
x x n2
Hint: tan = cot − 2 cot(x) n=1
2 2 ∞
∞ √ √ n2
n+1− n c) (−1)n
c) √ 1 + n2
2
n +n n=1
n=1
∞ ∞
(−1)n − 8 (−1)n
d) 7. Consider the series :
3n n=1
(n + 1)2
n=1
∞
n+1
1 n 1 a) Verify that it is convergent.
e) 1− − 1−
n=1
n n+1 b) Calculate its sum with an error less than
1
∞
.
f) 2−n − 2−3n 10000
n=1
∞ 8. Let an and bn be two convergent series, cn
arctan(n) + arctan(−n + 2)
g) and dn be two divergent series, and α = 0 be a
n=3
arctan(n) arctan(n − 2) real number. What can be said about the conver-
∞ gence of the following series?
3. Let an be a convergent series. Show that the a) (an + bn ) e) (an cn )
n=1
∞
a3 + 5n
series √n is divergent. b) (an bn ) f) (αcn )
n=1 n2 + 1
4. Find the values of x for which the following series c) (αan ) g) (cn + dn )
converge and, when possible, evaluate the sum of
the series: d) (an + cn ) h) (cn dn )
∞
8n 9. Test the convergence of the following series using a
a)
(x + 1)3n comparison test:
n=0
∞
∞
n 1
b) |x| − 1 a)
n=1 n=1
n3 + 3
162 Numerical Series
∞ ∞
n+1 n!
b) b)
n=1
n2 + 1 n=1
3n
∞
∞
1 n3
c) c)
n=1 n(n2 + 1) n=1
n!
∞ √ ∞
n n nn
d) √ d)
3 n3 + 1 (2n)!
n=1 (n + 1) n=1
∞ ∞
n!
e) n− n2 − 1 e) 2
n=1
n
n=1
∞ ∞
π 2 n n+1
f) sin f) 2−n+(−1) 3−n+(−1)
n=1
n n=1
∞ √ ∞
n!
n+ n g)
g)
n2 − n n=1
3 × 5 × 7 × · · · × (2n − 1)
n=2
∞ ∞
n2 + 1 2 × 4 × 6 × · · · × (2n)
h) h)
n3 + 1 n=1
2 × 5 × 8 × · · · × (3n − 1)
n=1
∞ ∞
n n
a) n
a)
n=1
3 n=1
(2n + 1)!
2.8. Proposed Exercises 163
∞
∞
1
b) n) ne log(n)
n=1
n(n + 1)(n + 2) n=2
∞ √ ∞
n−1 2n
c) o)
n n=0
(2n + 1)!
n=1
∞
∞
2 cos(nθ)
(2n + 1)! p) ,θ∈R
d) n5/2
n=1
n × n! n=1
∞
3n n2n
q)
14. Investigate the convergence or divergence of the n=0
(2n)!
following series. If convergence occurs, indicate ∞
log(n!) + n!
whether it is conditional or absolute: r)
n=1
nn + 2n
∞
∞ √
1 3
n2 + 1
a) √ s) (−1)n
n=0 n2 + 1 n+3
n=0
∞
1
b)
n=2 n(n2 − 1) 15. Examine the convergence or divergence of the follow-
∞ ing series. If convergence occurs, indicate whether
en n n
c) it is conditional or absolute:
n=1
n!
∞
∞
sin(n + 1)
2 a) 2 log(n + 1)
d) ,p∈R n=1
n
n2 + p 2
n=1 ∞
1 (n!)2
∞
b) +
cos(n π2 ) 3 n (2n)!
e) n=3
n2 ∞
n=1 1
∞ c)
52n (n + 1)(n + 2)
f) n=1
(n + 1)! ∞
n=0 e
d) cos(nπ) tan
∞
n
π n n=3
g) tan ∞
n=3
n 1 + (−1)n n
e)
∞ n2 + 5
1 nπ n=1
h) cos ∞ 3n−1
n=1
n 2 2n
f)
∞ 4n + 1
1 n=1
i) ∞
1)
(1+ n n! n 1
n=1 n g) 2 +
nn n2 + n
∞ n=1
n2 + 2n + 1 ∞
j) (n + p)!
n=1
3n2 + 2 h) , p, q ∈ N
n!(n + q)!
∞ n n=1
sin 3π ∞
k) 2 n!
n2 + 1 i)
n=1
n=1
(π + 1)(π + 2) . . . (π + n)
∞
1 ∞
l) (−1)n 1
nn log(n) j) +√
n=2 n=1
n(n + 1) 3n
∞
∞
1 2n + 3
m)
1
k) (−1)n
1+ log(n) (n + 1)(n + 2)
n=2 n n=0
164 Numerical Series
∞
n
a 19. Knowing that an is convergent and of positive
l) , a ∈ R \ {−3}
n=0
a+3 terms and bn > 0, ∀n ∈ N, what can be said about
an
∞ the convergence of the series ?
1 1 + bn
m) n+a
, a ∈ R+
0
n=0
2
∞ ! ! 20. Knowing that an and bn are convergent se-
! 1 !!
n) log !!log ! ries of positive terms, study the convergence of the
n=2
n!
following series:
∞
(3n)!
o) 1 1
n=0
27n (n!)2 a) +
an bn
1 1 1×3 1 1×3×5 1 n+1
p) 1 + · + · + · + ···
2 3 2×4 5 2×4×6 7 b) an
∞ √ n
2n − 1 log(4n + 1)
q)
n=1
n(n + 1)
∞
21. Let an be a divergent series, an ≥ 0, and sn be
3 × 5 × 7 × · · · × (2n + 1)
r) the sum of its first n terms. Show that the series
2n (2n − 1) n! √
n=1 √
∞
. sn+1 − sn
(−3)n
s)
n=1
3 × 5 × 7 × · · · × (2n + 1) is divergent.
∞
2 + sin3 (n + 1) 0a np +···+a
p
t) 22. Prove that the series b0 nq +···+bq
in which a0 ,
n=1
2n + n2
. . . , ap , b0 , . . . , bq are real numbers and a0 > 0,
∞
1 b0 > 0, is convergent if and only if q − p > 1.
u)
n=1
2n − 1 + sin2 (n3 )
23. Study the conditional and absolute convergence of
the series:
16. Indicate for which values of α the following series ∞
an
are conditionally or absolutely convergent: a) ,a>0
n=1
(1 + a)(1 + a2 ) . . . (1 + an )
∞
n ∞
a) 1 + sin(α) (α + 1)(α + 2) . . . (α + n)
b) , β ∈ R \ Z− :
n=1 (β + 1)(β + 2) . . . (β + n)
∞ n=1
1
b) (−1)n i. If α ∈ R \ Z−
(n + 1)α
n=0 ii. If α ∈ Z−
17. a) Test for convergence the series
∞ n2 un+1 2 1
2n n 24. Let un > 0 and ≤ 1 − + 2 ∀n ∈ N \ {1}.
e . un n n
n=1
n+3 Show that un is convergent.
b) Based on the previous item, indicate the limit
n2
n 25. Consider the series
of the sequence e2n .
n+3 ∞ ∞
(−1)n 1
. and √ :
18. Let an and bn be two convergent series of pos- n=0
n! n=0
(n + 1) n+1
√
itive terms. Show that the series an bn also
converges.
a) Calculate the partial sum of order three of
an + b n
Hint: Prove that ≥ an b n . Cauchy’s product of the two series.
2
b) Study the convergence of the product series.
2.8. Proposed Exercises 165
26. The figure shows a sequence of cubes constructed connect the midpoints of the sides of this square to
as follows: The edge of the larger cube has length 1, form another square inside it. Repeat this process
and the length of the edge of each of the following with the new square, connecting its midpoints to
cubes is half the length of the edge of the previous form another square. Keep repeating this process
cube. Find the sum of the volumes of the cubes. infinitely, and we will obtain infinite squares within
the first one.
Determine the sum of the areas of all the squares.
which is called the limit of .(fn ) on D. We also say that .(fn ) converges
pointwise to f on D.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 167
A. Alves de Sá, B. Louro, Sequences and Series,
https://doi.org/10.1007/978-3-031-67202-6 3
168 Series of Functions
x n
Example 3.1.1 The sequence of functions . 1 + , defined on .R, con-
n
verges whatever .x ∈ R (see Fig. 3.1). The limit function is .f (x) = ex :
x n
. lim 1 + = ex ∀x ∈ R.
n
Figure 3.1: The function .f (x) = ex and the first terms of the sequence
continuous: ⎧
⎨0, if 0 ≤ x < 1
n
.f (x) = lim x =
⎩1, if x = 1
that is,
. lim sup |fn (x) − f (x)| = 0.
n→+∞ x∈D
The definition of uniform convergence means that, for every fixed .δ > 0,
there is an order from which the images of all functions are in the .δ neigh-
borhood of .f (x) for any .x ∈ D, that is, there is an order from which the
images of all functions are in the region of the plane defined by .f (x) − δ
and .f (x) + δ, in the set D. This is illustrated in Fig. 3.3.
170 Series of Functions
Figure 3.4: The sequence of functions . x nα e−nx , .α > 1, .α = 1, and .α < 1
Figure 3.5: The sequence of functions . x n1/2 e−nx
172 Series of Functions
∞
the series is also represented by . fn or by . fn .
n=1
∞
x2
Example 3.2.1 Consider the series . .
n=0
(1 + x2 )n
If .x = 0, all terms are zero; therefore, the series is convergent.
If .x = 0, we can write
∞ ∞ ∞ n
x2 1 1
.
2 )n
= x2 2 )n
= x2 ,
n=0
(1 + x n=0
(1 + x n=0
1 + x2
1
and this is a geometric series with ratio .r = ; as .|r| < 1, the series
1 + x2
is convergent and
∞
x2 1
. = x2 · = 1 + x2 .
(1 + x 2 )n 1
n=0 1−
1 + x2
The sum function is
⎧
⎨1 + x2 , if x = 0
f (x) =
.
⎩0, if x = 0.
3.2. Pointwise and Uniform Convergence of Series of Functions 173
∞
xn
Example 3.2.2 Consider the series . . We can use a numerical series
n=0
n!
test to study the pointwise convergence of series of functions. In this case,
we will apply D’Alembert’s test to study the series
∞ n
x
.
.
n!
n=0
n+1
x
(n + 1)! |x|
. lim n = lim = 0, ∀x ∈ R.
x n +1
n!
∞
xn
. = ex , ∀x ∈ R.
n=0
n!
∞
Example 3.2.3 Consider the series . x (1 − x)n , x ∈ [0, 1].
n=0
If .x = 0, all terms are zero. Therefore, the series is convergent.
∞
If .x = 0, as the series . (1 − x)n is geometric with ratio .r = 1 − x and
n=0
|r| < 1 since .x ∈ ]0, 1], the series converges. In this case,
.
∞
1
. x (1 − x)n = x · = 1.
n=0
1 − (1 − x)
∞
Then, the series . x (1 − x)n , .x ∈ [0, 1], converges pointwise to the
n=0
function
⎧
⎨1, if 0 < x ≤ 1
f (x) =
.
⎩0, if x = 0.
174 Series of Functions
that is,
. lim sup |f (x) − Sn (x)| = 0.
n→+∞ x∈D
∞
x2
Example 3.2.4 We saw that the series . is pointwise conver-
n=0
(1 + x2 )n
gent to the function f defined by
⎧
⎨1 + x2 , if x = 0
.f (x) =
⎩0, if x = 0.
The function f and some partial sums are illustrated in Fig. 3.6.
3.2. Pointwise and Uniform Convergence of Series of Functions 175
f (x)
|fn (x)| ≤ an , ∀ x ∈ D, ∀ n ∈ N,
.
fn+1 (x) + · · · + fm (x) ≤ fn+1 (x) + · · · + fm (x)
.
≤ an+1 + · · · + am ∀x ∈ D
Then
∃ p ∈ N : m > n > p ⇒ |fn+1 (x) + · · · + fm (x)| < δ,
.
or still,
m
.∃ p ∈ N : m > n > p ⇒ fr (x) < δ.
r=n+1
∞
Example 3.2.5 Let k be a constant such that .|k| < 1. The series . fn (x),
n=1
where
1 × 3 × · · · × (2n − 1) 2n 2n
f (x) =
. n ·k sin(x) ,
2 × 4 × · · · × 2n
is uniformly convergent on .R. In fact,
fn (x) = 1 × 3 × · · · × (2n − 1) · k 2n sin(x) 2n
2 × 4 × · · · × 2n
.
1 × 3 × · · · × (2n − 1) 2n
≤ · k , ∀x ∈ R,
2 × 4 × · · · × 2n
and the numerical series
∞
1 × 3 × · · · × (2n − 1) 2n
. ·k
n=1
2 × 4 × · · · × 2n
∞
sin(nx) 1 1
Example 3.2.6 Since . ≤ 2 , .∀x ∈ R, and the series . is
n2 n n=1
n2
∞
sin(nx)
convergent, the series . is uniformly convergent on .R.
n=1
n2
∞
x2 + n
Example 3.2.7 For each .x ∈ R, the series . (−1)n is alternating,
n=1
n2
and using Leibniz’s test, we can prove that it is convergent. However, it is
not absolutely convergent because
2
n x + n
x2 + n 1
. (−1) = ≥ , ∀x ∈ R,
n2 n2 n
∞
1
and the series . is divergent. This makes it impossible to use the Weier-
n=1
n
strass’ test to determine the uniform convergence of the series. Instead, we
have to study it directly.
For each .x ∈ R, let .Sn (x) be the partial sum of order n of the series
∞
x2 + n
.S(x) = (−1)n . By Corollary 1 of Theorem 2.3.2,
n=1
n2
x2 + n + 1
|S(x) − Sn (x)| ≤
. .
(n + 1)2
If .X ⊂ R is a bounded set, that is, if there exists .M > 0 such that .|x| ≤ M ,
∀x ∈ X, then
.
M2 + n + 1
|S(x) − Sn (x)| ≤
. , ∀x ∈ X;
(n + 1)2
The importance of uniform convergence lies in the fact that a series that
converges uniformly behaves similarly to the sum of a finite number of func-
tions. For instance, when we add a finite number of continuous functions,
the result is also a continuous function. Similarly, with a series of functions
that converge uniformly, we obtain the following outcome:
We can write
|f (x) − f (x0 )| ≤ |f (x) − Sn (x)| + |Sn (x) − Sn (x0 )| + |Sn (x0 ) − f (x0 )|.
.
1
√
3
n +1 n3
. lim = lim =1
1 3
n +1
n3/2
∞
1
is finite and different from zero, we conclude that the series . √
+1 n3
n=1
∞
cos(nx)
is convergent. By inequality (3.1), the series . √
n3 + 1 converges,
n=1
∞
cos(nx)
∀x ∈ R, so the series
. √
. is convergent, .∀x ∈ R, that is, the
n=1
n3 + 1
domain of f is .R. To conclude that f is continuous on .R, it is sufficient
that the following two conditions are satisfied:
cos(nx)
(i) The functions .fn (x) = √ are continuous on .R.
n3 + 1
∞
cos(nx)
(ii) The series . √ is uniformly convergent to f on .R.
n=1
n3 + 1
The functions .cos(nx) are continuous on .R, .∀n ∈ N, so the first condition is
true. The second condition is a consequence of the inequality (3.1) and the
Weierstrass’ test. As the series converges to f , this function is continuous.
∞
x2
Example 3.2.9 Let us consider the series . , on the interval
n=0
(1 + x2 )n
.[−a, a], .a > 0. We proved in Example 3.2.1 that this series converges
180 Series of Functions
x2
Since f is discontinuous at .x = 0 and .fn (x) = is continuous,
(1 + x2 )n
∀n ∈ N, the series is not uniformly convergent.
.
b
∗
.∀δ > 0 ∃ p ∈ N : n > p ⇒ Sn − f (x) dx < δ.
a
b n b
= fi (x) dx − f (x) dx
a a
i=1
3.2. Pointwise and Uniform Convergence of Series of Functions 181
b n b n
= fi (x) − f (x) dx ≤ fi (x) − f (x) dx
a a
i=1 i=1
b
= |Sn (x) − f (x)| dx.
a
∞
But the series . fn (x) converges uniformly to f on .[a, b]; therefore,
n=1
δ
∃ p ∈ N : n > p ⇒ |f (x) − Sn (x)| <
. , ∀x ∈ [a, b],
b−a
which implies that
b b
δ
∃p ∈ N : n > p ⇒
. |Sn (x) − f (x)| dx < dx = δ.
a a b−a
∞
e−nx
Example 3.2.10 Let us consider the series . , on .[0, 1].
n=1
2n
−nx
e 1 1
.
2n = enx 2n ≤ 2n , ∀x ∈ [0, 1].
∞
1 1
The series .
n
is geometric with ratio . ; therefore, it is convergent.
n=1
2 2
According to Weierstrass’ test, the original series is uniformly convergent on
.[0, 1]. By Theorem 3.2.4,
1 ∞ ∞
e−nx 1
e−nx
dx = dx
0 n=1
2n n=1 0 2n
.
∞ −nx 1 ∞
1 e 1 − e−n
= − = .
n=1
2n n 0 n=1 n2n
∞
xn
Example 3.2.11 The series is uniformly convergent on every in-
.
n=0
n!
terval .[a, b], .a, b ∈ R, because on this interval
182 Series of Functions
n
x Mn
.
n! ≤ n! , where M = max |a|, |b| ,
∞
Mn xn
and the series . is convergent. As .fn (x) = is continuous .∀n ∈ N,
n=0
n! n!
∞
xn
the series . is integrable term by term and
n=0
n!
b ∞ n ∞ b n
x x
dx = dx
a n=0 n! n=0 a
n!
.
∞ b ∞
1 xn+1 bn+1 − an+1
= = .
n=0
n! n + 1 a n=0 (n + 1)!
∞
Example 3.2.12 The series .x + (xn − xn−1 ) is convergent on .[0, 1] to
n=2
the function f defined by
⎧
⎨0, if 0 ≤ x < 1
.f (x) =
⎩1, if x = 1.
2 1 ∞
n 1 ∞
x xn+1 x 1 1 1
. = + − = + −
2 0 n=2
n+1 n 0 2 n=2 n+1 n
∞
1 1 1
= − − = 0.
2 n=2 n n+1
3.2. Pointwise and Uniform Convergence of Series of Functions 183
∞
Corollary 2 Let .a, b ∈ R, .a < b. If the series . fn converges pointwise
n=1
on the interval .[a, b] to a function f , if on this interval the derivatives .fn
∞
exist and are continuous, and if the series . fn converges uniformly on
n=1
[a, b], then f is differentiable on .[a, b] and
.
∞
f (x) =
. fn (x), ∀x ∈ [a, b].
n=1
(The series is differentiable term by term.)
∞
Proof: Let .g(x) = fn (x), .x ∈ [a, b]. By Corollary 1,
n=1
∞ ∞ x ∞
x x
. g(t) dt = fn (t) dt = fn (t) = fn (x) − fn (a) ,
a a a
n=1 n=1 n=1
that is,
x
. g(t) dt = f (x) − f (a);
a
therefore,
x
f (x) =
. g(t) dt + f (a),
a
∞
which implies that .f (x) = g(x), that is, .f (x) = fn (x).
n=1
184 Series of Functions
∞
Example 3.2.13 Consider the series . xn . It is a geometric series of
n=0
ratio x. The series converges if and only if .|x| < 1 and, in this case,
∞
1
. xn = .
n=0
1−x
∞
Let .0 < r < 1. Then .|xn | ≤ rn , .∀x ∈ [−r, r]. As . rn is a convergent
n=0
∞
n
numerical series, then . x is uniformly convergent on .[−r, r]. Further-
n=0
more
r r ∞ ∞ r
1
dx = xn dx = xn dx
−r 1−x −r n=0 n=0 −r
r ∞ n+1
r
x
⇔ − log |1 − x| =
−r
n=0
n+1 −r
.
∞
rn+1 − (−r)n+1
⇔ − log |1 − r| + log |1 + r| =
n=0
n+1
∞ ∞
1+r r n+1 r2n+1
⇔ log = 1 − (−1)n+1 = 2 .
1−r n=0
n+1 n=0
2n + 1
∞ ∞
Differentiating term by term the series . xn , we obtain . (n + 1) xn .
n=0 n=0
Since
(n + 2)xn+1 (n + 2)|x|
. lim
(n + 1)xn = lim n + 1 = |x|,
by D’Alembert’s test, the series converges if .|x| < 1 and diverges if .|x| > 1;
∞ ∞
if .|x| = 1, we have the divergent series . (n + 1) and . (−1)n (n + 1).
n=0 n=0
∞
n
The series . (n + 1) x is uniformly convergent on every interval .[−r, r] if
n=0
∞
0 < r < 1 because .(n + 1)xn ≤ (n + 1)rn and the series .
. (n + 1) rn is
n=0
3.2. Pointwise and Uniform Convergence of Series of Functions 185
Notes:
1
Theorem 3.3.1 Let = r. If r ∈ R+ , the power series
lim n
|an |
∞
an x is absolutely convergent at each point x ∈ ] − r, r[ and di-
n
n=0
vergent at each point x ∈ ] − ∞, −r[ ∪ ]r, +∞[. If r = +∞, then the
power series is absolutely convergent whatever x ∈ R. If r = 0, the series
converges if x = 0 and diverges if x = 0.
∞
Proof: Consider the series |an xn |. Let r ∈ R+ . Using the fact that
n=0
lim
.
n
|an xn | = |x| lim n
|an |,
3.3. Power Series 187
we can apply Corollary 1 of the Root Test to state that if |x| lim n
|an | < 1
∞
(i.e., if |x| < r), the series converges; therefore, the series an xn con-
n=0
verges absolutely.
If |x| lim n |an | > 1 (i.e., if |x| > r), then, by the reasoning used in
Corollary 1 of the Root Test, there exists a subsequence of |an xn | that
takes values greater than or equal to 1, which implies that the sequence
|an xn | does not tend to zero; thus the sequence an xn does not tend
∞
to zero, and the series an xn diverges.
n=0
If r = 0, then lim |an | = +∞. It follows that, whatever x ∈ R, x = 0,
n
∞
|x| lim n |an | = +∞ > 1, and as before, the series an xn diverges.
n=0
Obviously, if x = 0, all terms are zero; thus, the series is convergent.
If r = +∞, then lim n |an | = 0. Therefore, whatever x ∈ R,
∞
|x| lim n |an | = 0 < 1, and we can conclude that the series an x n
n=0
converges absolutely.
an
Corollary 1 If lim = r is a positive real number, zero, or +∞,
an+1
then the radius of convergence of the power series is r.
|an |; that
n
Note that there always exists lim is, every power series has a
an
radius of convergence. However, lim may not exist, as illustrated in
an+1
the following example.
188 Series of Functions
∞ ∞
n
Example 3.3.1 Consider the series an x n = 3 + (−1)n xn . As
n=0 n=0
⎧
an ⎨2n−1 , if n is even
=
⎩ 1 ,
.
an+1 if n is odd,
2n+2
an 1
1
lim does not exist, but r = . =
an+1 lim |an | 4
n
1 1
The interval of convergence of the series is − , . The series converges
4 4
1 1 1 1
absolutely on the interval − , and diverges on −∞, − ∪ , +∞ .
4 4 4 4
nn n
The radius of convergence of the series is infinite; therefore, the series is
absolutely convergent ∀x ∈ R.
∞
Example 3.3.3 The series n! xn has a radius of convergence r = 0:
n=0
an n! 1
.r = lim
an+1 = lim (n + 1)! = lim n + 1 = 0,
∞
xn
Example 3.3.4 Consider the series . Taking into account that
n=1
n
1
an n n+1
.r = lim
an+1 = lim 1
= lim
n
= 1,
n+1
3.3. Power Series 189
∞
(−1)n
Example 3.3.5 Consider the series n
(x + 1)n . Let y = x + 1.
n=0
2
The power series of y
∞
(−1)n n
. y
n=0
2n
has radius of convergence
(−1)n
an
.r = lim = lim 2n
= 2.
an+1 (−1) n+1
2n+1
∞
x2n+1
Example 3.3.6 Consider the series (−1)n . Let y = x2 . The
n=0
(2n + 1)!
∞
yn
series (−1)n has radius of convergence
n=0
(2n + 1)!
(−1)n
(2n + 1)! (2n + 3)!
.r = lim
(−1)n+1 = lim (2n + 1)! = lim (2n + 3)(2n + 2) = +∞;
(2n + 3)!
∞
(−1)n
– If x = −1, we obtain the series , which is conditionally con-
n=1
n
vergent.
∞
1
– If x = 1, we obtain the series , which is divergent.
n=1
n
∞
Theorem 3.3.2 If the radius of convergence of the series an xn is
n=0
r > 0 and if 0 < ρ < r, then the series is uniformly convergent on
[−ρ, ρ].
1 ∞
x2n
Example 3.3.7 Let us evaluate f (x) dx, where f (x) = (−1)n .
0 n=0
(2n)!
∞
yn
Let y = x2 . The series (−1)n has an infinite radius of convergence:
n=0
(2n)!
(−1)n
(2n)! (2n + 2)!
.r = lim
(−1)n+1 = lim (2n)! = lim (2n + 2)(2n + 1) = +∞,
(2n + 2)!
which implies that the original series converges for all x in R; therefore, it is
uniformly convergent on [0, 1] and integrable term by term on that interval:
1 ∞ ∞ 1
x2n n x2n
(−1) dx = (−1)n dx
0 n=0 (2n)! n=0 0
(2n)!
.
∞ 2n+1 1 ∞
1 x (−1)n
= (−1)n = .
n=0
(2n)! 2n + 1 0 n=0 (2n + 1)!
∞
Proof: Consider the series an xn . It is pointwise convergent on ] − r, r[;
n=0
the functions (an xn ) = nan xn−1 are continuous on ] − r, r[, ∀n ∈ N.
∞
Moreover, nan xn−1 is a power series with radius of convergence r:
n=0
1 1 1
. = √ = = r;
lim |nan |
n
lim n |an |
n n
lim |an |
n
∞
Note: If the power series n=0 an xn has radius of convergence r, then
both the series of derivatives and the series of indefinite integrals have the
same radius of convergence, r.
∞
(x − 5)n
Example 3.3.8 Let us consider the power series (−1)n+1 .
n=1
n 5n
∞
yn
Let y = x − 5. The series (−1)n+1 n has radius of convergence
n=1
n5
(−1)n+1
n+1
.r = lim
n 5n = lim (n + 1) 5 = 5,
(−1)n+2 n 5n
(n + 1) 5n+1
which implies the absolute convergence of the original series on the interval
]0, 10[.
∞
−1
– If x = 0, we obtain the series , which is divergent.
n=1
n
∞
(−1)n+1
– If x = 10, we obtain the series , which is conditionally
n=1
n
convergent.
Then, the original series converges absolutely on the interval ]0, 10[, con-
verges conditionally at x = 10, and diverges on ] − ∞, 0] ∪ ]10, +∞[.
∞
α(α − 1) . . . (α − n + 1) n
. =1+ x .
n=1
n!
If .α ∈ N, the series is reduced to the development of Newton’s binomial.
Suppose that .α ∈ N0 . Then let us study the convergence of the series. The
radius of convergence is
α(α − 1) . . . (α − n + 1)
Collin Maclaurin (1698–
1764) was a Scottish lim n! = lim (n + 1)! 1
α(α − 1) . . . (α − n + 1)(α − n) n! |α − n|
mathematician. In 1717,
at the age of 19 years, he
.
(n + 1)!
was appointed professor n+1
in Aberdeen. In 1719, = lim = 1.
|α − n|
he published the work
Organic Geometry , a text Therefore, the series is absolutely convergent on .] − 1, 1[ and divergent on
that can be considered
] − ∞, −1[ ∪ ]1, +∞[.
.
the most important of
his works and contains, This series is usually referred to as the binomial series.
among others, an original
method of generating
A fundamental question in the Taylor series development of an indefinitely
conics. He developed
Newton’s work in calculus, differentiable function is as follows:
geometry, and gravitation.
(Source of image: Pencil
Is there a neighborhood V of .x0 such that
and chalk on paper by
David Steuart Erskine
f (x) = f (x0 ) + f (x0 )(x − x0 )
.
In reality, the mere existence of the derivatives .f (n) (x0 ) for all natural val-
ues of n, although it allows writing the Taylor series of f at .x0 , does not
guarantee that, in some neighborhood of .x0 , the equality is satisfied:
∞
f (n) (x0 )
f (x) =
. (x − x0 )n , (3.2)
n=0
n!
0 + 0x + 0x2 + · · · ,
.
which converges to the zero function on .R. Therefore, the only point where
f equals the sum of the series is 0, since .f (x) = 0 if .x = 0.
. lim Rn (x) = 0, ∀x ∈ V.
Proof: We know that the expression of the Lagrange remainder, .Rn (x), is
(x − x0 )n
Rn (x) = f (n) x0 + θ (x − x0 )
. , 0 < θ < 1.
n!
Then n
k|x − x0 |
. |Rn (x)| ≤ M , x∈V;
n!
n
k|x − x0 |
since the series of general term . is convergent, this sequence
n!
has limit 0. The desired result is an immediate consequence of Theo-
rem 3.4.1.
196 Series of Functions
|x|n
But .lim = 0, .∀x ∈ R, as it is the general term of a convergent series,
n!
which implies that
∞
x2n+1
. sin(x) = (−1)n , ∀x ∈ R.
n=0
(2n + 1)!
In Fig. 3.7, we can see how the partial sums are increasingly better ap-
proximations of the function .sin(x), thus illustrating the convergence of
n
x2k+1
the series. The polynomial . (−1)k is denoted by .P2n+1 in the
(2k + 1)!
k=0
figure. The black line represents the graph of the function .sin(x).
3.4. Taylor Series and Maclaurin Series 197
The Maclaurin series developments in the first two examples were obtained
directly from the formula:
f (0) 2
f (0) + f (0) x +
. x + ··· ,
2!
by evaluating all derivatives of the function at zero. Since this process
is quite laborious, it is not commonly employed. Instead, well-known de-
velopments are often utilized in practice. Furthermore, the next theorem
guarantees that these developments are indeed the Taylor series.
ment of .x − x0 is unique.
198 Series of Functions
∞
Proof: By hypothesis, .f (x) = an (x − x0 )n in a neighborhood V of .x0 ,
n=0
which implies that .f (x0 ) = a0 . Differentiating,
∞
f (x) =
. n an (x − x0 )n−1 ,
n=1
1
Example 3.4.6 Consider the function .f (x) = . Taking into ac-
2 + 3x
count that
1 1 1
. = ·
2 + 3x 2 1 − − 23 x
and that
∞
1
. = xn , ∀x ∈ ] − 1, 1[,
1 − x n=0
we can deduce that
∞ n
1 1 1 3
. · 3 = − x ,
2 1 − −2 x 2 n=0 2
3 2
and this equality is valid as long as .− x < 1, that is, .|x| < . Then, the
2 3
Maclaurin series of f is
∞
3n 2
. (−1)n xn , |x| < .
n=0
2n+1 3
1 1
Example 3.4.7 Let .f (x) = = . We have
x2 − x − 6 (x − 3)(x + 2)
1 1 1 1 1 1 1 1
.f (x) = − = − · − · .
5 x−3 x+2 5 3 1 − x3 2 1 − − x2
3.4. Taylor Series and Maclaurin Series 199
1
∞ x n
. = , |x| < 3
1− x
3 n=0
3
and
1 x n ∞ x n ∞
. x = − = (−1)n , |x| < 2,
1 − −2 n=0
2 n=0
2
we can write the Maclaurin series of f as follows:
∞ n
1 1 x 1
∞ n
n x
f (x) = − − (−1)
5 3 n=0 3 2 n=0 2
.
∞
1 (−1)n+1 1
= − n+1 xn , |x| < 2.
n=0
5 2n+1 3
f (x) = (1 + x)α , α ∈ R,
. x > −1,
Next, we prove that .f (x) = g(x), .∀x ∈ ] − 1, 1[, that is, f is the sum of its
Maclaurin series in that interval.
Since a power series is differentiable term by term on the interval of conver-
gence, we obtain
∞
α(α − 1) . . . (α − n + 1) n−1
g (x) =
. x ,
n=1
(n − 1)!
and multiplying by x
∞
α(α − 1) . . . (α − n + 1) n
x g (x) =
. x .
n=1
(n − 1)!
200 Series of Functions
Then
g (x) + x g (x) =
.
∞ ∞
α(α − 1) . . . (α − n + 1) n−1 α(α − 1) . . . (α − n + 1) n
= x + x
n=1
(n − 1)! n=1
(n − 1)!
∞ ∞
α(α − 1) . . . (α − n) n α(α − 1) . . . (α − n + 1) n
= x + x
n=0
n! n=1
(n − 1)!
∞
α(α − 1) . . . (α − n) α(α − 1) . . . (α − n + 1)
=α+ + xn
n=1
n! (n − 1)!
∞
. α(α − 1) . . . (α − n + 1) α−n
=α+ + 1 xn
n=1
(n − 1)! n
∞
α(α − 1) . . . (α − n + 1) α n
=α+ · x
(n − 1)! n
n=1
∞
α(α − 1) . . . (α − n + 1) n
=α 1+ x
n=1
n!
= α g(x),
that is,
(1 + x) g (x) = α g(x).
. (3.3)
g(x)
Let us consider the function . and calculate its derivative:
(1 + x)α
g(x) g (x)(1 + x)α − α(1 + x)α−1 g(x)
=
(1 + x)α (1 + x)2α
.
(1 + x)α−1 (1 + x) g (x) − α g(x)
= .
(1 + x)2α
By equality (3.3) the numerator of this fraction is zero; therefore,
g(x)
. = 0,
(1 + x)α
which implies that the function defined by
g(x)
.
(1 + x)α
is constant on .] − 1, 1[, that is, .g(x) = c (1 + x)α , for some constant c.
Since .g(0) = 1, we obtain .c = 1, and we have
g(x) = (1 + x)α , ∀x ∈ ] − 1, 1[ .
.
3.5. Introduction to Fourier Series 201
a cos(kx) + bk sin(kx) , k = 1, 2, . . . ,
. k
The constants .a0 , .an , and .bn , .n ∈ N, are the coefficients of the trigono-
metric series.
But π π
1
. cos(nx) dx = sin(nx) =0 (3.5)
−π n −π
and π π
1
. sin(nx) dx = − cos(nx) = 0; (3.6)
−π n −π
3.5. Introduction to Fourier Series 203
therefore, π
. f (x) dx = πa0 .
−π
π
f (x) cos(mx) dx =
−π
∞
π
a0
= + an cos(nx) + bn sin(nx) cos(mx) dx
−π 2 n=1
π ∞ π
a0
.= cos(mx) dx + an cos(nx) + bn sin(nx) cos(mx) dx
2 −π n=1 −π
π
a0
= cos(mx) dx+
2 −π
∞ π π
+ an cos(nx) cos(mx) dx + bn sin(nx) cos(mx) dx .
n=1 −π −π
1
. cos(nx) cos(mx) = cos (n + m)x + cos (n − m)x (3.7)
2
1
sin(nx) cos(mx) = sin (n + m)x + sin (n − m)x , (3.8)
2
it follows
π
0, if n = m
. cos(nx) cos(mx) dx = (3.9)
−π π, if n = m
and π
. sin(nx) cos(mx) dx = 0. (3.10)
−π
π
f (x) sin(mx) dx =
−π
∞
π
a0
= + an cos(nx) + bn sin(nx) sin(mx) dx
−π 2 n=1
∞ π
a0 π
.= sin(mx) dx + an cos(nx) + bn sin(nx) sin(mx) dx
2 −π n=1 −π
π
a0
= sin(mx) dx+
2 −π
∞ π π
+ an cos(nx) sin(mx) dx + bn sin(nx) sin(mx) dx .
n=1 −π −π
1
. sin(nx) sin(mx) = cos (n − m)x − cos (n + m)x , (3.11)
2
we conclude that
π
0, if n = m
. sin(nx) sin(mx) dx = (3.12)
−π π, if n = m,
and we obtain
π
. f (x) sin(mx) dx = bm π.
−π
In the case where f is explicitly defined on the interval .[0, 2π] instead of
[−π, π], it may be more convenient to compute the coefficients using the
.
f˜(x) = f (x), ∀x ∈ ] − π, π]
.
f˜(x + 2π) = f˜(x), ∀x ∈ R.
Before we study the properties of the Fourier series, let us examine some
examples of how these series are calculated.
π 0 π
1 1
a0 = f (x) dx = 1 dx + x dx
Figure 3.11: The graph of π −π π −π 0
the function of Example 3.5.3 0 2 π
1 x π
= x + =1+ ;
π −π 2 0 2
0 π
1
an = cos(nx) dx + x cos(nx) dx
π −π 0
0 π π
1 sin(nx) sin(nx) sin(nx)
= + x − dx
π n −π n 0 0 n
π
1 cos(nx) 1 cos(nπ) cos(0) (−1)n − 1
. = = − = ;
π n2 0 π n2 n2 πn2
0 π
1
bn = sin(nx) dx + x sin(nx) dx
π −π 0
0 π
π
1 cos(nx) cos(nx) cos(nx)
= − + −x + dx
π n −π 0 n 0 n
π
1 1 cos(−nπ) cos(nπ) sin(nx)
= − + −π +
π n n n n2 0
With this example, it is easy to verify that the sum of the series is not equal
to the function at all points of the interval .[−π, π]. In fact, at .x = −π and
at .x = π, we obtain the numerical series
∞
1 π 2
. + + ,
2 4 n=1 π (2n − 1)2
and .f (−π) = 1 = f (π) = π.
1 cos(nπ) cos(−n π) 2
= −π −π = (−1)n+1 .
π n n n
∞
2
The Fourier series of f is . (−1)n+1 sin(nx).
n=1
n
In the following example, we have a periodic function with period .2π, defined
on .]0, 2π[ (see Proposition 3.5.3).
The graph of f is shown in Fig. 3.14 Using equalities (3.8) and (3.11), we
can calculate the Fourier coefficients as follows:
1 2π 1 π 2
a0 = f (x) dx = sin(x) dx = ;
π 0 π 0 π
1 2π 1 π
Figure 3.14: The graph of an = f (x) cos(nx) dx = sin(x) cos(nx) dx
the function of Example 3.5.6 π 0 π 0
⎧
⎪
⎨0, if n is odd
= 2
. ⎪
⎩ , if n is even;
π(1 − n2 )
1 2π 1 π
bn = f (x) sin(nx) dx = sin(x) sin(nx) dx
π 0 π 0
⎧
⎨0, if n = 1
= 1
⎩ , if n = 1.
2
We now introduce some concepts that will allow us to approach the study
of the convergence of the Fourier series, which is an essential problem of
this theory.
.f (x−
k ) = lim− f (x) and f (x+
k ) = lim+ f (x).
x→xk x→xk
f (x−
.
k ) = lim f (x) and f (x+
k ) = lim f (x)
x→x−
k x→x+
k
and
f (x−
.
k ) = lim f (x) and f (x+
k ) = lim f (x).
x→x−
k x→x+
k
continuity of f , the sum of the series is equal to the value of the function.
At each point of discontinuity of f , the sum of the series is equal to the
arithmetic mean of the left- and right-hand limits of the function at that
point.
Figure 3.16 shows some partial sums of the Fourier series of the function .f˜.
3.5. Introduction to Fourier Series 213
Figure 3.18 shows some partial sums of the Fourier series of the function f .
Example 3.5.9 In Example 3.5.3, we obtained the Fourier series for the
real-valued function of a real variable
⎧
⎨1, if − π ≤ x < 0
.f (x) =
⎩x, if 0 ≤ x ≤ π.
Let f be an even function. Taking into account that the functions .cos(nx)
are even and .sin(nx) are odd, for all .n ∈ N, by Proposition 3.5.5, we find Figure 3.20: An even func-
that .f (x) cos(nx) are even and .f (x) sin(nx) are odd, for all .n ∈ N. tion
216 Series of Functions
Example 3.5.10 Consider the even function .f (x) = |x| on the interval
.[−π, π] (Fig. 3.21). The Fourier coefficients of f are
2 π 2 π
a0 = f (x) dx = x dx = π;
π 0 π 0
2 π 2 π
an = f (x) cos(nx) dx = x cos(nx) dx
. π 0 π 0
⎧ 4
⎨− , if n is odd
= n2 π
⎩
0, if n is even, n = 0.
The analytical expression of the periodic extension of f to .R is
f˜(x) = |x − 2kπ|, x ∈ [(2k − 1)π, (2k + 1)π] k ∈ Z.
.
Let f be an odd function. Considering that .cos(nx) are even and .sin(nx)
are odd for all .n ∈ N, by Proposition 3.5.5, we obtain that .f (x) cos(nx)
are odd and .f (x) sin(nx) are even for all .n ∈ N. By Proposition 3.5.6, the
Fourier coefficients of f are
1 π
a0 = f (x) dx = 0;
π −π
1 π
. an = f (x) cos(nx) dx = 0;
π −π
π
1 2 π
bn = f (x) sin(nx) dx = f (x) sin(nx) dx.
π −π π 0
We can state the following proposition:
Figure 3.22: Graph of the extension of the function from Example 3.5.11
Then, we apply everything that has been stated for extensions as an even
function. Therefore, the Fourier coefficients can be computed by the for-
mulas of Proposition 3.5.7.
To develop f in a Fourier sine series, we take an odd extension of f from
the interval .[0, π] to the interval .[−π, π] (see Fig. 3.23):
⎧
⎨−f (−x), if − π < x < 0
.fodd (x) =
⎩f (x), if 0 ≤ x < π.
Example 3.5.12 Let us determine the Fourier sine series and the Fourier
cosine series of the function f defined by .f (x) = x3 −x, on .[0, π[ (Fig. 3.24).
The analytical expression of the even extension to the interval .] − π, π[ is
⎧
⎨−x3 + x, if − π < x < 0
.feven (x) =
⎩x3 − x, if 0 ≤ x < π.
Since .f˜even is continuous on .R, Theorem 3.5.1 allows us to affirm that the
following equality is satisfied, whatever .x ∈ R:
∞
π3 π 2 3π 2 6 + n2
f˜
. even (x) = − + (−1)n 2
+ 4
1 − (−1)n cos(nx).
4 2 π n=1
n n
f˜
. odd (x) = x3 − x, x ∈ ] − π, π[ .
3.5. Introduction to Fourier Series 221
f˜
. odd (x) = (x − 2kπ)3 − x − 2kπ, x ∈ ](2k − 1)π, (2k + 1)π[, k ∈ Z,
⎧
∞
1−π 2
6 ⎨f˜odd (x), if x = (2k + 1)π
.2 (−1)n + 3 sin(nx) =
n n ⎩0, if x = (2k + 1)π,
n=1
for all .k ∈ Z.
222 Series of Functions
Thus far, we have dealt with periodic functions of period .2π. Let us now
focus on functions f with period .T = 2l, .l > 0.
lt lt
Let .x = . The function .g(t) = f has period .2π. In fact,
π π
l(t + 2π) lt lt
g(t + 2π) = f
. =f + 2l =f = g(t).
π π π
where π
1
a0 = g(t) dt;
π −π
π
1
. an = g(t) cos(nt) dt;
π −π
π
1
bn = g(t) sin(nt) dt.
π −π
lt
Then the Fourier series of .f (x) = f is
π
a0 ∞ nπx nπx
. + an cos + bn sin ,
2 n=1
l l
where
π l
1 1 π 1 l
a0 = g(t) dt = f (x) ·
dx. = f (x) dx ;
π −π −l π l l −l
nπx π
1 π 1 l
an = g(t) cos(nt) dt = f (x) cos · dx
π −π π −l l l
l
1 nπx
. = f (x) cos dx ;
l −l l
nπx π
1 π 1 l
bn = g(t) sin(nt) dt = f (x) sin · dx
π −π π −l l l
l
1 nπx
= f (x) sin dx.
l −l l
3.5. Introduction to Fourier Series 223
The graph of f is shown in Fig. 3.27. The coefficients of its Fourier series
are given by
1 2 1 2
a0 = f (x) dx = 2 dx = 2;
2 −2 2 0
2 nπx 2 nπx
Figure 3.27: The graph of
1 1 the function of Example
an = f (x) cos dx = 2 cos dx = 0; 3.5.13
2 −2 2 2 0 2
. nπx nπx
1 2 1 2
bn = f (x) sin dx = 2 sin dx
2 −2 2 2 0 2
⎧
⎨0, if n is even
= 4
⎩ , if n is odd.
nπ
The Fourier series of f is
a0 ∞ nπx nπx
+ an cos + bn sin
2 n=1
2 2
.
∞
4 (2n − 1)πx
= 1+ sin .
n=1
(2n − 1)π 2
224 Series of Functions
We want to develop f into a Fourier sine series. For this, we take an odd
extension of f , which has period .2l = 4 (Fig. 3.28).
We often want to differentiate and integrate term by term the Fourier series
and, as we saw in Sect. 3.2, sufficient conditions to guarantee this procedure
involve the concept of uniform convergence. We have the following results.
∞
Theorem 3.5.3 If the series . |an |+|bn | converges, then the trigono-
n=1
metric series
a0
∞ nπx nπx
. + an cos + bn sin
2 n=1
l l
converges absolutely and uniformly on .R, and its sum is a continuous
function. The trigonometric series is the Fourier series of the sum func-
tion.
.[a, b].
226 Series of Functions
It should be noted that the previous theorem remains valid even if the
Fourier series is not pointwise convergent for the function f . By integrating
1
term by term, a factor of . is introduced in the series, which results in its
n
convergence.
ating term by term the Fourier series of f . In addition, the Fourier series
of .f converges to .f at the points of continuity of .f , and at each point
of discontinuity, the sum of the series is equal to the arithmetic mean of
the left- and right-hand derivatives of the function f at that point.
Note that the theorem regulating term by term differentiation is more de-
manding than that of term by term integration. Recall what we saw in
Sect. 3.2: Term by term differentiation of a series of functions can only be
guaranteed if the series of derivatives is uniformly convergent.
The previous theorems allow us to operate on the Fourier series similarly to
what was done with the Taylor series.
∞
2
.x= (−1)n+1 sin(nx), −π < x < π,
n=1
n
3.5. Introduction to Fourier Series 227
that is,
∞
x sin(nx)
. = (−1)n+1 , −π < x < π.
2 n=1 n
We can integrate this series term by term, obtaining
x ∞ x
t sin(nt)
dt = (−1)n+1 dt, −π < x < π
0 2 n=1 0 n
.
∞ ∞
x2 (−1)n+1 cos(nx)
⇔ = 2
− (−1)n+1 , −π < x < π.
4 n=1
n n=1
n2
The value of the numerical series can be computed by integrating this result
between .−π and .π:
π 2 ∞ ∞
x (−1)n+1 (−1)n+1 π
dx = 2π − cos(nx) dx
−π 4 n=1
n2 n=1
n2 −π
. ∞
(−1)n+1
= 2π .
n=1
n2
Therefore,
∞
π3 (−1)n+1
. = 2π ,
6 n=1
n2
that is,
∞
π2 (−1)n+1
. = .
12 n=1 n2
Thus, the Fourier series of .f (x) = x2 is
∞
π2 cos(nx)
x2 =
. +4 (−1)n , −π < x < π,
3 n=1
n2
and this series diverges whatever .x ∈ R (the general term is not a null
sequence).
This result does not contradict Theorem 3.5.8 because .f (−π) = f (π),
which shows the need for the condition of equality. The extension by peri-
odicity of f is discontinuous (see Example 3.5.1). We can observe that, in
Theorem 3.5.8, the continuity of f on .[−l, l] and the condition .f (−l) = f (l)
ensure the continuity of the extension by periodicity of f .
3.6. Solved Exercises 229
1
a) Find the domain of f . Find f (x) dx.
0
b) Check if f is continuous on its domain.
7. Consider the series of functions defined, for x ≥ 0,
2. Consider the real-valued function f of a real variable, ∞ x
1 n+ n
which is defined by by x+ .
n=1
n
∞
1
.f (x) = . a) Determine the values of x for which the series
n=0
1 + (3x)n
is convergent.
b) Let 0 < α < 1. Show that the series con-
a) Determine the domain of f .
verges uniformly on [0, α].
b) Show that f is continuous on [1, +∞[.
8. Let (an ) be a sequence of positive terms such that
3. Consider the real-valued function f of a real variable,
which is defined by an+1 1
.lim = .
∞
an 2
x
.f (x) = .
n=1
nx If, for each n ∈ N, fn : R → R is a function such
that
1
.|fn (x)| ≤ , ∀ x ∈ R,
a) Find the domain of f . n
b) Show that f is continuous on ]1, +∞[. ∞
show that the series of functions an fn (x) con-
4. Consider the real-valued function f of a real variable, n=1
which is defined by verges uniformly on R.
∞
1 9. Let fn : [0, 1] → R, n ∈ N, and f : [0, 1] → R be
.f (x) = √ .
(n x + 1) n3 + 1 continuous functions such that
n=1
2 ∞
Calculate f (x) dx. .f (x) = fn (x), ∀x ∈ [0, 1].
1 n=1
5. Consider the real-valued function f of a real variable,
Assuming that
which is defined by
1
∞ 1 1
cos(nx) (i) fn (x) dx = −
.f (x) = . n n+1
n2 0
n=1 1
Show that (ii) f (x) dx = 0
0
π/2 ∞
(−1)n
. f (x) dx = . show that the series of functions does not converge
0 (2n + 1)3
n=0 uniformly on [0, 1].
230 Series of Functions
∞
10. Consider the real-valued function f of a real variable, 3n log(n) n
c) x
which is defined by n=1
7n n2
∞
∞ 2 × 5 × · · · × (3n − 1) n
arccot(nx) d) x
.f (x) = √ . (n + 2)!
n=1 n5 n=1
∞
log(n) n
a) Determine the domain of f . e) x
n=1
n
b) Demonstrate that f is continuous on its do- ∞
3 n2
main. f) xn
n=0
4n (n3 + 2)
c) Show that f is differentiable on its domain.
∞
1 × 4 × · · · × n2 n
g) x
11. Let (an ) be a sequence of positive terms. Show (2 n)!
∞ n=1
that an cos(nx) converges uniformly on R if, ∞ n
1
n=0 h) nn sin 2
xn
∞
n=1
n
and only if, an converges. ∞
1
n=0 i) xn
n=1
2n n arctan(n) − π
2
12. Consider a sequence of functions fn , continu- ∞
n∈N
3 × 6 × · · · × 3n n
ous on R, such that j) x
n=1
(n + 3)!
.|fn (x)| ≤ an , ∀n ∈ N, ∀x ∈ R, ∞
n+1
k) (−1)n √ xn
where an is the general term of a convergent series n=0 e2n + 5
of positive terms. Let (bn )n∈N be a sequence such ∞
1 × 3 × 5 × · · · × (2n − 1) 1 n
∞ l) · x
that the series (bn − bn+1 ) converges. Show n=1
2 × 4 × 6 × · · · × 2n n
n=1
∞
that the series of functions bn fn (x) defines a 15. Find the real values of x for which the following se-
n=1 ries are absolutely convergent, conditionally conver-
continuous function on R.
gent, and divergent. Explain the reasoning behind
∞
the response.
13. a) Show that the series (−1)n (1 − x)xn con-
n=0 ∞
2n
n
verges uniformly on [0, 1]. a) (x − 1)n
n=1
2n + 1
∞
(1 − x)xn converges ∞
b) Show that the series (x − 2)n
n=0 b) (−1)n
n log(n)
pointwise, but not uniformly, on [0, 1]. n=2
∞
n
14. Find the real values of x for which the following series c) (x + 4)n
n=1
(−3)n (2n2 − 1)
are absolutely convergent, conditionally convergent,
∞
and divergent. Give a justification for the answers en
d) (x − e)n
provided. log(n)
n=2
∞ ∞ π
π n xn cos 2n
a) e)
n=2
n log2 (n) n=1
(2n − 1)2 (x − 1)n
∞
∞
2n nn (2 x2 − 5)n
b) xn f)
n=1
(n + 1)n+1 n=1
n 3n+1
3.6. Solved Exercises 231
∞ √ √ ∞
n+1− n
g) √ √ (x + 2)n 20. Assume that the series an 5n converges. Is the
n=1
n+1+ n n=1
n ∞
∞
log2 (n) 2x series an (−2)n convergent? Provide a justifi-
h)
n=1
n2 x2 + 1 n=1
cation for the answer.
∞
1 1 ∞
i) √ · n
2n n + 1 x 21. Can the power series an (x − 3)n converge at
n=1
n=0
∞
2 1 n the point x = 0 and diverge at the point x = 1?
j) + (2x − 3)n
3 2n
n=1 ∞
∞ 2 bn xn be a power series whose radius of con-
23n (n − 1)! 22. Let
k) x2n n=0
n=2
(2n)! vergence is R > 1. Let fn n∈N be a sequence of
∞
functions such that
2n − 1 2n
l) (x − 3)n
n=0
3n + 1 .|fn (x)| ≤ |bn |, ∀x ∈ R.
∞ π
sin n What can be said about the uniform convergence of
m) (x − 1)2n ∞
4n (n + 1)
n=1 the series of functions fn ?
n=0
16. Find the values
of x for which the series
∞ 23. Let (an ) be a sequence of positive terms such that
x2n x2 ∞
√ log 1 + √ is convergent. Justify the
n=1
n n the series an (x − 1)n is conditionally conver-
answer. n=1
gent at x = −1.
17. Compute the radius of convergence of the series a) What is the radius of convergence of the power
∞ π n ∞
cos + 2 − (−1)n x2n and indicate the series an (x − 1)n ? Justify.
n=1
n
n=1
largest open interval on which the series converges.
b) Show that x = −1 is the only point at which
∞
the series is conditionally convergent.
4n
18. Consider the power series n + 5n
xn . Know-
2 ∞
n=1 1
ing that the radius of convergence of this series is 24. Consider the series (x − 2)n .
5 n=1
n 5n
, indicate, without resorting to additional calcu-
4 a) Find the real values of x for which the series
lations, whether the series converges or diverges at
points 1 and 2. Investigate the convergence of the is absolutely convergent, conditionally conver-
5 gent, and divergent. Justify.
series at the point .
4 b) Let f be the sum of the series. Find the set
19. Find the real values of x for which the series of points where f is differentiable and deter-
∞
mine the analytical expression of f . Justify
an xn is absolutely convergent, conditionally
the answer given.
n=0
convergent, and divergent, where ∞
(−1)n n+1
25. Consider the series log(x) .
n=0
n + 1
1, if n is even
.an =
2, if n is odd. a) Find the real values of x for which the series
is absolutely convergent, conditionally conver-
Justify the answer. gent, and divergent. Justify.
232 Series of Functions
5x − 1 x
b) Using the series of derivatives, calculate, for a) g) cos(t2 ) dt
x2 − x − 2 0
each x, the sum of the series.
x
b)
2
et dt x + x2
h)
26. Consider the real-valued function f of a real variable, 0 (1 − x)2
which is defined by 1 1 − e−tx i) x arctan(x)
c) dt
0 t
∞
2x
(−1)n
.f (x) = x2n . x j)
22n (n!)2 d) arctan(t) dt (1 + x2 )2
n=0
0
k) log(2 + x3 )
1
e)
(2x + 5)2 l) log 1+x
a) Determine the domain of f . 1−x
2
b) Show that x2 f (x) + xf (x) + x2 f (x) = 0. f) + e2x m) log(4 − x2 )
3 + 2x
Justify the answer given.
30. a) Determine Maclaurin series of the function
a) Compute the Maclaurin series of f , indicating 36. Represent in a power series of x − 3 the real-valued
the values of x for which the development is function f of a real variable defined by
valid. 1
.f (x) = + log(4 − x).
b) Using the expansion obtained in item a), indi- 4−x
cate the value of f (17) (0). Determine the values of x for which the development
is valid. Give a clear explanation of the answer.
33. Using only the Maclaurin expansions of the functions
1
f (x) = sin(x), g(x) = ex , and h(x) = , find 37. Expand in a power series of x − 3 the real-valued
1−x function f of a real variable defined by
the sum of the following series:
1
∞
.f (x) = .
1 n x2
a) (−1)
(2n + 1)!
n=0 Determine the values of x for which the development
∞ is valid. Provide a clear explanation of the answer.
2n
b)
n=0
n! Hint: Use term by term differentiation.
∞
1 38. Develop in a power series of x + 1 the real-valued
c) (−1)n
n=1
2n n function of a real variable f defined by
1
34. Consider the function f : R → R defined by .f (x) = .
x2 + x − 6
.f (x) = log(x2 + 1) + cos(2x). Indicate the values of x for which the expansion is
valid. Justify.
b) Using item a) compute the sum of the series Find the power series of x − 1 of f , indicating the
values of x for which the development is valid.
∞
π 2n 1 22n
.1 + (−1)n−1 − . 40. Consider the function
n=1
42n n (2n)! x
arctan(t − 2) − t + 2
.f (x) = dt.
2 (t − 2)2
35. Consider the function f : R → R defined by
Expand f in a power series of x − 2, indicating the
2x values of x for which the development is valid.
.f (x) = sin(2x) + .
(1 − x2 )2 41. Let f be a function of class C ∞ in a neighborhood
of the origin that satisfies the following conditions:
b) 2 − x2 , 0 < x < 1
50. Consider the function f (x) = π − x. 54. Consider the Fourier series of the periodic function
of period 2π, f˜, defined on the interval [−π, π] by
a) Develop the function into a Fourier sine series f (x) = |x|,
on ]0, π[. Use the result to calculate the sum
∞
of the series π 2 (−1)n − 1
. + cos(nx).
∞
sin(nx) 2 π n=1 n2
. , 0 < x < 2π.
n=1
n
∞
b) Using the result from the previous item, cal- a) Show that the series n an sin(nx), where
culate the sum of the series n=1
an are the Fourier coefficients of the original
∞
cos(nx) series, is convergent on R.
. , 0 < x < 2 π. ∞
n=1
n2
b) Let g(x) = − n an sin(nx). Sketch the
n=1
c) Develop the function into a Fourier cosine se- graph of g on the interval [−2π, 2π].
ries on ]0, π[.
∞ c) Calculate the sum of the series
(−1)n
d) Calculate the sum of the series . ∞ ∞
n=1
n2 1 (−1)n
. and .
n=1
(2n − 1)2 n=1
(2n − 1)3
51. Consider the function f (x) = x.
a) Develop the function into a Fourier sine series 55. Determine, using Fourier series, the function f such
on ]0, π[. Use the result to calculate the sum that
of the series
∞
f (x) + 3f (x) = 3x, 0<x<1
sin(nx) .
. (−1)n+1 , −π < x < π. f (0) = f (1) = 0.
n=1
n
b) Develop the function into a Fourier cosine se- 56. Consider the periodic function with period 2 defined
ries on ]0, π[. by f (x) = x2 on the interval ]0, 2].
52. Develop into a Fourier series the periodic function of a) Determine the Fourier series of f .
period 2π, defined on ] − π, π[ by f (x) = | sin(x)|, b) Show that the term by term differentiation of
and use the result to show that the series obtained in item a) does not con-
∞
1 1 ∞
(−1)n 1 π verge to f .
. = and = − .
n=1
4n2 −1 2 n=1
4n 2−1 2 4 57. Consider the periodic function with period 6 defined
on ] − 3, 3[ by f (x) = x − x2 .
Hint: sin(a) cos(b) = 1
2
sin(a + b) + sin(a − b) .
a) Find the Fourier series of f .
53. Show that on the interval ]0, π[ we have
b) Let S(x) be the sum of the series
∞
π4 cos(2nx)
.x (π − x) = −3
2 2
a0 x 3 nπx nπx
, ∞
30 n=1
n4 . + an sin − bn cos ,
2 n=1
nπ 3 3
and from this result, we obtain
∞ ∞
where a0 , an , and bn are the Fourier coeffi-
1 π4 1 π4
. = and = . cients of f . Determine the expression of S(x),
n 4 90 (2n + 1) 4 96
n=1 n=0 x ∈ ] − 3, 3[.
236 Series of Functions
SOLUTIONS
∞
sin(n2 x)
1. a) The terms of the series √5
are functions of domain R. The function f is the sum of the series;
n=1 n7 + 3
therefore, its domain is the subset of R where the series converges. The general term satisfies
sin(n2 x)
. √ ≤ √ 1 , ∀x ∈ R, ∀n ∈ N, (3.13)
5
n + 3
7 5
n7 + 3
since | sin(x)| ≤ 1, ∀x ∈ R.
∞
1
We test the convergence of the series of positive terms, √
5
, by comparison with the series
n=1 n7 + 3
∞
1
, which we know to be convergent. As the limit
n=1
n7/5
1
√
5
n7 + 3 5 n7
. lim = lim =1
1 n7 + 3
n7/5
is finite and different from zero, we conclude, by Corollary 2 of the General Comparison Test, that the series
∞ ∞
1 sin(n2 x)
√ √
5
n7 + 3
is convergent. By inequality (3.13), the series 5 n7 + 3 converges, ∀x ∈ R. Then, the
n=1 n=1
∞
sin(n2 x)
series √5
is convergent, ∀x ∈ R, that is, the domain of f is R.
n=1 n7 + 3
b) To conclude that f is continuous on R, it is sufficient (see Theorem 3.2.3) that the following two conditions
are satisfied:
sin(n2 x)
(i) The functions fn (x) = √ 5
are continuous on R.
n7 + 3
∞
sin(n2 x)
(ii) The series √5
is uniformly convergent to f on R.
n=1 n7 + 3
The first condition is immediately satisfied because the functions hn (x) = sin(n2 x) are continuous on
R. The second is a consequence of the previous item: Inequality (3.13) and the convergence of the se-
∞ ∞
1 sin(n2 x)
ries √
5
are sufficient to conclude the uniform convergence of the series √5
by the
7
n +3 n7 + 3
n=1 n=1
Weierstrass’ test. As the series converges to f , this function is continuous.
∞
1 1
2. a) The terms of the series n
are functions with domain R, if n is even, and R \ − , if n is
n=0
1 + (3x) 3
1
odd. The function f is the sum of the series, so its domain is the subset of R \ − where the series
3
converges. For the series to converge, its general term must be a null sequence. But
⎧
⎪
⎪1, if |3x| < 1
⎪
⎨
1 0, if |3x| > 1
. lim =
1 + (3x)n ⎪
⎪1
⎪
⎩ , if 3x = 1;
2
3.6. Solved Exercises 237
1 1
therefore, if |x| ≤ , x = − , the series diverges.
3 3
∞
1
Let x be such that |3x| > 1. We study the series of positive terms
1 + (3x)n using D’Alembert’s
n=0
test:
1 1 1
+
1 + (3x)n+1 1 + (3x)
n (3x)n+1
. lim = lim = lim 3x = 1 .
1
1 + (3x)n+1
1
|3x|
1 + (3x)n (3x)n+1 + 1
1 1
We conclude that the series converges because < 1, that is, |x| > .
|3x| 3
1 1
The domain of f is −∞, − ∪ , +∞ .
3 3
b) To show that f is continuous on [1, +∞[, it is sufficient (see Theorem 3.2.3) that the following two conditions
are met:
1
(i) The functions fn (x) = are continuous on [1, +∞[.
1 + (3x)n
∞
1
(ii) The series is uniformly convergent to f on [1, +∞[.
n=1
1 + (3x)n
The first condition is immediately satisfied. The second condition follows from the inequality
n
1
. ≤ 1 ≤ 1 , ∀x ∈ [1, +∞[, ∀n ∈ N,
1 + (3x)n (3x)n 3
∞ n
1
since the series is convergent, which, by the Weierstrass’ test, is sufficient to conclude the
n=1
3
∞
1
uniform convergence of the series . As the series converges to f , this function is continuous.
n=1
1 + (3x)n
∞
x
3. a) Let f (x) = . The terms of the series are functions with domain R. The function f is the sum of
n=1
nx
the series; therefore, its domain is the subset of R where the series converges.
The functions gn (x) = e−x log(n) are continuous, since the exponential is continuous on R. Let us look at
the second condition. Let a > 1 be a real constant. The inequality
1
. ≤ 1 , ∀x ∈ [a, +∞[, ∀n ∈ N,
nx na
∞
1
combined with the fact that the series is convergent, is sufficient to conclude, by Weierstrass’ test,
n=1
na
∞
1
the uniform convergence of the series x
on [a, +∞[. As the series converges to g, this function is
n=1
n
continuous on [a, +∞[. Since a is an arbitrary constant greater than 1, we also have g continuous on
]1, +∞[.
∞
1
4. The terms of the series √ are continuous functions on [1, 2]. As
n=1 (nx + 1) n3 + 1
1 1 1 1 1 1
. √ = ·√ ≤ ·√ ≤ √ , ∀x ∈ [1, 2], ∀n ∈ N,
(nx + 1) n3 + 1 |nx + 1| n3 + 1 n+1 n3 + 1 n3
∞
1 3
and the series of positive terms √ is convergent because it is a p-series with p = , by Weierstrass’ test,
n=1 n3 2
∞
1
the series √ converges uniformly to f , on [1, 2]. It follows from Theorem 3.2.3 that f is
n=1 (nx + 1) n3 + 1
continuous on [1, 2]; by Theorem 3.2.4, the series is integrable term by term on this interval and
2 2 ∞
∞ 2
1 1 1
f (x) dx = √ dx = √ dx
1 1 n=1 (nx + 1) n3 + 1 n=1 n 3 + 1 1 nx + 1
∞
2 ∞
1 log(nx + 1) log(2n + 1) − log(n + 1)
. = √ = √
n=1 n3 +1 n 1 n=1 n n3 + 1
∞
1 2n + 1
= √ log .
n=1 n n3 + 1 n+1
∞
cos(nx)
5. The terms of the series are functions of domain R. The function f is the sum of the series; therefore,
n=1
n2
its domain is the subset of R where the series converges. Since we have
cos(nx)
. ≤ 1 , ∀x ∈ R, ∀n ∈ N,
n2 n2
∞
1
because | cos(nx)| ≤ 1, ∀x ∈ R, ∀n ∈ N, and the series of positive terms 2
is convergent as it is a p-
n=1
n
∞
cos(nx)
series with p = 2, by Weierstrass’ test, the series converges uniformly to f , on R. Consequently, by
n=1
n2
cos(nx)
Theorem 3.2.3, given that the functions are continuous on R, f is continuous on R; by Theorem 3.2.4,
n2
3.6. Solved Exercises 239
π
the series is integrable term by term on every closed bounded interval of R. In particular, f is integrable on 0, ,
2
and
π/2 π/2 ∞ ∞ π/2
cos(nx) 1
f (x) dx = dx = cos(nx) dx
0 0 n=1
n2 n=1
n2 0
.
∞ π ∞ ∞ ∞
1 sin(nx) 2 sin( nπ
2
) (−1)n−1 (−1)n
= = = = ,
n=1
n 2 n 0 n=1
n 3
n=1
(2n − 1) 3
n=0
(2n + 1)3
because
⎧
nπ ⎪
⎨0, if n is even
. sin = −1, if n = 3, 7, 11, . . .
2 ⎪
⎩
1, if n = 1, 5, 9, . . .
∞
e−nx
6. The terms of the series f (x) = are functions of domain R. As
n=0
n4 + 5
−nx
e 1 1
. ≤ ≤ 4, ∀x ∈ [0, 1], ∀n ∈ N0 ,
n + 5
4 n4 + 5 n
∞
1
since |e−nx | = e−nx ≤ e0 = 1, ∀x ∈ [0, 1], ∀n ∈ N0 , and the series of positive terms 4
is convergent
n=1
n
∞
e−nx
as it is a p-series with p = 4. By Weierstrass’ test, the series 4+5
converges uniformly to f , on [0, 1].
n=0
n
e−nx
Therefore, by Theorem 3.2.3 since the functions fn (x) = are continuous on [0, 1], f is continuous on
n4 + 5
[0, 1]. Theorem 3.2.4 shows that the series is integrable term by term on this interval. We have
∞
∞ 1
1 1 1 e−nx 1 1
f (x) dx = + 4
dx = + 4
e−nx dx
0 0 5 n=1 n + 5 5 n=1 n + 5 0
.
∞
−nx 1 ∞
1 1 e 1 1 − e−n
= + 4
− = + .
5 n=1 n + 5 n 0 5 n=1 n (n4 + 5)
∞
n+ x
1 n
. x+
n=1
n
is of positive terms. Let us study this series by applying Cauchy’s Root Test:
n+ x x
n 1 n n 1 n 1 n
lim x+ = lim x+ x+
n n n
.
x x
1 n 1 n 1 1 n2
= lim x + x+ = lim x + x+ = x.
n n n n
240 Series of Functions
By Cauchy’s Root Test, the series converges if x < 1 and diverges if x > 1. It remains to see what happens
∞ 1
1 n+ n
if x = 1. In this case, we obtain the numerical series 1+ , which is divergent because its
n=1
n
general term is not a null sequence. In fact,
1 1
1 n+ n 1 n 1 n
. lim 1+ = lim 1 + 1+ = e.
n n n
∞ α
1 n+ n
As shown in item a), the numerical series α+ is convergent. By Weierstrass’ test, the original
n=1
n
series converges uniformly on [0, α].
∞
8. We will use Weierstrass’ test to prove that the series of functions an fn (x) converges uniformly on R. Since
n=1
1
|fn (x)| ≤ , ∀n ∈ N, ∀ x ∈ R, and (an ) is a sequence of positive terms, we have
n
an
.|an fn (x)| ≤ , ∀n ∈ N, ∀ x ∈ R.
n
∞
an
We prove that the series of positive terms is convergent by applying D’Alembert’s test:
n=1
n
an+1
n+1 n an+1 n an+1 1
. lim a
n
= lim = lim · =
(n + 1) an n+1 an 2
n
an+1 1
because, by hypothesis, lim = . As the limit is less than 1, the series is convergent. By Weierstrass’ test,
an 2
∞
the series of functions an fn (x) converges uniformly on R.
n=1
9. Since the functions fn : [0, 1] → R are continuous, if the convergence was uniform, we would have, by Theo-
rem 3.2.4,
1
∞ ∞ 1
. fn (x) dx = fn (x) dx.
0 n=1 n=1 0
But
∞
1 1
. fn (x) dx = f (x) dx = 0
0 n=1 0
and
∞
1 ∞
1 1
. fn (x) dx = − = 1.
n=1 0 n=1
n n+1
∞
arccot(nx)
10. a) The terms of the series √ are functions with domain R. The function f is the sum of the
n=1 n5
series; therefore, its domain is the subset of R where the series converges. The terms of the series satisfy
arccot(nx)
. √ ≤ √π , ∀x ∈ R, ∀n ∈ N, (3.14)
n5 n5
∞
1
since |arccot(nx)| ≤ π, ∀x ∈ R. The series of positive terms √ is convergent because it is a
n=1 n5
∞
5 π
p-series with p = . Therefore, the series √ is also convergent. By inequality (3.14), the series
2 n=1 n5
∞ ∞
arccot(nx) arccot(nx)
√ converges, ∀x ∈ R. Then the series √ is convergent, ∀x ∈ R, that is, the
n=1 n5 n=1 n5
domain of f is R.
b) To conclude that f is continuous on R, it is sufficient, according to Theorem 3.2.3, that the following two
conditions are satisfied:
arccot(nx)
(i) The functions fn (x) = √ are continuous on R.
n5
∞
arccot(nx)
(ii) The series √ is uniformly convergent to f on R.
n=1 n5
The functions hn (x) = arccot(nx) are continuous on R, which implies the first condition. The second condi-
∞
π
tion is a consequence of the previous item: Inequality (3.14) and the fact that the series √ is conver-
n=1 n5
∞
arccot(nx)
gent are sufficient, by Weierstrass’ test, to conclude the uniform convergence of the series √ .
n=1 n5
As the series converges to f , this function is continuous.
c) To conclude that f is differentiable on R, it is sufficient, by Corollary 2 of Theorem 3.2.4, that the following
conditions are fulfilled:
∞
∞
1 3 arccot(nx)
and the series √ is convergent because it is a p-series with p = , so the series √
n=1 n3 2 n=1 n5
is uniformly convergent. We can conclude that the series converges to f , that is,
242 Series of Functions
∞ ∞
arccot(nx) n
.f (x) = √ = − √ ,
n 5 (1 + n 2 x2 ) n 5
n=1 n=1
∞
11. If the series an cos(nx) converges uniformly on R, then it converges pointwise on R. In particular, the series
n=0
∞
is convergent at x = 0. At this point, we obtain the numerical series an , which is convergent.
n=0
∞
Conversely, suppose that the series an is convergent. We have
n=0
∞
By Weierstrass’ test the series an cos(nx) is uniformly convergent on R.
n=0
+∞
n
12. We know that the series (bn − bn+1 ) converges, that is, the sequence Sn = (bk − bk+1 ) = b1 − bn+1
n=1 k=1
converges. Therefore, there exists b = lim bn = b1 − lim Sn , so the sequence (bn ) is bounded, that is, there exists
M > 0 such that |bn | ≤ M, ∀n ∈ N. Based on this fact, we have the following inequalities:
∞
Thus, by Weierstrass’ test, the series bn fn (x) is uniformly convergent on R, defining a function f of domain
n=1
R. Since, by hypothesis, the functions fn are continuous on R, ∀n ∈ N, the function f is continuous on its domain.
∞
13. a) Consider the series (−1)n (1 − x) xn , x ∈ [0, 1].
n=0
∞
1−x
Then the series (−1)n (1 − x) xn converges pointwise to the function f (x) = on [0, 1]. This
n=0
1+x
convergence will be uniform if
. lim sup |f (x) − Sn (x)| = 0,
n→+∞ x∈[0,1]
1 − (−x)n+1 1−x
– If x = 1, Sn (x) = (1 − x) · = · (1 − (−x)n+1 ), and therefore,
1 − (−x) 1+x
1 − x 1−x
|f (x) − Sn (x)| = − · 1 − (−x)n+1
1+x 1+x
1−x
= · 1 − 1 − (−x)n+1
.
1+x
1 − x 1−x
= · (−x)n+1 = · xn+1 .
1+x 1+x
1−x
Let g(x) = |f (x) − Sn (x)| = · xn+1 .
1+x
The function g is continuous on [0, 1]; therefore, by Weierstrass’ Theorem, it has a maximum and a minimum
on this interval. As g is nonnegative and g(0) = g(1) = 0, the maximum is attained at some point within
the interior of the interval. The function g is differentiable on ]0, 1[, so this point corresponds to a zero of
the first derivative.
1−x xn
.g (x) = · xn+1 = − 2
· (n + 1)x2 + 2x − (n + 1) ,
1+x (1 + x)
and on the interval ]0, 1[,
−1 + 1 + (n + 1)2
.g (x) = 0 ⇔ (n + 1)x2 + 2x − (n + 1) = 0 ⇔ x = .
n+1
Then
−1 +1 + (n + 1)2
1−
−1 + 1 + (n + 1)2 n+1
. sup |f (x) − Sn (x)| = g ≤ ,
x∈[0,1] n+1 −1 + 1 + (n + 1)2
1+
n+1
and, as
−1 +1 + (n + 1)2
1−
n+1
. lim = 0,
n→+∞ −1 + 1 + (n + 1)2
1+
n+1
we can conclude that
. lim sup |f (x) − Sn (x)| = 0,
n→+∞ x∈[0,1]
∞
that is, the series (−1)n (1 − x) xn converges uniformly to the function f , on the interval [0, 1].
n=0
∞
b) Let us consider the series (1 − x) xn , x ∈ [0, 1].
n=0
– If x = 1, all terms are zero. Hence, the series converges.
∞
– If x = 1, as the series (1 − x) xn is geometric with ratio r = x, the series converges because
n=0
x ∈ [0, 1[. In this case,
∞
1
. (1 − x) xn = (1 − x) · = 1.
n=0
1−x
244 Series of Functions
∞
Then, the series (1 − x) xn , x ∈ [0, 1], converges pointwise to the function
n=0
0, if x = 1
.f (x) =
1, if x ∈ [0, 1[.
Let fn (x) = (1 − x)xn . The functions fn are continuous on [0, 1], ∀n ∈ N0 , because they are polynomial
functions. If the series was uniformly convergent on the interval [0, 1], f would be continuous on this interval.
∞
Since f is not continuous at x = 1 because lim f (x) = 1 and f (1) = 0, the series (1 − x) xn does
x→1−
n=0
not converge uniformly to f on [0, 1].
∞
π n xn πn
14. a) Let us study the power series of x, 2
. Setting an = , the radius of convergence of
n=2
n log (n) n log2 (n)
this series is, by Corollary 1 of Theorem 3.3.1
πn
an 2 n + 1 log2 (n + 1)
n log (n) 1
.r = lim = lim
= lim · = ,
an+1 π n+1 π n log 2
(n) π
(n + 1) log2 (n + 1)
1 1
which implies, by the same theorem, that the series is absolutely convergent if x ∈ − , and divergent
π π
1 1
if x ∈ −∞, − ∪ , +∞ .
π π
1
At x = ± , we have to study the corresponding numerical series.
π
∞ n ∞
1 πn 1 1
– If x = , we obtain the series 2
= 2
. Let us study the convergence of
π n=2
n log (n) π n=2
n log (n)
this series using the Integral Test.
1
Let f (x) = . We have:
x log2 (x)
(i) f (x) > 0, ∀x ≥ 2.
(ii) f is continuous on [2, +∞[.
(iii) f is decreasing on [2, +∞[ because x log2 (x) is positive and increasing.
∞ +∞
1 1
So the numerical series 2
converges if and only if the improper integral 2
dx
n=2
n log (n) 2 x log (x)
converges. We have
x
+∞ 1 1 x 1 −2
dx = lim dt = lim log(t) dt
. 2 x log2 (x) x→+∞ 2 t log2 (t) x→+∞ 2 t
x
1 1 1 1
= lim − = lim − + = ,
x→+∞ log(t) 2 x→+∞ log(x) log(2) log(2)
that is, the improper integral is convergent, so the same happens to the series. As it is a series of
positive terms, it converges absolutely.
3.6. Solved Exercises 245
∞ ∞
1 πn 1 n 1
– If x = − , we obtain the series 2
− = (−1)n 2
, which is alternating.
π n=2
n log (n) π n=2
n log (n)
As we just saw, the series of modules is convergent. Then the series is absolutely convergent.
1 1
Conclusion: The original series diverges if x ∈ −∞, − ∪ , +∞ and is absolutely convergent if
π π
1 1
x∈ − , .
π π
2
log2 (n + 1) log(n + 1)
Remark: To calculate the lim = lim . we proceed as follows: Consider the func-
log2 (n) log(n)
log(x + 1) log(x + 1)
tion g(x) = ; we use L’Hôpital’s Rule applied to the calculation of the limit lim ,
log(x) x→+∞ log(x)
∞
where the indeterminate form of type arises:
∞
1
log(x + 1) x+1 x
. lim = lim = lim = 1.
x→+∞ log(x) x→+∞ 1 x→+∞ x + 1
log(x + 1) log2 (n + 1)
Therefore, lim = 1, which implies that lim = 1.
x→+∞ log(x) log2 (n)
∞
2n nn 2n nn
b) Let us study the power series of x, n+1
xn . Setting an = , the radius of conver-
n=1
(n + 1) (n + 1)n+1
gence is by Corollary 1 of Theorem 3.3.1
2n nn
an 2n nn (n + 2)n+2
(n + 1) n+1
r = lim = lim n+1 = lim
an+1 2 (n + 1)n+1 (n + 1)n+1 2n+1 (n + 1)n+1
. (n + 2)n+2
n
1 n n + 2 n+2 1 1 1
= lim · = · ·e= .
2 n+1 n+1 2 e 2
1 1 1 1
Then, the series is absolutely convergent if x ∈ − , and divergent if x ∈ −∞, − ∪ , +∞ .
2 2 2 2
Let us study the convergence of the numerical series corresponding to the endpoints of the interval of
convergence.
∞ n ∞
1 2n nn 1 nn
– If x = , we obtain the series of positive terms n+1
· = .
2 n=1
(n + 1) 2 n=1
(n + 1)n+1
We will study this series by comparing it with the harmonic series, which is divergent. The limit
nn
n+1 n
(n + 1)n+1 n n n 1
. lim = lim = lim · =
1 n+1 n+1 n+1 e
n
is finite and different from zero; therefore, by Corollary 2 of the General Comparison Test, the series is
divergent.
246 Series of Functions
∞ ∞
1 2n nn 1 n nn
– If x = − , we obtain the series n+1
· − = (−1)n , which is alter-
2 n=1
(n + 1) 2 n=1
(n + 1)n+1
n n
nating, since > 0, ∀n ∈ N. We proved in the previous case that the series of modules is
(n + 1)n+1
divergent. Therefore, we need to verify whether the following conditions hold to apply Leibniz’s test:
n
nn n 1 1
(i) lim = lim · = × 0 = 0.
(n + 1)n+1 n+1 n+1 e
nn
(ii) > 0, ∀n ≥ 1.
(n + 1)n+1
n
nn n 1
(iii) The sequence of general term n+1
= · is decreasing, as it is the product
(n + 1) n+1 n+1
of two decreasing positive sequences.
Therefore, the series is convergent. It is conditionally convergent since the series of modules diverges.
1 1
Conclusion: The power series of x is absolutely convergent if x ∈ − , , conditionally convergent if
2 2
1 1 1
x = − , and divergent if x ∈ −∞, − ∪ , +∞ .
2 2 2
∞
3n log(n)
c) By Corollary 1 of Theorem 3.3.1, the power series of x, an xn , with an = , has a radius of
n=1
7n n2
convergence
3n log(n)
an 7(n + 1)2 log(n)
7n n2 7
.r = lim = lim
= lim = .
a 3n+1 log(n + 1) 3n2 log(n + 1) 3
n+1
7n+1 (n + 1)2
7 7 7 7
Then, the series converges absolutely if x ∈ − , and diverges if x ∈ −∞, − ∪ , +∞ .
3 3 3 3
log(n)
. lim
n2 = lim log(n) = 0,
1 n1/2
n3/2
∞
log(n)
by Corollary 3 of the General Comparison Test, the series is convergent. As it is the series
n=1
n2
∞
log(n)
of modules of (−1)n , this series is absolutely convergent.
n=1
n2
3.6. Solved Exercises 247
∞
log(n)
7
– If x = , we obtain the series , which we proved in the previous case to be convergent. As
3 n=1
n2
it is of positive terms, it is absolutely convergent.
7 7 7 7
Conclusion: The series converges absolutely if x ∈ − , and diverges if x ∈ −∞, − ∪ , +∞ .
3 3 3 3
∞
2 × 5 × · · · × (3n − 1) n 2 × 5 × · · · × (3n − 1)
d) Let us study the power series of x, x . Setting an = ,
n=1
(n + 2)! (n + 2)!
the radius of convergence is by Corollary 1 of Theorem 3.3.1
2 × 5 × · · · × (3n − 1)
an
.r = lim = lim (n + 2)! = lim n + 3 = 1 .
2 × 5 × · · · × (3n − 1)(3n + 2)
an+1 3n + 2 3
(n + 3)!
1 1 1 1
Then, the series converges absolutely if x ∈ − , and diverges if x ∈ −∞, − ∪ , +∞ .
3 3 3 3
Let us study the behavior of the numerical series at the endpoints of the convergence interval.
1
– If x = − , we obtain the alternating series
3
∞ ∞
2 × 5 × · · · × (3n − 1) 1 n 2 × 5 × · · · × (3n − 1)
. · − = (−1)n .
n=1
(n + 2)! 3 n=1
3n (n + 2)!
∞ ∞
2 × 5 × · · · × (3n − 1)
. = bn ,
n=1
3n (n + 2)! n=1
∞
showing the series is convergent. Then, (−1)n bn is absolutely convergent.
n=1
∞
1 2 × 5 × · · · × (3n − 1)
– If x = , we obtain , which is the series of modules of the previous series
3 n=1
3n (n + 2)!
that we have seen to be convergent. As it is of positive terms, it is absolutely convergent.
1 1 1 1
Conclusion: The series converges absolutely if x ∈ − , and diverges if x ∈ −∞, − ∪ , +∞ .
3 3 3 3
248 Series of Functions
∞ ∞
log(n) n log(n) n log(n)
e) Let us study the power series of x, x = x . Setting an = , the radius of
n=1
n n=2
n n
convergence of this series is by Corollary 1 of Theorem 3.3.1
log(n)
an
.r = lim = lim n = lim n + 1 · log(n) = 1;
log(n + 1)
an+1 n log(n + 1)
n+1
therefore, the series converges absolutely if x ∈ ] − 1, 1[ and diverges if x ∈ ] − ∞, −1[ ∪ ]1, +∞[. Regarding
the endpoints of the convergence interval, we need to study the respective numerical series.
∞
log(n)
– If x = −1, we obtain the alternating series (−1)n . Let us start by studying the series of
n=1
n
∞
log(n)
modules, , using the Integral Test.
n=1
n
log(x)
Let f (x) = , x ≥ 1:
x
(i) If x > 1, then log(x) > 0, therefore, f (x) > 0, ∀x > 1.
(ii) f is continuous, ∀x ≥ 1.
(iii) We can study the monotonicity of f by analyzing its derivative. We have
1
x· − log(x) 1 − log(x)
x
.f (x) = = ,
x2 x2
Then, the series is convergent. As the series of modules is divergent, the original series is conditionally
convergent.
∞
log(n)
– If x = 1, we obtain the series which, as we saw before, is divergent.
n=1
n
Conclusion: The power series is absolutely convergent if x ∈ ] − 1, 1[, conditionally convergent if x = −1,
and divergent if x ∈ ] − ∞, −1[ ∪ [1, +∞[.
∞ ∞
log(n) log(n)
Remark: The series and are both divergent because they only differ in a finite
n=1
n n=3
n
number of terms.
∞
3 n2 3 n2
f) Let us study the power series of x, xn . Setting an = n 3 , the radius of conver-
n=0
4n 3
(n + 2) 4 (n + 2)
gence of this series is, by Corollary 1 of Theorem 3.3.1,
3 n2
an 3n2 4n+1 (n + 1)3 + 2
= lim
n 3
4 (n + 2)
r = lim = lim
an+1 3 (n + 1)2 3 (n + 1)2 4n (n3 + 2)
. 4n+1 (n + 1)3 + 2
2
n (n + 1)3 + 2
= lim 4 = 4.
n+1 n3 + 2
3 n2
n3 + 2 3 n3
. lim = lim 3 =3
1 n +2
n
is finite and different from zero; therefore, the series is divergent by Corollary 2 of the General Com-
parison Test.
∞
∞
3 n2 3 n2
– If x = −4, we obtain the series n (n3 + 2)
(−4) n
= (−1)n 3 , which is alternating since
n=0
4 n=0
n +2
3 n2
> 0, ∀n ∈ N. The series of modules is the series we studied for the case x = 4, which we
n3 + 2
proved to be divergent. Being an alternating series, we apply Leibniz’s test:
3 n2
(i) lim 3 = 0.
n +2
3n 2
(ii) 3 > 0, ∀n ≥ 1.
n +2
3 n2
(iii) The sequence is decreasing, because
n3 + 2
Then, the series is convergent. As the series of modules is divergent, the series is conditionally conver-
gent.
Conclusion: The power series is absolutely convergent if x ∈ ] − 4, 4[, conditionally convergent if x = −4,
and divergent if x ∈ ] − ∞, −4[ ∪ [4, +∞[.
∞
1 × 4 × · · · × n2 (n!)2 (n!)2 n
g) Let an = = . By Corollary 1 of Theorem 3.3.1, the power series of x, x ,
(2 n)! (2 n)! n=1
(2 n)!
has radius of convergence
(n!)2
an
.r = lim = lim (2 n)!
= lim
(2n + 2)(2n + 1)
= 4.
a (n + 1)! 2 (n + 1)2
n+1
(2 n + 2)!
2
(n + 1)!
4n+1
bn+1 (2 n + 2)! 2n + 2
. lim = lim = lim = 1.
bn (n!)2 n 2n + 1
4
(2 n)!
Since the limit is 1, but for values greater than 1, taking into account the Note after D’Alembert’s
test, we can conclude that the series is divergent.
∞ ∞
(n!)2 (n!)2 n
– If x = −4, we obtain the alternating series (−4)n = (−1)n 4 . The series of
n=1
(2 n)! n=1
(2 n)!
∞
(n!)2 n
modules 4 is divergent. Since we establish the divergence of the series of modules by
n=1
(2 n)!
D’Alembert’s test, we can affirm that the alternating series is divergent (refer to the Note on page 105
for further details).
Conclusion: The power series converges absolutely if x ∈ ] − 4, 4[ and diverges if x ∈ ] − ∞, −4] ∪ [4, +∞[.
∞
n n
1 1
h) Let us study the power series of x, nn
sin 2
xn . Let an = nn sin 2
. The radius of
n=1
n n
convergence of this series is, by Theorem 3.3.1,
1 1 1
.r = = n = = +∞.
lim n |an | 1 1
lim n nn sin lim n sin 2
n 2 n
1 In general, this test is not useful for studying the endpoints of a power series because the limit is always 1.
3.6. Solved Exercises 251
∞
1 1
i) Let us study the power series of x, xn . Let an = . The
2n n arctan(n) − π 2n n arctan(n) − π
n=1 2 2
radius of convergence of this series is, by Corollary 1 of Theorem 3.3.1,
1
an 2n n arctan(n) − π2 2(n + 1) π2 − arctan(n + 1)
.r = lim = lim = lim = 2.
an+1
1
n π2 − arctan(n)
2n+1 (n + 1) arctan(n + 1) − π
2
Then, the series is absolutely convergent if x ∈ ] − 2, 2[ and diverges if x ∈ ] − ∞, −2[ ∪ ]2, +∞[.
At the endpoints of the convergence interval, we obtain two numerical series.
∞ ∞
1 1
– If x = 2, we have the series 2n = . Let us calculate
2n n arctan(n) − π
n arctan(n) − π2
n=1 2 n=1
the limit of its general term:
1
. lim = −1.
n arctan(n) − π2
Since the general term of the series is not a null sequence, by Theorem 2.2.1, the series is divergent.
∞
∞
(−2)n 1
– If x = −2, we obtain the series = (−1)n+1 π , which is
2 n n arctan(n) − π
n − arctan(n)
n=1 2 n=1 2
1
alternating because an = π > 0, ∀n ∈ N. The limit of the general term does not
n 2 − arctan(n)
exist, as the subsequence of even order terms has limit −1 and the subsequence of odd order terms
has limit 1. Then, the series is divergent.
Conclusion: The power series converges absolutely if x ∈ ] − 2, 2[ and diverges if x ∈ ] − ∞, −2] ∪ [2, +∞[.
∞
∞
3 × 6 × · · · × 3n n 3n n! 3n n!
j) Consider the power series x = xn . If an = , then the radius
n=1
(n + 3)! n=1
(n + 3)! (n + 3)!
of convergence of this series is, by Corollary 1 of Theorem 3.3.1,
3n n!
an (n + 3)! n+4 1
.r = lim = lim
= lim = ;
a
n+1 3n+1 (n + 1)! 3(n + 1) 3
(n + 4)!
1 1 1 1
therefore, the series converges absolutely if x ∈ − , and diverges if x ∈ −∞, − ∪ , +∞ .
3 3 3 3
Let us study the numerical series at the endpoints of the convergence interval.
∞ ∞
1 n! 1
– If x = , we obtain the series = . Let us study this series by
3 n=1
(n + 3)! n=1
(n + 3)(n + 2)(n + 1)
∞
1
comparing it with 3
which we know to be convergent (p-series with p = 3):
n=1
n
1
(n + 3)(n + 2)(n + 1) n3
. lim = lim = 1.
1 (n + 3)(n + 2)(n + 1)
n 3
Since this value is finite and different from zero, the series is convergent by Corollary 2 of the General
Comparison Test. As it is a series of positive terms, it is absolutely convergent.
252 Series of Functions
∞
1 n!
– If x = − , we obtain the alternating series (−1)n . As the series of modules is convergent,
3 n=1
(n + 3)!
the alternating series is absolutely convergent.
1 1 1 1
Conclusion: The series converges absolutely if x ∈ − , and diverges if x ∈ −∞, − ∪ , +∞ .
3 3 3 3
∞
n+1 n+1
k) Let us study the power series of x, (−1)n √ xn . Let an = (−1)n √ . The radius of
e 2n + 5 e 2n + 5
n=0
convergence of this series is, by Corollary 1 of Theorem 3.3.1,
n+1
(−1)n √ √
an e2n + 5 (n + 1) e2n+2 + 5 n+1 e2n+2 + 5
.r = lim
an+1
= lim
n + 2 = lim (n + 2)√e2n + 5 = lim n + 2 · e2n + 5
= e;
(−1) n+1
√
2n+2
e +5
then, the series is absolutely convergent if x ∈ ] − e, e[ and divergent if x ∈ ] − ∞, −e[ ∪ ]e, +∞[.
At the endpoints of the convergence interval, we obtain two numerical series that we will proceed to study.
∞
∞
n+1 (n + 1) en
– If x = −e, we obtain the series (−1)n √ (−e)n = √ . As
n=0 e2n +5 n=0 e2n + 5
(n + 1) en
. lim √ = +∞,
e2n + 5
the general term is not a null sequence; therefore, by Theorem 2.2.1, the series diverges.
∞ ∞
n+1 (n + 1) en
– If x = e, we obtain the series (−1)n √ en = (−1)n √ . As the limit of the
2n
e +5 e2n + 5
n=0 n=0
general term does not exist, the series is divergent.
Conclusion: The power series converges absolutely if x ∈ ] − e, e[ and diverges if x ∈ ] − ∞, −e] ∪ [e, +∞[.
∞
1 × 3 × 5 × · · · × (2n − 1) 1 n
l) Let us study the power series of x, · x . Setting
n=1
2 × 4 × 6 × · · · × 2n n
1 × 3 × 5 × · · · × (2n − 1) 1
.an = · ,
2 × 4 × 6 × · · · × 2n n
the radius of convergence of this series is, by Corollary 1 of Theorem 3.3.1,
1 × 3 × 5 × · · · × (2n − 1) 1
·
an × × × · · · ×
.r = lim = lim 2 4 6 2n n = lim (n + 1)(2n + 2) = 1.
1 × 3 × 5 × · · · × (2n − 1)(2n + 1) 1
an+1 · n(2n + 1)
2 × 4 × 6 × · · · × 2n(2n + 2) n+1
∞ ∞
1 × 3 × 5 × · · · × (2n − 1) 1
. · = an ,
n=1
2 × 4 × 6 × · · · × 2n n n=1
Conclusion: The power series converges absolutely if x ∈ [−1, 1] and diverges if x ∈ ] − ∞, −1[ ∪ ]1, +∞[.
∞
2n
n
15. a) Let us study the power series of x − 1, (x − 1)n . Let y = x − 1, and consider the
n=1
2n + 1
∞
2n 2n
n n
series y n . Setting an = , the radius of convergence of this series is, by
n=1
2n +1 2n + 1
Theorem 3.3.1,
1 1 1 1
.r = = 2n = 2 = = 4.
lim n |an | n n n2
lim n lim lim
2n + 1 2n + 1 4n2 + 4n + 1
As 2n
2n 1
. lim = ,
2n + 1 e
the general term does not have limit; therefore, by Theorem 2.2.1, the series is divergent.
– If x = 5, we obtain the series
∞
2n ∞
2n
n 2n
. 4n = ,
n=1
2n + 1 n=1
2n + 1
whose general term does not tend to zero, as we saw earlier; therefore, the series is divergent.
254 Series of Functions
In summary, the power series of x − 1 is divergent if x ∈ ] − ∞, −3] ∪ [5, +∞[ and absolutely convergent if
x ∈ ] − 3, 5[.
∞
(x − 2)n
b) Let us study the power series of x − 2, (−1)n . Let y = x − 2, and consider the series
n=2
n log(n)
∞
n yn (−1)n
(−1) . Setting an = , the radius of convergence of this series is, by Corollary 1 of
n=2
n log(n) n log(n)
Theorem 3.3.1,
(−1)n
an
n log(n) n + 1 log(n + 1)
.r = lim = lim = lim · = 1;
an+1 (−1)n+1 n log(n)
(n + 1) log(n + 1)
therefore, the series converges absolutely if y ∈ ]−1, 1[ and diverges if y ∈ ]−∞, −1[ ∪ ]1, +∞[. As y = x−2,
∞
(x − 2)n
the series (−1)n is absolutely convergent if x ∈ ]1, 3[ and divergent if x ∈ ] − ∞, 1[ ∪ ]3, +∞[.
n=2
n log(n)
Let us study the numerical series obtained at the endpoints of the convergence interval.
∞
∞
(−1)n 1
– If x = 1, we obtain the series (−1)n = . Let us study it using the Integral
n=2
n log(n) n=2
n log(n)
Test.
1
Let f (x) = :
x log(x)
(i) As x ≥ 2 ⇒ log(x) > 0, we have f (x) > 0, ∀x ≥ 2.
(ii) f is continuous on [2, +∞[.
(iii) f is decreasing on [2, +∞[ because x log(x) is increasing and positive.
∞ +∞
1 1
Then the numerical series and the improper integral dx are both conver-
n=2
n log(n) 2 x log(x)
gent or both divergent. Since
+∞ x x 1
1 1 t
dx = lim dt = lim dt
. 2 x log(x) x→+∞ 2 t log(t) x→+∞ 2 log(t)
x
= lim log log(t) = lim log log(x) − log log(2) = +∞,
x→+∞ 2 x→+∞
Then, we conclude that the series is convergent. As the series of modules is divergent, the series is
conditionally convergent.
Conclusion: The power series of x − 2 is absolutely convergent if x ∈ ]1, 3[, conditionally convergent if
x = 3, and divergent if x ∈ ] − ∞, 1] ∪ ]3, +∞[.
∞
n
c) Let us study the power series of x + 4, n (2n2 − 1)
(x + 4)n . We can rewrite it in the form
n=1
(−3)
∞ ∞
n n
(−1)n n 2 − 1)
(x + 4) n
. Let y = x + 4, and consider the series (−1)n n 2 − 1)
y n . If
n=1
3 (2n n=1
3 (2n
n
an = (−1)n n , the radius of convergence of this series is, by Corollary 1 of Theorem 3.3.1,
3 (2n2 − 1)
n
(−1)n n
an 2 − 1) 2
.r = lim = lim 3 (2n = lim 3n(2n + 4n + 1) = 3.
n+1 (n + 1)(2n2 − 1)
an+1 (−1)n+1
3n+1 2(n + 1)2 − 1
At the endpoints of the convergence interval, we have to study the respective numerical series.
∞
∞
n n
– If x = −7, we obtain the series (−1)n (−3)n = .
n=1
−
3n (2n2 1) n=1
2n 2−1
It is of positive terms, and we can compare it with the harmonic series, which is divergent. Since the
limit
n
2n2 − 1 n2 1
. lim = lim =
1 2n2 − 1 2
n
is finite and different from zero, the series is divergent by Corollary 2 of the General Comparison Test.
∞ ∞
n n
– If x = −1, we obtain the alternating series (−1)n n 3n = (−1)n 2 . We apply
n=1
3 (2n − 1)
2
n=1
2n − 1
Leibniz’s test since the series of modules is divergent:
n
(i) lim = 0.
2n2 − 1
n
(ii) > 0, ∀n ≥ 1.
2n2 − 1
n+1 n 2n2 + 2n + 1
(iii) − = − < 0, ∀n ∈ N, which implies that
2(n + 1) − 1 2n2 − 1 (2n2 + 4n + 1)(2n2 − 1)
2
n
is decreasing.
2n2 − 1
We can conclude that the series is convergent. Since the series of modules is divergent, the series is
conditionally convergent.
∞
en
d) Let us study the power series of x − e, (x − e)n . Let y = x − e, and consider the series
n=2
log(n)
∞
en en
y n . If an = , the radius of convergence of this series is, by Corollary 1 of Theorem 3.3.1,
n=2
log(n) log(n)
en
an log(n) 1 log(n + 1) 1
.r = lim = lim
= lim · = ,
a
n+1 en+1 e log(n) e
log(n + 1)
1 1 1 1
so the series is absolutely convergent if y ∈ − , and divergent if y ∈ −∞, − ∪ , +∞ . As
e e e e
∞
en 1 1
y = x − e, the series (x − e)n is absolutely convergent if x ∈ e − , e + and divergent if
log(n) e e
n=2
1 1
x ∈ −∞, e − ∪ e + , +∞ .
e e
Let us study the numerical series corresponding to the endpoints of the interval.
∞ ∞
1 en 1 n (−1)n
– If x = e − , we get the series − = . It is an alternating series because
e n=2
log(n) e n=2
log(n)
∞
1 1
an = > 0, ∀n > 1. Let us consider the series of modules, , and study it by
log(n) n=2
log(n)
comparing it with the harmonic series. As the limit
1
log(n) n
. lim = lim = +∞
1 log(n)
n
∞
1
by Corollary 4 of the General Comparison Test, the series is divergent. Let us apply Leibniz’s
n=2
log(n)
test:
1
(i) lim = 0.
log(n)
1
(ii) > 0, ∀n ≥ 2.
log(n)
1
(iii) The sequence log(n) is increasing; therefore, is a decreasing sequence for n ≥ 2.
log(n)
Then, the series is convergent. Since the series of modules is divergent, the alternating series is
conditionally convergent.
∞
n ∞
1 en 1 1
– If x = e + , we get the series = which, as we have seen, is divergent.
e n=2
log(n) e n=2
log(n)
1 1
Conclusion: The power series of x − e is absolutely convergent if x ∈ e − , e + , conditionally conver-
e e
1 1 1
gent if x = e − and divergent if x ∈ −∞, e − ∪ e + , +∞ .
e e e
3.6. Solved Exercises 257
∞ π
1 cos 2n
and let y = . We study the power series of y, y n . Note that if n > 1, then
x−1 (2n − 1)2
n=2
π π π π
cos 2n
0< < ; therefore, cos > 0. Let an = . The radius of convergence of this series
2n 2 2n (2n − 1)2
is, by Corollary 1 of Theorem 3.3.1,
cos π
2n π
an − 1)2 cos 2n (2n + 1)2
.r = lim = lim (2n
= lim = 1.
an+1 cos
π π
(2n − 1)2
cos 2(n+1)
2(n+1)
(2n + 1)2
Therefore, the series converges absolutely if y ∈ ] − 1, 1[ and diverges if y ∈ ] − ∞, −1[ ∪ ]1, +∞[. If
y = ±1, we obtain two numerical series we will study.
∞ π
cos 2n
– If y = 1, we obtain the series , which is of positive terms. We are going to compare it
n=2
(2n − 1)2
∞
1
with 2
, which we know to be convergent, as it is a p-series with p = 2. The limit
n=1
n
π
cos 2n π
(2n − 1)2 n2 cos 2n 1
. lim = lim =
1 (2n − 1)2 4
n2
is finite and different from zero; therefore, by Corollary 2 of the General Comparison Test, the series is
convergent. As it is of positive terms, it is absolutely convergent.
∞ π
cos 2n
– If y = −1, we obtain the alternating series (−1)n , whose series of modules is the one
n=2
(2n − 1)2
we studied for the case y = 1, which we proved to be convergent. Therefore, the series is absolutely
convergent.
Conclusion: The power series of y is divergent if y ∈ ] − ∞, −1[ ∪ ]1, +∞[ and absolutely convergent if
y ∈ [−1, 1].
1
Concerning the original series, we have, because y = ,
x−1
1
.|y| ≤ 1 ⇔ ≤ 1 ⇔ |x − 1| ≥ 1 ⇔ x ≥ 2 ∨ x ≤ 0,
x − 1
which implies that the series converges absolutely if x ∈ ] − ∞, 0] ∪ [2, +∞[ and diverges if x ∈ ]0, 2[ \{1},
1
because 1 does not belong to the domain of .
x−1
258 Series of Functions
∞
(2x2 − 5)n
f) Let us study the power series of 2x2 − 5, . Let y = 2x2 − 5 and consider the series
n=1
n 3n+1
∞
1
yn .
n=1
n 3n+1
1
Setting an = , the radius of convergence of this series is, by Corollary 1 of Theorem 3.3.1,
n 3n+1
1
an
.r = lim
= lim n 3n+1
= lim 3(n + 1) = 3;
1
an+1 n
(n + 1) 3n+2
therefore, the series converges absolutely if y ∈ ] − 3, 3[ and diverges if y ∈ ] − ∞, −3[ ∪ ]3, +∞[.
At the endpoints of the interval of convergence, it is necessary to study the respective numerical series.
∞ ∞ ∞
1 (−1)n 1 (−1)n
– If y = −3, we obtain the series (−3) n
= = , which is the
n=1
n 3n+1 n=1
3n 3 n=1 n
alternating harmonic series that we know to be conditionally convergent.
∞ ∞
1 1 1
– If y = 3, we obtain the series = , which is divergent.
n=1
3n 3 n=1 n
But
|y| < 3 ⇔ |2x2 − 5| < 3 ⇔ −3 < 2x2 − 5 < 3 ⇔ 1 < x2 < 4
.
⇔ x ∈ ] − 2, 2[ ∩ ] − ∞, −1[ ∪ ]1, +∞[ ⇔ x ∈ ] − 2, −1[ ∪ ]1, 2[.
In addition,
.2x
2
− 5 = 3 ⇔ x = −2 ∨ x = 2
and
.2x
2
− 5 = −3 ⇔ x = −1 ∨ x = 1.
Conclusion: The original series is absolutely convergent if x ∈ ] − 2, −1[ ∪ ]1, 2[, conditionally convergent
if x = −1 or x = 1 and divergent if x ∈ ] − ∞, −2] ∪ ] − 1, 1[ ∪ [2, +∞[.
√ ∞ √
n+1− n
g) Let us study the power series of x + 2, √ √ (x + 2)n . We can rewrite the series in the form
n=1
n+1+ n
∞ ∞
1 1
(x + 2)n . Let y = x + 2 and consider the series yn .
n=1 2n + 1 + 2 n(n + 1) n=1 2n + 1 + 2 n(n + 1)
1
Let an = . The radius of convergence of this series is, by Corollary 1 of Theorem 3.3.1,
2n + 1 + 2 n(n + 1)
1
an 2n + 1 + 2 n(n + 1) 2n + 3 + 2 (n + 1)(n + 2)
.r = lim = lim
= lim = 1;
a 1 2n + 1 + 2 n(n + 1)
n+1
2n + 3 + 2 (n + 1)(n + 2)
therefore, the series absolutely converges if y ∈ ] − 1, 1[ and diverges if y ∈ ] − ∞, −1[ ∪ ]1, +∞[. As
∞
1
y = x + 2, we can conclude that the series (x + 2)n is absolutely convergent if
n=1 2n + 1 + 2 n(n + 1)
x ∈ ] − 3, −1[ and divergent if x ∈ ] − ∞, −3[ ∪ ] − 1, +∞[.
Let us study the numerical series at the endpoints of the interval of convergence.
3.6. Solved Exercises 259
∞
1
– If x = −1, we obtain the series . As it is of positive terms and the limit
n=1 2n + 1 + 2 n(n + 1)
1
2n + 1 + 2 n(n + 1) n 1
. lim = lim =
1 2n + 1 + 2 n(n + 1) 4
n
is finite and different from zero, the series is divergent because the harmonic series is divergent (see
Corollary 2 of the General Comparison Test).
∞
(−1)n
– If x = −3, we obtain the series . It is an alternating series, and we apply
n=1 2n + 1 + 2 n(n + 1)
Leibniz’s test since the series of modules is divergent:
1
(i) lim = 0.
2n + 1 + 2 n(n + 1)
1
(ii) > 0, ∀n ≥ 1.
2n + 1 + 2 n(n + 1)
1 1
(iii) − < 0, which implies that the sequence is
2n + 3 + 2 (n + 1)(n + 2) 2n + 1 + 2 n(n + 1)
decreasing.
Then, the alternating series is convergent. It is conditionally convergent because the series of modules
is divergent.
∞ n
log2 (n) 2x 2x
h) Let us consider the series 2
· 2
and set y = 2 . We are going to study the power
n=1
n x + 1 x +1
∞
log2 (n) n
series of y, y . The radius of convergence of this series is, by Corollary 1 of Theorem 3.3.1,
n=1
n2
log2 (n)
with an = ,
n2
log2 (n)
an (n + 1)2 log2 (n)
n2
.r = lim = lim
= lim 2 = 1.
a log2 (n + 1) n log2 (n + 1)
n+1
(n + 1)2
∞ ∞
log2 (n) log2 (n)
– If x = 1, we obtain the series 2
= , which is of positive terms, and we will study
n=1
n n=2
n2
∞
1
it by comparing it with the convergent series . The limit
n=1
n3/2
log2 (n)
n2 log2 (n)
. lim = lim √ =0
1 n
n3/2
allows us to conclude, by Corollary 3 of the General Comparison Test, that the series is convergent. As
it is of positive terms, it is absolutely convergent.
∞
log2 (n)
– If x = −1, we obtain the series (−1)n , which is alternating. As we have seen, the series of
n=2
n2
modules is convergent, and the alternating series is absolutely convergent.
∞
1 1 1
i) Let us study the power series of , √ · . It is important to note that the set where the
x n=1 2n n + 1 xn
series converges does not include the point x = 0 since the functions are not defined at that point. Rather,
∞
1 1 1
we can set y = , and consider the series n
√ y n . By setting an = n √ , we can apply
x n=1
2 n + 1 2 n+1
Corollary 1 of Theorem 3.3.1 to determine the radius of convergence of this series:
1
√ √
an 2n n+1 2 n+2
.r = lim = lim
= lim √ = 2.
a 1 n+1
n+1
n+1
√
2 n+2
1
The series converges absolutely if y ∈ ] − 2, 2[ and diverges if y ∈ ] − ∞, −2[ ∪ ]2, +∞[. Since y = ,
x
∞
1 1 1
we conclude that the series √ · n converges absolutely if ∈ ] − 2, 2[ and diverges if
2n n + 1 x x
n=1
1 1 1
∈ ] − ∞, −2[ ∪ ]2, +∞[, that is, it is absolutely convergent if x ∈ −∞, − ∪ , +∞ and divergent
x 2 2
1 1
if x ∈ − , \ {0}.
2 2
1 1
Let us study the numerical series obtained at x = − and x = .
2 2
∞
∞
1 1 1
– If x = , we obtain the divergent series √ = √ .
2 n=1
n + 1 n=2
n
∞
1 1
– If x = − , we obtain the alternating series (−1)n √ . We will apply Leibniz’s test since the
2 n=1
n+1
series of modules is divergent:
1
(i) lim √ = 0.
n+1
1
(ii) √ > 0, ∀n ≥ 1.
n+1
3.6. Solved Exercises 261
√ √
1 1 n− n+1 1
(iii) √ −√ = √ √ < 0, ∀n ∈ N, which implies that √ is decreasing.
n+1 n n n+1 n
It follows that the series is convergent. As the series of the modules is divergent, the series is condi-
tionally convergent.
1 1
Conclusion: The power series is absolutely convergent if x ∈ −∞, − ∪ , +∞ , conditionally
2 2
1 1 1
convergent if x = − , and divergent if x ∈ − , \ {0}.
2 2 2
∞
n
2 1
j) Let us study the power series + (2x − 3)n . Let y = 2x − 3, and consider the series
n=1
3 2n
∞
n n
2 1 2 1
+ y n . If an = + , the radius of convergence of this series is, by Theorem 3.3.1,
n=1
3 2n 3 2n
1 1 1 3
.r = = n = = .
lim n |an | 2 1 2 1 2
lim n + lim +
3 2n 3 2n
3 3 3 3
The series converges absolutely if y ∈ − , and diverges if y ∈ −∞, − ∪ , +∞ . However,
2 2 2 2
∞
2 1 n n 3 9
y = 2x − 3, so the series + (2x − 3) is absolutely convergent if x ∈ , and divergent
3 2n 4 4
n=1
3 9
if x ∈ −∞, ∪ , +∞ .
4 4
Regarding the endpoints of the interval of convergence, we have to study the respective numerical series.
∞ ∞
9 2 1 n 3 n 3 n
– If x = , we obtain the series + = 1+ . As
4 n=1
3 2n 2 n=1
4n
1
3 n 3 4n 4 3
. lim 1 + = lim 1+ = e4 ,
4n 4n
the general term is not a null sequence; therefore, by Theorem 2.2.1, the series diverges.
∞ ∞
3 2 1 n 3 n 3 n
– If x = , we obtain the series + − = (−1)n 1 + . As the limit of the
4 n=1
3 2n 2 n=1
4n
general term does not exist, the series is divergent.
3 9
Conclusion: The power series of x is divergent if x ∈ −∞, ∪ , +∞ and absolutely convergent if
4 4
3 9
x∈ , .
4 4
∞ 2
23n (n − 1)!
k) Let us study the power series of x2 , x2n . Let y = x2 . The radius of convergence of
n=2
(2n)!
262 Series of Functions
∞ 2 2
23n (n − 1)! 23n (n − 1)!
the series y n is, by Corollary 1 of Theorem 3.3.1, and setting an = ,
n=2
(2n)! (2n)!
23n (n − 1)!2
an (2n)! (2n + 2)(2n + 1) 1
.r = lim = lim = lim = .
an+1 23n+3 (n!)2
23 n2 2
(2n + 2)!
1 1 1 1
The series is absolutely convergent if y ∈ − , and diverges if y ∈ −∞, − ∪ , +∞ . Since
2 2 2 2
∞ 2
23n (n − 1)! 1 1
2
y = x , the series x is absolutely convergent if x ∈ − √ , √
2n
and divergent if
(2n)! 2 2
n=2
1 1
x ∈ −∞, − √ ∪ √ , +∞ .
2 2
At the endpoints of the interval of convergence, we obtain two numerical series.
∞ 2 ∞
1 23n (n − 1)! 1
– If x = − √ , we obtain the series · n = bn . Using Raabe’s test,
2 n=1
(2n)! 2 n=1
⎛ 2n 2 ⎞
2 (n − 1)!
⎜ ⎟
bn ⎜ (2n)! ⎟ (2n + 2)(2n + 1) 3
. lim n − 1 = lim n ⎜ − 1⎟ = lim n − 1 = > 1,
bn+1 ⎝ 22n+2 (n!)2 ⎠ 4n2 2
(2n + 2)!
2n
6n − 3 5
. lim = e− 3 ,
6n + 2
the limit of the general term does not exist, and, by Theorem 2.2.1, the series diverges.
∞ ∞
21 2n − 1 2n 9 n 6n − 3 2n
– If x = , we obtain the series = . The general term is not
4 n=0
3n + 1 4 n=1
6n + 2
a null sequence, so the series is divergent.
3 21
Conclusion: The power series of x − 3 is divergent if x ∈ −∞, ∪ , +∞ and absolutely convergent
4 4
3 21
if x ∈ , .
4 4
∞ π ∞ π
sin n sin n
2n
m) Let us study the power series of x−1, (x−1) = (x−1)2n . Let y = (x−1)2 ,
4n (n + 1) 4n (n + 1)
∞
n=1 n=2
sin n π
and consider the series y n . The radius of convergence of this series is, by Corollary 1 of
4n (n + 1)
n=2 π
sin n
Theorem 3.3.1 and setting an = n ,
4 (n + 1)
sin π
n
n π n+1
an 4 (n + 1) sin n 4 (n + 2)
.r = lim = lim
π = lim π = 4;
a sin sin 4n (n + 1)
n+1 n+1 n+1
n+1
4 (n + 2)
Since this limit is finite and different from zero, the series is convergent by Corollary 2 of the General
Comparison Test. Being of positive terms, it is absolutely convergent.
264 Series of Functions
∞ π ∞ π
sin sin
– If x = 3, we again obtain the series n
22n = n
, which we have seen to be
n=2
4n (n + 1) n=2
n+1
absolutely convergent.
Conclusion: The power series of x − 1 is divergent if x ∈ ] − ∞, −1[ ∪ ]3, +∞[ and absolutely convergent if
x ∈ [−1, 3].
∞
x2n x2
16. The series √ log 1 + √ is of positive terms. For it to be convergent, its general term must be a null
n=1
n n
sequence (see Theorem 2.2.1). But
x2n x2 0, if |x2 | ≤ 1
. lim √ log 1 + √ =
n n +∞, if |x2 | > 1;
we conclude, by Corollary 2 of the General Comparison Test, that the series is divergent. The original series
converges if x ∈ ] − 1, 1[ and diverges if x ∈ ] − ∞, −1] ∪ [1, ∞[.
∞
π n π n
17. Let y = x2 . The series cos + 2 − (−1)n y n setting an = cos n + 2 − (−1)n has radius of
n=1
n
convergence
1 1 1 1
. = ! π = π = .
lim n |an | n
lim cos + 2 − (−1) n 4
lim n cos + 2 − (−1)n n
n
1 1 1 1
The series is absolutely convergent if y ∈ − , and diverges if y ∈
∪ , +∞ .−∞, −
4 4 4 4
∞
π
n 1 1 1 1
The series cos + 2 − (−1)n (x2 )n is absolutely convergent if x2 ∈ − , , that is, x ∈ − , ,
n=1
n 4 4 2 2
and this is the largest open interval on which the series converges.
∞
4n 5
18. Consider the power series of x, n + 5n
xn . If the radius of convergence of the series is , then, by Theo-
2 4
n=1
5 5 5 5
rem 3.3.1, the series is absolutely convergent on the interval − , and diverges on −∞, − ∪ , +∞ .
4 4 4 4
5 5
But 1 ∈ − , , therefore, the series is absolutely convergent at this point. The series is divergent at x = 2
4 4
5 5
because 2 ∈ −∞, − ∪ , +∞ .
4 4
5
At x = , the series may be convergent or divergent as it is one of the endpoints of the interval of convergence.
4 n
∞ ∞
4n 5 5n
Let us study the series n + 5n
= n + 5n
. Its general term is not a null sequence. In fact,
n=1
2 4 n=1
2
5n 1
. lim = lim n = 1;
2n + 5n 2
+1
5
∞
19. The series an xn is a power series of x. We have
n=0
√ 1, if n is even
.n an = √n
2, if n is odd,
√ √
n
which implies that lim n an = 1, because lim 2 = 1. Then the radius of convergence of this series is, by
Theorem 3.3.1,
1
.r = √ = 1;
lim n an
therefore, the series is absolutely convergent if x ∈ ] − 1, 1[ and divergent if x ∈ ] − ∞, −1[ ∪ ]1, +∞[.
266 Series of Functions
∞
If x = −1, we obtain the series (−1)n an , whose general term does not have a limit; therefore, by Theo-
n=1
rem 2.2.1, the series is divergent.
∞
If x = 1, we obtain the series an . As lim an does not exist, the series is divergent.
n=1
Conclusion: The power series is absolutely convergent if x ∈ ] − 1, 1[ and diverges if x ∈ ] − ∞, −1] ∪ [1, +∞[.
∞
∞
20. Let us consider the power series an xn . If, by hypothesis, the series an 5n converges, then either the power
n=1 n=1
series converges absolutely on R or there exists r ∈ R+ such that the series converges absolutely if x ∈ ] − r, r[
∞
and diverges if x ∈ ] − ∞, −r[ ∪ ]r, +∞[. In the first case, it is evident that the series an (−2)n is absolutely
n=1
convergent. In the second case, if the series converges at x = 5, then r ≥ 5. As −2 ∈ ] − 5, 5[ ⊂ ] − r, r[ , the
∞
series an (−2)n is absolutely convergent.
n=1
21. Suppose the radius of convergence of the power series of x − 3 is r > 0. In that case, the series is absolutely
convergent on the interval ]3 − r, 3 + r[, which is the largest open interval where the series is convergent. The
∞
series may also be convergent at the endpoints of the interval. If the series an (x − 3)n converges at x = 0,
n=0
then 0 ∈ [3 − r, 3 + r], which implies that r ≥ 3. But in that case, 1 ∈ ]3 − r, 3 + r[, so the series cannot diverge
at x = 1.
If the radius of convergence of the series is +∞, then the series converges absolutely at all points of R, with no
points where it can be divergent.
∞
22. If the power series of x, bn xn , has a radius of convergence R > 1, then it is absolutely convergent if x = 1,
n=0
∞
that is, the series bn is absolutely convergent. If, for each n ∈ N, |fn (x)| ≤ |bn |, ∀x ∈ R, Weierstrass’ test
n=0
∞
allows us to conclude that the series of functions fn (x) is uniformly convergent on R.
n=0
∞
23. a) Let us consider a power series an (x − x0 )n with a radius of convergence r > 0. By Theorem 3.3.1, the
n=1
series is absolutely convergent if x ∈ ]x0 − r, x0 + r[ and divergent if x ∈ ] − ∞, x0 − r[ ∪ ]x0 + r, +∞[,
which implies that it can only be conditionally convergent at the endpoints of the interval of convergence,
that is, at x = x0 − r or x = x0 + r.
∞
If the series an (x − 1)n is conditionally convergent at x = −1, then x0 − r = −1. Since x0 = 1, we
n=1
have r = 2.
b) From the previous item, we know that the interval of convergence of the series is ] − 1, 3[, so the series
∞
can only be conditionally convergent at x = −1 or x = 3. If x = 3, we get the series an 2n . As,
n=1
by hypothesis, (an ) is a sequence of positive terms, this is a series of positive terms that, if convergent,
∞
will be absolutely convergent. Therefore, the only point where the series an (x − 1)n is conditionally
n=1
convergent is x = −1.
3.6. Solved Exercises 267
∞
∞
1 n 1
24. a) Let us study the power series of x−2, n
(x−2) . Let y = x−2, and consider the series yn .
n=1
n 5 n=1
n 5n
1
The radius of convergence of this series is, by Corollary 1 of Theorem 3.3.1 and setting an = ,
n 5n
1
an n 5n (n + 1) 5n+1
. lim = lim
= lim = 5;
a 1 n 5n
n+1
(n + 1) 5n+1
the series is absolutely convergent if y ∈ ] − 5, 5[ and diverges if y ∈ ] − ∞, −5[ ∪ ]5, +∞[. As y = x − 2, the
∞
1
series (x − 2)n is absolutely convergent if x ∈ ] − 3, 7[ and divergent if x ∈ ] − ∞, −3[ ∪ ]7, +∞[.
n=1
n 5n
∞
∞
1 1
If x = 7, we get the series n
5n = , which is divergent.
n=1
n5 n=1
n
∞
∞
1 (−1)n
If x = −3, we get the series n
(−5)n = , which is conditionally convergent.
n=1
n5 n=1
n
∞
∞ ∞
1 1 1
f (x) = (x − 2)n = (x − 2) n
= n (x − 2)n−1
n=1
n 5n n=1
n 5 n
n=1
n 5 n
.
∞ ∞ ∞
1 1 x−2 n 1 x−2 n
= (x − 2) n−1
= = , ∀x ∈ ] − 3, 7[.
n=1
5n n=0
5 5 5 n=0 5
x−2
This series is geometric with r = . Then
5
∞
x−2 n 1 5
. = =
5 x−2 7 − x
n=0 1−
5
and, finally,
1 5 1
.f (x) = · = , ∀x ∈ ] − 3, 7[.
5 7−x 7−x
∞
(−1)n n+1
25. Let us consider the series log(x) .
n=0
n + 1
268 Series of Functions
∞
(−1)n n+1
a) Let y = log(x), and let us study the power series y . The radius of convergence of this series
n=0
n+1
(−1) n
is, by Corollary 1 of Theorem 3.3.1 and setting an = ,
n+1
(−1)n
an n+1 n+2
. lim = lim = lim = 1;
an+1 (−1)n+1
n+1
n+2
so the series is absolutely convergent if y ∈ ] − 1, 1[ and diverges if y ∈ ] − ∞, −1[ ∪ ]1, +∞[. Since
∞
(−1)n n+1
y = log(x), we conclude that the series log(x) is absolutely convergent if log(x) ∈ ] − 1, 1[
n=0
n + 1
and divergent if log(x) ∈ ] − ∞, −1[ ∪ ]1, +∞[, that is, it is absolutely convergent if x ∈ ]e−1 , e[ and
divergent if x ∈ ]0, e−1 [ ∪ ]e, +∞[, because the domain of log(x) is R+ .
∞ ∞ ∞ ∞
(−1)n (−1)2n+1 1 1
– If x = e−1 , we obtain the series (−1)n+1 = = − = − ,
n=0
n+1 n=0
n+1 n=0
n+1 n=1
n
which is divergent.
∞ ∞ ∞
(−1)n (−1)n+1 (−1)n
– If x = e, we obtain the series =− =− , which is the alternating
n=0
n + 1 n=0
n + 1 n=1
n
harmonic series, therefore conditionally convergent.
Conclusion: The original series is absolutely convergent if x ∈ ]e−1 , e[, conditionally convergent if x = e
and divergent if x ∈ ]0, e−1 ] ∪ ]e, +∞[.
∞
(−1)n n+1
b) The function log(x) is the sum of a power series with interval of convergence ]e−1 , e[, so
n=0
n + 1
it is differentiable term by term on this interval (see Theorem 3.3.3), that is, if x ∈ ]e−1 , e[, we have
∞ ∞ ∞
(−1)n n+1 (−1)n n+1 1 n
. log(x) = log(x) = (−1)n log(x) .
n=0
n + 1 n=0
n + 1 x n=0
∞
n
But the series (−1)n log(x) is geometric with ratio r = − log(x); therefore,
n=0
∞
1 n 1 1
. (−1)n log(x) = · ,
x n=0 x 1 + log(x)
so
∞
(−1)n n+1 1 1
. log(x) = · .
n=0
n + 1 x 1 + log(x)
Then
∞
(−1)n n+1 1 1
. log(x) = · dx = log 1 + log(x) + C.
n=0
n + 1 x 1 + log(x)
Let us determine the value of the constant C. For x = 1 (note that this value belongs to the interval of
convergence), we obtain
∞
(−1)n n+1
.0 = log(1) = log 1 + log(1) + C ⇔ C = 0;
n=0
n + 1
3.6. Solved Exercises 269
thus,
∞
(−1)n n+1
. log(x) = log 1 + log(x) .
n=0
n + 1
26. a) The domain of f is the subset of R where the series is convergent. Let us consider the power series of x2 ,
∞ ∞
(−1)n (−1)n
2n 2
x2n . Let y = x2 . The radius of convergence of the series 2n 2
y n is, by Corollary 1
n=0
2 (n!) n=0
2 (n!)
(−1)n
of Theorem 3.3.1 and setting an = 2n ,
2 (n!)2
(−1)n
an 22n (n!)2
. lim = lim
= lim 22 (n + 1)2 = +∞.
a (−1)n+1
n+1
22n+2 (n + 1)! 2
∞
(−1)n
Then, the series converges absolutely for all y ∈ R. As y = x2 , the series x2n is absolutely
n=0
22n(n!)2
convergent for all x ∈ R; therefore the domain of f is R.
b) As the power series is differentiable term by term on its interval of convergence (see Theorem 3.3.3), then,
for all x ∈ R,
∞ ∞
(−1)n (−1)n 2n 2n−1
2n
.f (x) = x = x
2n
2 (n!) 2 22n (n!)2
n=0 n=1
and
∞
∞
(−1)n 2n 2n−1 (−1)n 2n(2n − 1) 2n−2
.f (x) = x = x .
n=1
22n (n!)2 n=1
22n (n!)2
Then
∞ ∞
2 (−1)n (−1)n
.x f (x) = x2 2n (n!)2
x 2n
= 2n (n!)2
x2n+2
n=0
2 n=0
2
∞ ∞
(−1)n 2n 2n−1 (−1)n 2n 2n
.x f (x) = x x = x
22n (n!)2 22n (n!)2
n=1 n=1
and
∞ ∞
(−1)n 2n(2n − 1) 2n−2 (−1)n 2n(2n − 1) 2n
.x
2
f (x) = x2 2n 2
x = x .
n=1
2 (n!) n=1
22n (n!)2
270 Series of Functions
Therefore,
∞ ∞ ∞
(−1)n 2n(2n − 1) 2n (−1)n 2n 2n (−1)n
= x + x + x2n+2
n=1
22n (n!)2 n=1
22n (n!)2 n=0
22n (n!)2
∞ ∞
(−1)n 2n(2n − 1) + 2n (−1)n
= x2n + x2n+2
n=1
22n (n!)2 n=0
2 (n!)2
2n
∞ ∞
.
(−1)n 4n2 2n (−1)n−1 2n
= 2n 2
x + 2 x
n=1
2 (n!) n=1 22n−2 (n − 1)!
∞ ∞
(−1)n 4n2 2n (−1)n 4
= 2n (n!)2
x − 2 x
2n
n=1
2 n=1 2
2n (n − 1)!
∞
∞
(−1)n 4 (−1)n 4
2 x −
2n 2n
= x = 0.
22n (n − 1)! 2n (n − 1)! 2
n=1 n=1 2
∞
27. Let (an ) be a sequence of real numbers such that the series an 2n is conditionally convergent.
n=1
∞
∞
a) If the series an 2n is conditionally convergent, then the radius of convergence of the series an xn
n=1 n=1
is 2, which implies that the power series of x is absolutely convergent if x ∈ ] − 2, 2[ and divergent if
∞
x ∈ ] − ∞, −2[ ∪ ]2, +∞[. It follows that the series an is absolutely convergent (x = 1) and the series
n=1
∞
n
an 3 is divergent (x = 3).
n=1
∞
sin(an )
b) The series an is convergent; therefore, by Theorem 2.2.1, lim an = 0. Then lim = 1.
n=1
an
∞
28. Let f (x) = an xn .
n=0
a) We know, by Theorem 3.4.3, that the power series of x is unique and is equal to the Maclaurin series of f ,
∞
f (n) (0) n f (n) (0)
that is, f (x) = x , ∀n ∈ N0 . Therefore, an = . We will prove, by induction, that
n=0
n! n!
.f
(n)
(x) = f (x), ∀n ∈ N, ∀x ∈ R.
f (n) (0) 1
Then f (n) (0) = f (0) = f (0) = 1 and an = = , ∀n ∈ N.
n! n!
3.6. Solved Exercises 271
∞
xn
b) Taking into account part a), we have f (x) = , ∀x ∈ R. However, this series is the Maclaurin series
n=0
n!
of ex , that is, f (x) = ex .
5x − 1
29. a) Let f (x) = . As x2 − x − 2 = (x + 1)(x − 2), we have
x2 − x − 2
5x − 1 A B A(x + 1) + B(x − 2) (A + B)x + A − 2B
.f (x) = = + = = .
x2 − x − 2 x−2 x+1 (x − 2)(x + 1) (x − 2)(x + 1)
Then
A+B =5 A=3
. ⇔
A − 2B = −1 B = 2;
therefore,
3 2 3 2 3 1 2
.f (x) = + =− + =− · + .
x−2 x+1 2−x 1 − (−x) 2 1− x 1 − (−x)
2
1 1
Since is the sum of a geometric series with ratio r = −x and the first term equal to 1 and x
1 − (−x) 1−
2
x
is the sum of a geometric series with ratio r = and the first term equal to 1, we have
2
∞ ∞
1
. = (−x)n = (−1)n xn
1 − (−x) n=0 n=0
and
∞
x n 1
x = 2
..
1− n=0
2
x
These equalities are valid if | − x| < 1 and < 1, respectively, that is, |x| < 1 and |x| < 2. We can write
2
the Maclaurin series expansion of f :
∞
3 x n
∞ ∞ 3
.f (x) = − +2 (−1)n xn = 2 (−1)n − n+1 xn ,
2 n=0 2 n=0 n=0
2
∞ ∞ 2n
t2 (t2 )n t
.e = = , ∀t ∈ R.
n=0
n! n=0
n!
As it is a power series, by Corollary 1 of Theorem 3.3.2, it is integrable term by term on every closed bounded
interval of R. Therefore,
x x ∞ 2n ∞ x 2n ∞ x
2 t t 1
f (x) = et dt = dt = dt = t2n dt
0 0 n=0 n! n=0 0
n! n=0
n! 0
.
∞ x ∞ ∞
1 t2n+1 1 x2n+1 x2n+1
= = · = ,
n=0
n! 2n + 1 0 n=0
n! 2n + 1 n=0
n!(2n + 1)
∞ ∞
−tx (−tx)n (−1)n tn xn
.e = = .
n=0
n! n=0
n!
Then
∞ ∞
(−1)n tn xn (−1)n tn xn
1− −
1 1 − e−tx 1
n=0
n! 1
n=1
n!
f (x) = dt = dt = dt
. 0 t 0 t 0 t
∞
1 ∞
1
(−1)n tn−1 xn (−1)n+1 xn n−1
= − dt = t dt.
0 n=1 n! 0 n=1 n!
For each x ∈ R \ {0}, the power series of t has an infinite radius of convergence:
(−1)n xn |x|n
n! n+1
. lim = lim n! = lim = +∞,
(−1)n+1 xn+1 |x|n+1 |x|
(n + 1)! (n + 1)!
∞ ∞
1 n
.
2
= − x2 = (−1)n x2n
1+x n=0 n=0
if, and only if, |x2 | < 1, that is, |x| < 1. Then
x ∞
x ∞
x
1
arctan(x) = dt = (−1)n t2n dt = (−1)n t2n dt
0 1 + t2 0 n=0 n=0 0
.
∞
x ∞
t2n+1 x2n+1
= (−1)n = (−1)n
n=0
2n + 1 0 n=0
2n + 1
3.6. Solved Exercises 273
because power series are integrable term by term on every interval contained on its interval of convergence
(see Corollary 1 of Theorem 3.3.2).
By the same theorem, we have
x x ∞
∞ x
t2n+1 t2n+1
arctan(t) dt = (−1)n dt = (−1)n dt
0 0 n=0
2n + 1 n=0 0 2n + 1
∞
(−1) n 2n+2 x ∞
(−1) n 2n+2
t x
. = = ·
n=0
2n + 1 2n + 2 0 n=0
2n + 1 2n +2
∞
(−1)n
= x2n+2 .
n=0
(2n + 1)(2n + 2)
Therefore,
∞
(−1)n
.f (x) = x2n+2 ;
n=0
(2n + 1)(2n + 2)
We know that
1
∞ n ∞
1 5 1 2 2n
. = = − x = (−1)n n+1 xn ,
2x + 5 2 5 n=0 5 5
1+ x n=0
5
2 5 5
and this development is valid if − x < 1, that is, x ∈ − , . It is a power series, so it is differentiable
5 2 2
term by term on its interval of convergence (see Theorem 3.3.3), that is,
∞
∞ ∞
1 2n 2n n2n n−1 5 5
. = (−1)n xn = (−1)n xn = (−1)n x , ∀x ∈ − , .
2x + 5 n=0
5n+1 n=0
5n+1 n=1
5n+1 2 2
Then
∞ ∞
1 1 n2n n2n−1
f (x) = 2
=− (−1)n n+1 xn−1 = (−1)n+1 n+1 xn−1
(2x + 5) 2 n=1 5 n=1
5
.
∞ n
2 (n + 1) n 5 5
= (−1)n n+2
x , ∀x ∈ − , .
n=0
5 2 2
2
f) Let f (x) = + e2x . We know that
3 + 2x
∞ ∞
2 2 1 2 2x n 2n+1 xn
= · . = − = (−1)n n+1 ,
3 + 2x 3 1 + 2x 3 n=0 3 n=0
3
3
2x 3 3
and this development is valid if − < 1, that is, x ∈ − , .
3 2 2
274 Series of Functions
Then
∞ ∞ ∞
2 2n+1 xn 2n xn (−1)n 2 1
.f (x) = + e2x = (−1)n n+1 + = 2n n+1
+ xn ,
3 + 2x n=0
3 n=0
n! n=0
3 n!
3 3
development valid on − , .
2 2
g) By Example 3.4.4, we have the Maclaurin series representation of the function sin(x):
∞
x2n+1
. sin(x) = (−1)n , ∀x ∈ R.
n=0
(2n + 1)!
∞
x2n+1
. cos(x) = sin(x) = (−1)n , ∀x ∈ R,
n=0
(2n + 1)!
that is,
∞
∞
(2n + 1) x2n x2n
. cos(x) = (−1)n = (−1)n , ∀x ∈ R.
n=0
(2n + 1)! n=0
(2n)!
∞ 2 2n ∞
2 t4n
n t
. cos t = (−1) = (−1)n ,
n=0
(2n)! n=0
(2n)!
and
x ∞
x t4n
.f (x) = cos(t2 ) dt = (−1)n dt.
0 0 n=0
(2n)!
The interval of convergence of the series is R; therefore, by Corollary 1 of Theorem 3.3.2, the series is
integrable term by term on every closed interval of R. We have
∞
x ∞
t4n (−1)n x 4n
f (x) = (−1)n dt = t dt
0 n=0
(2n)! n=0
(2n)! 0
.
∞
x ∞
(−1)n t4n+1 (−1)n
= = x4n+1 .
n=0
(2n)! 4n + 1 0 n=0
(2n)!(4n + 1)
∞
(−1)n
.f (x) = x4n+1 , ∀x ∈ R.
n=0
(2n)!(4n + 1)
3.6. Solved Exercises 275
1 1
h) Let us consider the following derivative: = . Based on this, we can express f (x) as
1−x (1 − x)2
follows:
x + x2 1
.f (x) = = (x + x2 ) .
(1 − x)2 1−x
We know that
∞
1
. = xn .
1−x n=0
Moreover, this development is valid on the interval ] − 1, 1[. According to Theorem 3.3.3, this power series
is differentiable term by term on its interval of convergence, that is,
∞
∞ ∞
1
. = xn = (xn ) = n xn−1 , ∀x ∈ ] − 1, 1[.
1−x n=0 n=0 n=1
Then
∞
∞
∞
1
f (x) = (x + x2 ) = (x + x2 ) n xn−1 = x n xn−1 + x2 n xn−1
1−x n=1 n=1 n=1
∞
∞
∞
∞
. = n xn + n xn+1 = (n + 1) xn+1 + n xn+1
n=1 n=1 n=0 n=1
∞
∞
= x+ (2n + 1) xn+1 = (2n + 1) xn+1 , ∀x ∈ ] − 1, 1[.
n=1 n=0
1
i) Let f (x) = x arctan(x). We know that arctan(x) = , and this function is the sum of a geometric
1 + x2
series with ratio r = −x2 and the first term equal to 1. Therefore,
∞ ∞
1
.
2
= (−x2 )n = (−1)n x2n
1+x n=0 n=0
if and only if | − x2 | < 1, that is, if and only if |x| < 1. Then, if x ∈ ] − 1, 1[,
x ∞
x ∞
x
1
arctan(x) = dt = (−1)n t2n dt = (−1)n t2n dt
0 1 + t2 0 n=0 n=0 0
.
∞
x ∞
t2n+1 x2n+1
= (−1)n = (−1)n
n=0
2n + 1 0 n=0
2n + 1
because power series, by Corollary 1 of Theorem 3.3.2, are integrable term by term on every closed interval
contained in their interval of convergence. Therefore,
∞
∞
x2n+1 x2n+2
.f (x) =x (−1)n = (−1)n ,
n=0
2n + 1 n=0
2n +1
this development is valid if | − x2 | < 1, that is, if x ∈ ] − 1, 1[. This is a power series that, by Theorem 3.3.3,
is differentiable term by term on its interval of convergence, that is,
∞
∞ ∞
1
.
2
= (−1) xn 2n
= (−1)n x2n = (−1)n 2n x2n−1 , ∀x ∈ ] − 1, 1[.
1+x n=0 n=0 n=1
Then
∞ ∞
2x
.f (x) = 2 2
=− (−1)n 2n x2n−1 = (−1)n+1 2n x2n−1 , ∀x ∈ ] − 1, 1[.
(1 + x ) n=1 n=1
3x2
k) Let f (x) = log 2 + x3 . We have f (x) = log 2 + x3 and =
2 + x3
∞ n ∞
1 1 1 1 x3 (−1)n 3n
.
3
= · 3 = − = x ;
2+x 2 x 2 n=0 2 2n+1
1− − n=0
2
x3 √
this development is valid if, and only if, − < 1, that is, if and only if |x| < 3 2. Thus,
2
∞ ∞
(−1)n 3n 3 (−1)n 3n+2
.f (x) = 3x2 n+1
x = x ,
n=0
2 n=0
2n+1
√ √
if and only if x ∈ ] − 3 2, 3 2[.
x
Since f (t) dt = f (x) − f (0) = f (x) − log(2), we have
0
x ∞
x 3 (−1)n 3n+2
f (x) = f (t) dt + log(2) = log(2) + t dt
0 0 n=0
2n+1
.
∞
x
3 (−1)n
= log(2) + t3n+2 dt
n=0
2n+1 0
because, by Corollary 1 of Theorem 3.3.2, a power series is integrable term by term on its interval of
convergence. Then
∞
x ∞
3 (−1)n t3n+3 3 (−1)n x3n+3
f (x) = log(2) + n+1
= log(2) +
n=0
2 3n + 3 0 n=0
2n+1 3n + 3
.
∞
(−1)n
= log(2) + x3n+3 ,
n=0
2n+1 (n + 1)
√ √
and the development is valid on ] − 3
2, 3
2[.
3.6. Solved Exercises 277
!
1+x
l) Let f (x) = log , and calculate its derivative:
1−x
1+x
! 1−x
!
! 1+x 1+x 2
2
1+x 1−x 1−x (1 − x)2 1
.f (x) = log = ! = ! = = .
1−x 1+x 1+x 2
1+x 1 − x2
1−x 1−x 1−x
1
Knowing that is the sum of a geometric series with ratio r = x2 and the first term equal to 1, we
1 − x2
have
∞
1
. = x2n , |x2 | < 1,
1−x 2
n=0
the development is valid on ] − 1, 1[. As, by Corollary 1 of Theorem 3.3.2, a power series is term by term
integrable on its interval of convergence, we have
x ∞ ∞ x ∞ 2n+1 x ∞
t x2n+1
.f (x) − f (0) = f (x) = t2n dt = t2n dt = = .
0 n=0 n=0 0 n=0
2n + 1 0 n=0
2n + 1
Then
∞
1
.f (x) = log(4) − x2n+2 ;
n=0
22n+1 (2n + 2)
this development is valid on ] − 2, 2[.
278 Series of Functions
1 1
30. a) Since = , we can write
1−x (1 − x)2
x 1
.f (x) = =x .
(1 − x)2 1−x
1
We know that is the sum of a geometric series with ratio r = x and the first term equal to 1.
1−x
Therefore,
∞
1
. = xn ;
1−x n=0
this development is valid on ]−1, 1[. It is a power series that, by Theorem 3.3.3, is term by term differentiable
on its interval of convergence, that is,
∞
∞ ∞
1
. = xn = (xn ) = n xn−1 , ∀x ∈ ] − 1, 1[.
1−x n=0 n=0 n=1
Then
∞ ∞
x
.f (x) = =x n xn−1 = n xn , ∀x ∈ ] − 1, 1[.
(1 − x) 2
n=1 n=1
∞ ∞ n
n 1
.
n
= n .
n=0
3 n=1
3
1
By item a), with x = ,
3
n 1
∞
1 1 3 3
. n =f = 2 = .
3 3 1 4
n=1 1−
3
⎧
⎨ sin(x)
, if x = 0
31. Let f (x) = x
⎩1, if x = 0.
a) We know that
∞
x2n+1
. sin(x) = (−1)n , ∀x ∈ R;
n=0
(2n + 1)!
therefore,
∞
sin(x) x2n
. = (−1)n , ∀x ∈ R \ {0}.
x n=0
(2n + 1)!
Observing that for x = 0, the previous series is convergent and has sum 1, we conclude that
∞
x2n
.f (x) = (−1)n , ∀x ∈ R.
n=0
(2n + 1)!
3.6. Solved Exercises 279
b) Given the uniqueness of the power series development (see Theorem 3.4.3), the series obtained in item a)
f (n) (0)
is the Maclaurin series of f , so, for each n ∈ N, the coefficient an of xn in the previous series is ,
n!
that is, if ⎧
⎪ k
⎨ (−1) , if n = 2k, k ∈ N
0
.an = (2k + 1)!
⎪
⎩0, if n = 2k + 1, k ∈ N0 ,
then ⎧ k
⎨ (−1)
(n) , if n = 2k, k ∈ N
.f (0) = n! an = 2k + 1
⎩
0, if n = 2k + 1, k ∈ N0 .
2
32. a) Let f (x) = ex + cos(2x). Taking into account that
∞
x2n
. cos(x) = (−1)n , ∀x ∈ R,
n=0
(2n)!
we can write
∞
(2x)2n
. cos(2x) = (−1)n , ∀x ∈ R.
n=0
(2n)!
Knowing that
∞
xn
.e
x
= , ∀x ∈ R,
n=0
n!
we have
∞
x2 (x2 )n
.e = , ∀x ∈ R.
n=0
n!
∞ ∞ ∞
(x2 )n (2x)2n 1 22n
.f (x) = + (−1)n = + (−1)n x2n ,
n=0
n! n=0
(2n)! n=0
n! (2n)!
which is valid on R.
b) By Theorem 3.4.3, the power series of x of a function f is unique. Therefore, the series we obtained in the
previous item is the Maclaurin series of f , that is,
∞ ∞
f (n) (0) n 1 22n
. x = + (−1)n x2n .
n=0
n! n=0
n! (2n)!
Considering that in this series, all terms of odd order are zero, all derivatives of odd order of f at x = 0 are
zero. In particular, f (17) (0) = 0.
∞
∞
x2n+1 1
33. a) We know that sin(x) = (−1)n , ∀x ∈ R. If x = 1, we obtain (−1)n = sin(1).
n=0
(2n + 1)! n=0
(2n + 1)!
∞ ∞
xn 2n
b) Upon substituting x = 2 in the equality ex = , which holds true on R, we obtain e2 = .
n=0
n! n=0
n!
280 Series of Functions
∞
1
c) We have the following equality, valid on ] − 1, 1[ : = xn . By Corollary 1 of Theorem 3.3.2, we
1−x n=0
can integrate this series term by term obtaining
∞ ∞
xn+1 xn
. − log(1 − x) + log(1) = = .
n=0
n+1 n=1
n
1
If x = − , we arrive at the following result:
2
∞ ∞
3 (− 12 )n 1
. − log = = (−1)n n .
2 n=1
n n=1
2 n
2x
.f (x) = − 2 sin(2x).
x2 + 1
we can write
∞
(2x)2n+1
. sin(2x) = (−1)n , ∀x ∈ R.
n=0
(2n + 1)!
1
Knowing that is the sum of a geometric series with ratio r = −x2 and the first term equal to 1, we
1 + x2
have
∞
1
.
2
= (−1)n x2n , | − x2 | < 1.
1+x n=0
∞
∞
(2x)2n+1
f (x) = 2x (−1)n x2n − 2 (−1)n
n=0 n=0
(2n + 1)!
∞
∞
22n+2 x2n+1
. = (−1)n 2 x2n+1 − (−1)n
n=0 n=0
(2n + 1)!
∞
22n+2
= (−1) 2− n
x2n+1 ,
n=0
(2n + 1)!
x ∞
22n+2
f (x) − f (0) = f (x) − 1 = (−1)n 2− t2n+1 dt
0 n=0 (2n + 1)!
∞ x
22n+2
= (−1)n 2 − t2n+1 dt
n=0
(2n + 1)! 0
.
∞
2 22n+2
= (−1)n − x2n+2
n=0
2n + 2 (2n + 2)!
∞
1 22n+2
= (−1)n − x2n+2 .
n=0
n+1 (2n + 2)!
∞
∞
1 22n+2 1 22n
.1 + (−1)n − x2n+2 = 1 + (−1)n−1 − x2n ,
n=0
n+1 (2n + 2)! n=1
n (2n)!
∞
π 2n 1 22n
.1 + (−1)n−1 − .
n=1
42n n (2n)!
π π
π 2 π 2
This series has sum f = log + 1 + cos 2 = log +1 .
4 4 4 4
Considering that
∞
x2n+1
. sin(x) = (−1)n , ∀x ∈ R,
n=0
(2n + 1)!
we get
∞
(2x)2n+1
. sin(2x) = (−1)n , ∀x ∈ R.
n=0
(2n + 1)!
1
Since is the sum of a geometric series with ratio r = x2 and the first term equal to 1, we have
1 − x2
∞
1
. = x2n , |x2 | < 1.
1−x 2
n=0
282 Series of Functions
As, by Theorem 3.3.3, power series are differentiable term by term on their interval of convergence,
∞
∞
(2x)2n+1
f (x) = (−1)n + x2n
n=0
(2n + 1)! n=0
∞
∞
(2x)2n+1
= (−1)n + 2nx2n−1
n=0
(2n + 1)! n=1
.
∞ ∞
(2x)2n+1
= (−1)n + 2(n + 1)x2n+1
n=0
(2n + 1)! n=0
∞
22n+1
= (−1)n + 2(n + 1) x2n+1 ,
n=0
(2n + 1)!
this development is valid if |x − 3| < 1, that is, x ∈ ]2, 4[. It is a power series so, by Corollary 1 of Theorem 3.3.2,
it is integrable term by term on its interval of convergence, that is,
x x ∞
∞ x ∞
1 (x − 3)n+1
. log(4 − x) = − dt = − (t − 3) dt = − (t − 3)n dt = −
n
.
3 4 − t 3 n=0 n=0 3 n=0
n+1
Then
∞ ∞
1 (x − 3)n+1
f (x) = + log(4 − x) = (x − 3)n −
4−x n=0 n=0
n+1
.
∞
∞ ∞
(x − 3)n n−1
= 1+ (x − 3)n − =1+ (x − 3)n , ∀x ∈ ]2, 4[.
n=1 n=1
n n=1
n
1 1 1
37. Let f (x) = . Considering that = − , we can write
x2 x2 x
1 1 1 1
.f (x) = − =− =− .
x 3+x−3 3 1 + x−3
3
3.6. Solved Exercises 283
Since
∞
n ∞
1 x−3 (x − 3)n
.
x−3
= − = (−1)n ,
1+ 3 n=0
3 n=0
3n
we also have
∞
∞
1 (x − 3)n (x − 3)n
.f (x) =− (−1)n = (−1)n+1 ;
3 n=0
3n n=0
3n+1
this development is valid if |x − 3| < 3, that is, x ∈ ]0, 6[. It is a power series that, by Theorem 3.3.3, is
differentiable term by term on its interval of convergence, that is,
∞
∞
(x − 3)n n (x − 3)n−1
.f (x) = (−1)n+1 = (−1)n+1 .
n=0
3n+1 n=1
3n+1
1
38. Let f (x) = . As x2 + x − 6 = (x − 2)(x + 3), we get
x2 + x − 6
so ⎧
⎪ 1
A+B =0 ⎨A =
. ⇔ 5
3A − 2B = 1 ⎪ 1
⎩B = − ;
5
therefore,
1 1 1 1 1 1 1 1
f (x) = · − · =− · − ·
5 x−2 5 x+3 5 2−x 5 3+x
1 1 1 1
= − · − ·
. 5 3 − (x + 1) 5 2 + (x + 1)
1 1 1 1
= − · − · .
15 x+1 10 x+1
1− 1− −
3 2
1 x+1
Knowing that is the sum of a geometric series with ratio r = and the first term equal to 1 and
x+1 3
1−
3
1 x+1
is the sum of a geometric series with ratio r = − and the first term equal to 1, we have
x+1 2
1− −
2
∞
1 x+1 n
. =
x+1 3
1− n=0
3
and
∞ ∞
1 x+1 n x+1 n
. = − = (−1)n ,
x+1 2 2
1− − n=0 n=0
2
284 Series of Functions
x + 1
equalities valid if < 1 and − x + 1 < 1, respectively. If x ∈ ] − 3, 1[, we can write the development of f
3 2
in powers of x + 1:
∞ ∞ ∞
1 x+1 n 1 x+1 n 1 (−1)n+1 1
.f (x) =− − (−1)n = n+1
− n+1 (x + 1)n .
15 n=0 3 10 n=0 2 n=0
5 2 3
1 1 1
39. Let g(t) = log(t). Since log(t)
and = = is the sum of a geometric series with ratio
t t 1 + (t − 1)
r = −(t − 1) and the first term equal to 1, we have
∞ ∞
1 n
. = − (t − 1) = (−1)n (t − 1)n
1 + (t − 1) n=0 n=0
if, and only if, |t − 1| < 1, that is, t ∈ ]0, 2[. Then
∞
t ∞
t
log(t) − log(1) = log(t) = (−1)n (x − 1)n dx = (−1)n (x − 1)n dx
1 n=0 n=0 1
.
∞
t ∞
(x −
1)n+1 (t − 1)n+1
= (−1)n = (−1)n ,
n=0
n+1 1 n=0
n+1
because, by Corollary 1 of Theorem 3.3.2, power series are integrable term by term on every interval contained on
their interval of convergence. Again, by Corollary 1 of Theorem 3.3.2, we have
∞
∞
(t − 1)n+1 (t − 1)n+1
(−1)n − (t − 1) (−1)n
x
n=0
n+1 x
n=1
n+1
f (x) − f (1) = f (x) = dt = dt
1 t−1 t−1 1
x∞ ∞
(t − 1)n x (t − 1)n
= (−1)n dt = (−1)n dt
. 1 n=1 n+1 n=1 1 n+1
∞
(−1)n (t − 1)n+1 x ∞
(−1)n (x − 1)n+1
= = ·
n=1
n + 1 n + 1 1 n=1
n+1 n+1
∞
(−1)n
= (x − 1)n+1 .
n=1
(n + 1)2
Therefore,
∞
(−1)n
.f (x) = (x − 1)n+1 .
n=1
(n + 1)2
∞ ∞
1 n
. = − (t − 2)2 = (−1)n (t − 2)2n
1 + (t − 2) 2
n=0 n=0
3.6. Solved Exercises 285
if, and only if, |(t − 2)2 | < 1, that is, t ∈ ]1, 3[. Then, by Corollary 1 of Theorem 3.3.2,
∞
t ∞
t
arctan(t − 2) − arctan(0) = arctan(t − 2) = (−1)n (x − 2)2n dx = (−1)n (x − 2)2n dx
2 n=0 n=0 2
.
∞
t ∞
(x − 2)2n+1 (t − 2)2n+1
= (−1)n = (−1)n .
n=0
2n + 1 2 n=0
2n + 1
∞
∞
(t − 2)2n+1 (t − 2)2n+1
(−1)n − (t − 2) (−1)n
x
n=0
2n + 1 x
n=1
2n + 1
f (x) = dt = dt
2 (t − 2)2 2 (t − 2)2
∞
x ∞ x
(t − 2)2n−1 (t − 2)2n−1
= (−1)ndt = (−1)n dt
. 2 n=1 2n + 1 n=1 2 2n + 1
∞ x ∞
(−1) n (t − 2) 2n (−1) n (x − 2) 2n
= = ·
n=1
2n + 1 2n 2 n=1
2n + 1 2n
∞
(−1)n
= (x − 2)2n .
n=1
2n (2n + 1)
Therefore,
∞
(−1)n
.f (x) = (x − 2)2n .
n=1
2n (2n + 1)
∞
f (n) (0) n
41. We know that the Maclaurin series is x . Therefore, we need to calculate f (n) (0), ∀n ∈ N. We will
n=0
n!
prove, by induction, that f (n) (x) = f (x), ∀n > 2, ∀x ∈ R.
By hypothesis, f (x) = f (x) + x, ∀x ∈ R, which implies that f (x) = f (x) + 1 and f (x) = f (x), ∀x ∈ R
which proves the equality for n = 3.
Suppose that f (n) (x) = f (x), ∀x ∈ R. We prove that f (n+1) (x) = f (x), ∀x ∈ R.
.f
(n+1)
(x) = f (n) (x) = f (x) = f (x) = f (x).
Then, f (0) = f (0) + 0 = 1 and f (n) (0) = f (0) = f (0) + 1 = 2, so the Maclaurin series of f is
∞ ∞
f (n) (0) n 2 n
. x =1+x+ x .
n=0
n! n=2
n!
π, if − π < x ≤ 0
42. a) Let f (x) =
0, if 0 < x < π.
286 Series of Functions
(a) The function f (b) The periodic extension, f˜, of the function f
The graph of f can be seen in Fig. 3.29. By Definition 3.5.3, the Fourier coefficients are
1 π 1 0 " #0
a0 = f (x) dx = π dx = x −π = π;
π −π π −π
0 0
1 sin(nx)
an = π cos(nx) dx = = 0;
. π −π n −π
⎧
0 ⎨0, if n is even
1 0 cos(nx) 1 cos(−n π) (−1)n − 1
bn = π sin(nx) dx = − =− + = =
π −π n −π n n n ⎩− 2 , if n is odd.
n
Since f is a piecewise C 1 function, by Theorem 3.5.1 the series converges to f˜(x) if x = kπ, k ∈ Z, and
π
takes the value , if x = kπ, k ∈ Z. In Fig. 3.30, we see an illustration of the sum of the series.
2
0, if −π <x≤0
b) Let us consider the function f (x) =
ex , if 0 < x < π.
The graph of f can be seen in Fig. 3.31. By Definition 3.5.3, the Fourier coefficients are
1 π 1 π x 1 x π 1
a0 = f (x) dx = e dx = e = (eπ − 1);
π −π π 0 π 0 π
1 π x
an = e cos(nx) dx;
π 0
. π π
1 π x 1 x
bn = e sin(nx) dx = e sin(nx) − n ex cos(nx) dx
π 0 π 0 0
n π x
= − e cos(nx) dx.
π 0
We have
π π π π
ex cos(nx) dx = ex cos(nx) + n ex sin(nx) dx = eπ cos(nπ) − e0 + n ex sin(nx) dx
0 0 0 0
π π
. = (−1)n eπ − 1 + n ex sin(nx) − n ex cos(nx) dx
0 0
π
= (−1)n eπ − 1 − n2 ex cos(nx) dx.
0
therefore, π (−1)n eπ − 1
. ex cos(nx) dx = .
0 n2 + 1
Using this equality, the coefficients are
1 π (−1)n eπ − 1
an = ex cos(nx) dx = ;
π 0 π(n2 + 1)
.
n (−1)n eπ − 1
bn = − · .
π n2 + 1
The Fourier series of f is
∞
eπ − 1 (−1)n eπ − 1 n (−1)n eπ − 1
+ cos(nx) − · sin(nx)
2π n=1
π(n2 + 1) π n2 + 1
.
∞
eπ − 1 1 (−1)n eπ − 1
= + cos(nx) − n sin(nx) .
2π π n=1 n2 + 1
Since f is a piecewise C 1 function, by Theorem 3.5.1 the series converges pointwise to the function
⎧˜
⎪
⎪ f (x), if x = kπ, k ∈ Z
⎪
⎪
⎪
⎨1
.S(x) = , if x = 2kπ, k ∈ Z
⎪
⎪ 2
⎪
⎪
⎪ eπ
⎩ , if x = (2k + 1)π, k ∈ Z.
2
Figure 3.32 displays the sum of the series.
288 Series of Functions
cos(x), if − π < x ≤ 0
c) Let f be the function defined by f (x) =
sin(x), if 0 < x < π.
The graph of f can be seen in Fig. 3.33. Using equalities (3.7), (3.8), and (3.11), we can compute the
Fourier coefficients:
π 0 π
1 1
. a0 = f (x) dx = cos(x) dx + sin(x) dx
π −π π −π 0
π
1 0 2
= sin(x) + − cos(x) = ;
π −π 0 π
0 π
1
an = cos(x) cos(nx) dx + sin(x) cos(nx) dx
π −π 0
1 1 0 1 π
= cos (n + 1)x + cos (n − 1)x dx + sin (n + 1)x + sin (1 − n)x dx
π 2 −π 2 0
⎧ ⎛$ %0 $ %π ⎞
⎪
⎪ 1 ⎝ sin (n + 1)x sin (1 − n)x cos (n + 1)x cos (1 − n)x
⎪
⎪ + − − ⎠, if n = 1
⎪
⎨ 2π
+
n+1 1−n n+1 1−n
= −π 0
⎪ 0
⎪
⎪ 1 sin(2x) cos(2x) π
⎪
⎪ −
⎩ + x + , if n = 1
2π 2 −π 2 0
⎧
⎪
⎪ 1 cos (n + 1)π cos (1 − n)π 1 1
⎪
⎨ − − + + , if n = 1
2π n+1 1−n n+1 1−n
=
⎪
⎪
⎪
⎩1, if n = 1
2
3.6. Solved Exercises 289
⎧
⎪
⎪ 1 (−1)n + 1 (−1)n + 1 1 (−1)n + 1
⎪
⎨ 2π − =− · , if n = 1
n+1 n−1 π n2 − 1
=
⎪
⎪
⎪ 1
⎩ , if n = 1;
2
0 π
1
bn = cos(x) sin(nx) dx + sin(x) sin(nx) dx
π −π 0
0
1 1 1 π
= sin (n + 1)x + sin (n − 1)x dx + cos (n − 1)x − cos (n + 1)x dx
π 2 −π 2 0
⎧ ⎛$ %0 $ %π ⎞
⎪
⎪ 1 ⎝ cos (n + 1)x cos (n − 1)x sin (n − 1)x sin (n + 1)x
⎪
⎪ − − − ⎠, if n = 1
⎪
⎨ 2π
+
n+1 n−1 n−1 n+1
= −π 0
⎪
⎪
⎪ 1 cos(2x) 0 sin(2x) π
⎪
⎪ − −
⎩ + x , if n = 1.
2π 2 −π 2 0
⎧
⎪
⎪ 1 1 1 cos (n + 1)(−π) cos (n − 1)(−π)
⎪
⎨ − − + + , if n = 1
2π n+1 n−1 n+1 n−1
=
⎪
⎪
⎪
⎩1, if n = 1
2
⎧
⎪
⎪ 1 (−1)n+1 − 1 (−1)n−1 − 1 1 n 1 + (−1)n
⎪
⎨ 2π + = − · , if n = 1
n+1 n−1 π n2 − 1
=
⎪
⎪1
⎪
⎩ , if n = 1.
2
∞
1 1 1 1 (−1)n + 1
. + cos(x) + sin(x) − cos(nx) + n sin(nx)
π 2 2 π n=2 n2 − 1
∞
1 1 2 1
. = + cos(x) + sin(x) − cos(2nx) + 2n sin(2nx) .
π 2 π n=1 4n2 − 1
Since f is a piecewise C 1 function, by Theorem 3.5.1 the series converges pointwise to the function
⎧
⎪
⎪ f˜(x), if x = kπ, k ∈ Z
⎪
⎪
⎨1
.S(x) = , if x = 2kπ, k ∈ Z
⎪
⎪ 2
⎪
⎪
⎩− 1 , if x = (2k + 1)π, k ∈ Z.
2
d) The function f (x) = x3 defined on ] − π, π[ is odd (see Fig. 3.35); therefore, an = 0, ∀n ∈ N0 . Let us
calculate the remaining Fourier coefficients.
Since f is a piecewise C 1 function, by Theorem 3.5.1 the series converges pointwise to the function
e) The function f (x) = 4 − |x| defined on ] − π, π[ is even (see Fig. 3.37); therefore, bn = 0, ∀n ∈ N. The
Fourier coefficients are
π π π
2 2 2 x2
a0 = f (x) dx = 4x −
(4 − x) dx = = 8 − π;
π0 0 π π 2 0
π π π
2 2 4 sin(nx)
. an = (4 − x) cos(nx) dx = − x cos(nx) dx
π 0 π n 0 0
π
2 x sin(nx) π sin(nx) 2 cos(nx) π 2(1 − (−1)n )
= − − dx = − 2
= .
π n 0 0 n π n 0 n2 π
∞ ∞
π 2 1 − (−1)n π 4 1
.4 − + cos(nx) = 4 − + cos (2n − 1)x .
2 π n=1 n2 2 π n=1 (2n − 1)2
As f˜ is a piecewise C 1 function and extendable by continuity to R, by Theorem 3.5.1 the series converges
pointwise to the extension by continuity of f˜ on R.
43. a) Let us consider the periodic function of period 2π defined on the interval ]0, 2π[ by f (x) = x (see Fig. 3.38).
292 Series of Functions
By Proposition 3.5.3, the coefficients of the Fourier series can be calculated by the formulas
2π 2π
1 1 x2
a0 = x dx = = 2π;
π 0 π 2 0
2π
1 2π 1 sin(nx) 2π sin(nx)
an = x cos(nx) dx = x − dx = 0;
π 0 π n 0 0 n
.
2π
1 2π 1 cos(nx) 2π cos(nx)
bn = x sin(nx) dx = −x + dx
π 0 π n 0 0 n
2π
1 2π sin(nx) 2
= − + =− .
π n n2 0 n
By Theorem 3.5.2, the series converges pointwise to the function, S, periodic of period 2π, illustrated in
Fig. 3.39 and whose analytical expression is
b) Let us consider the periodic function of period 2π defined on ]0, 2π[ by f (x) = x2 (refer to Fig. 3.40).
3.6. Solved Exercises 293
Using the results from the previous item and Proposition 3.5.3, the coefficients of the Fourier series are
2π 2π
1 1 x3
8π 2
a0 = x2 dx = ; =
π 0 0 π 33
2π 2π 2π
1 1 2 sin(nx) sin(nx) 4
. an = 2
x cos(nx) dx = x − 2x dx = 2 ;
π 0 π n 0 0 n n
2π 2π 2π
1 1 cos(nx) cos(nx) 4π
bn = 2
x sin(nx) dx = −x 2
+ 2x dx = − .
π 0 π n 0 0 n n
By Theorem 3.5.2, the series converges pointwise to the function, S, periodic of period 2π, illustrated in
Fig. 3.41 and whose analytical expression is
π 3π
c) Let us consider the periodic function of period 2π defined on − , by f (x) = |x| (see Fig. 3.42).
2 2
294 Series of Functions
By Theorem 3.5.1, the series converges pointwise to the function, S, periodic with period 2π, illustrated in
Fig. 3.43 and whose analytical expression is
⎧
⎪ 1 3
⎨|x − 2kπ|, if 2k − π < x < 2k + π, k ∈ Z
.S(x) = 2 3 2
⎪
⎩π, if x = 2k + π, k ∈ Z.
2
π 3π
d) Let f be the periodic function with period 2π defined on − , by
2 2
⎧ π π
⎪
⎨x, if − ≤x≤
2 2
.
⎪
⎩π − x, π 3π
if <x≤ .
2 2
3.6. Solved Exercises 295
It is easy to verify that it is an odd function (see Fig. 3.44). As it is periodic with period 2π, it follows from
Propositions 3.5.3 and 3.5.6 that an = 0, ∀n ∈ N0 . Let us calculate the remaining coefficients.
2 π/2 π
bn = x sin(nx) dx + (π − x) sin(nx) dx
π 0 π/2
π/2 π
2 cos(nx) cos(nx) π/2 cos(nx) π
= −x − dx + π −
− − x sin(nx) dx
π n
0 0 n n π/2 π/2
π/2 π
2 −π nπ sin(nx) π nπ
= cos + − cos(nπ) − cos − x sin(nx) dx
π 2n 2 n2 0 n 2 π/2
π
. 2 π nπ 1 nπ π cos(nx) π cos(nx)
= cos + 2 sin − (−1) − −xn
+ − dx
π 2n 2 n 2 n n π/2 π/2 n
2 π nπ 1 nπ π π π nπ sin(nx) π
= cos + 2 sin − (−1) + n
cos(nπ) − cos −
π 2n 2 n 2 n n 2n 2 n2 π/2
⎧
nπ ⎪
⎨ 0, if n is even
4
= sin =
πn 2 2 ⎪
⎩ 4 (−1)
(n−1)/2
, if n is odd.
n2 π
Since f is a continuous function and piecewise C 1 , Theorem 3.5.1 allows us to conclude that the series
converges pointwise to f on R.
π 3π
e) Let f be the periodic function of period 2π defined on − , by
2 2
⎧ π π
⎪ 2
⎨x , if − ≤x≤
2 2
.
⎪ 2
⎩π , π 3π
if <x≤ .
4 2 2
It is easy to verify that it is an even function (see Fig. 3.45). As it is periodic with period 2π, it follows from
Propositions 3.5.3 and 3.5.6 that bn = 0, ∀n ∈ N. Let us calculate the remaining coefficients.
π/2
1 3π/2 2 π/2 1 3π/2 π2 2 x3 π 3π/2 π2
a0 = f (x) dx = x2 dx + dx = + x = ;
π −π/2 π 0 π π/2 4 π 3 0 4 π/2 3
π/2
2 1 3π/2 π 2
an = x2 cos(nx) dx + cos(nx) dx
π 0 π π/2 4
π/2 π/2
2 2 sin(nx) sin(nx) π sin(nx) 3π/2
= x − 2x dx +
π n 0 0 n 4 n π/2
π/2
2 π 2 nπ 2 cos(nx) 2
π/2
cos(nx) π 3nπ nπ
. = sin + x − dx + sin − sin
π 4n 2 n n 0 n 0 n 4n 2 2
2 π2 nπ π nπ 2 sin(nx) π/2
π nπ
= sin + 2 cos − 2 − sin
π 4n 2 n 2 n n 0 2n 2
2
2 π nπ π nπ 2 nπ π nπ
= sin + 2 cos − 3 sin − sin
π 4n 2 n 2 n 2 2n 2
2 nπ 4 nπ
= cos − 3 sin .
n2 2 n π 2
∞ nπ nπ
π2 2 4
. + 2
cos − sin cos(nx).
6 n=1
n 2 n3 π 2
Since f is a continuous function and piecewise C 1 , by Theorem 3.5.1, the series converges pointwise to f
on R.
3.6. Solved Exercises 297
44. a) Let f be the periodic function of period 1 defined on the interval ]0, 1[ by f (x) = π sin(π x). It is easy to
verify that it is an even function (see Fig. 3.46) and, therefore, bn = 0, ∀n ∈ N. As it is periodic with period
2l = 1, it follows from Proposition 3.5.9 that the coefficients of its development in Fourier series are given
by
1/2
1 l " #1/2
a0 = f (x) dx = 4 π sin(πx) dx = 4 − cos(πx) 0 = 4;
l −l 0
.
nπx 1/2
1 l
an = f (x) cos dx = 4 π sin(πx) cos(2nπx) dx.
l −l l 0
$ %1/2
1/2 cos (2n + 1)π x cos (2n − 1)π x
.2 π sin (2n + 1)π x − sin (2n − 1)π x dx = 2 − +
0 2n + 1 2n − 1
0
and so
4
.an =− .
(2n − 1)(2n + 1)
Since f is a continuous function and piecewise C 1 , Theorem 3.5.2 allows us to conclude that the series
converges pointwise to f on R.
b) Let f be the periodic function of period 1 defined on the interval ]0, 1[ by f (x) = 2 − x2 (see Fig. 3.47).
As it is periodic with period 2l = 1, it follows from Propositions 3.5.9 and 3.5.3 that the coefficients of its
Fourier series development are given by
l 1/2 1 1
1 x3 10
a0 = f (x) dx = 2 f (x) dx = 2 (2 − x2 ) dx = 2 2x − = ;
l −l −1/2 0 3 0 3
l nπx 1
1
an = f (x) cos dx = 2 (2 − x2 ) cos(2nπx) dx
l −l l 0
1 1
sin(2nπx) 1 sin(2nπx) 1 sin(2nπx)
= 4 −2 x2 cos(2nπx) dx = −2 x2 − 2x dx
2nπ 0 0 2nπ 0 0 2nπ
1
2 1 2 cos(2nπx) 1 cos(2nπx)
= x sin(2nπx) dx = −x + dx
nπ 0 nπ 2nπ 0 0 2nπ
1
. 2 cos(2nπ) sin(2nπx) 1
= − + =− ;
nπ 2nπ (2nπ)2 0 (nπ)2
l nπx 1
1
bn = f (x) sin dx = 2 (2 − x2 ) sin(2nπx) dx
l −l l 0
1 1
cos(2nπx) 1 cos(2nπx) 1 cos(2nπx)
= 4 − −2 x2 sin(2nπx) dx = −2 −x2 + 2x dx
2nπ 0 0 2nπ 0 0 2nπ
1 1
1 1 sin(2nπx) 1 sin(2nπx)
= −2 − + x − dx
2nπ nπ 2nπ 0 nπ 0 2nπ
1 1 cos(2nπ) 1 1
= + − = .
nπ (nπ)2 2nπ 0 nπ
Since f is piecewise C 1 , by Theorem 3.5.2 the series converges pointwise to the function, S, periodic with
period 1, illustrated in Fig. 3.48 and analytically defined by
⎧
⎪
⎨f (x), if x ∈/Z
.S(x) =
⎪
⎩ ,
3
if x ∈ Z.
2
c) Let f be the periodic function with period 4 defined on the interval ]0, 4] by f (x) = x − 3 (see Fig. 3.49).
As it is periodic with period 2l = 4, it follows from Proposition 3.5.9 that the coefficients of its development
in Fourier series are given by
l 2 4 4
1 1 1 1 x2
a0 = f (x) dx = f (x) dx = (x − 3) dx = − 3x = −2;
l −l 2 −2 2 0 2 2 0
. l nπx 4 nπx
1 1
an = f (x) cos dx = (x − 3) cos dx
l −l l 2 0 2
3.6. Solved Exercises 299
nπx nπx 4 4
1 4 6 nπx 4 1 nπx
= x cos dx − sin = x sin − sin dx
2 0 2 nπ 2 0 nπ 2 0 0 2
2 nπx 4
= cos = 0;
(nπ)2 2 0
nπx nπx
1 l 1 4
. bn = f (x) sin dx = (x − 3) sin dx
l −l l 2 0 2
nπx nπx 4 nπx 4 4
1 4 6 1 nπx
= x sin cos dx + = −x cos + cos dx
2 0 2 nπ 2 0 nπ 2 0 0 2
1 2 nπx 4 4
= −4 + sin =− .
nπ nπ 2 0 nπ
Since f is piecewise C 1 , Theorem 3.5.2 allows us to conclude that the series converges pointwise to the
function, S, periodic with period 4, illustrated in Fig. 3.50.
300 Series of Functions
f (x), if x = 4k, k ∈ Z
.S(x) =
−1, if x = 4k, k ∈ Z.
(see Fig. 3.51). As it is periodic with period 2l = 4, it follows from Proposition 3.5.9 that the coefficients
of its Fourier series development are given by
l 2 1 2
1 1 1 1
.a0 = f (x) dx = f (x) dx = x dx + 1 dx = ;
l −l 2 −2 2 −1 1 2
1 nπx 2 1 nπ
= sin =− sin ;
nπ 2 1 nπ 2
3.6. Solved Exercises 301
l nπx 1 nπx 2 nπx
1 1
bn = f (x) sin dx = x sin dx + sin dx
l −l l 2 −1 2 1 2
nπ −nπ nπ
1 2 nπx 1
= − cos − cos + sin − cos(nπ) + cos
nπ 2 2 nπ 2 −1 2
nπ nπ
1 4
= − cos + sin − (−1)n .
nπ 2 nπ 2
∞
nπ nπx nπ nπ nπx
1 1 1 4
. + − sin cos + − cos + sin − (−1)n sin .
4 n=1 nπ 2 2 nπ 2 nπ 2 2
As f is piecewise C 1 , by Theorem 3.5.2 the series converges pointwise to the function, S, periodic with
period 4, illustrated in Fig. 3.52 and whose analytical expression is
⎧
⎪
⎪f (x), if x = 4k + 2 ∧ x = 4k + 3, k ∈ Z
⎪
⎪
⎪
⎨1
.S(x) =
, if x = 4k + 2, k ∈ Z
⎪ 2
⎪
⎪
⎪
⎪
1
⎩− , if x = 4k + 3, k ∈ Z.
2
e) Let us consider the function f , periodic with period 6, defined on the interval ]−3, 3] by
⎧ x
⎪
⎨− , if − 3 < x ≤ 0
.f (x) =
3
⎪ x2
⎩2x − , if 0 < x ≤ 3
3
302 Series of Functions
As it is periodic with period 2l = 6 (see Fig. 3.53), it follows from Proposition 3.5.9 that the coefficients of
its Fourier series development are given by
l 0 3
1 1 3 1 x x2
a0 = f (x) dx = f (x) dx = − dx + 2x − dx
l −l 3 −3 3 −3 3 0 3
0 3
1 x2 x3 5
= − + x2 − = ;
3 6 −3 9 0 2
l 0 nπx 3 nπx
1 nπx 1 x x2
an = f (x) cos dx = − cos dx + 2x − cos dx
l −l l 3 −3 3 3 0 3 3
0 3 2
1 x nπx 0 1 nπx 1 x nπx
= − sin + sin dx + 2x − cos dx
nπ 3 3 −3 3 −3 3 3 0 3 3
1 nπx 0 1 x2 nπx 3 3 2x nπx
= − cos + 2x − sin − 2 − sin dx
(nπ)2 3 −3 nπ 3 3 0 0 3 3
3 3
1 1 2x nπx nπx
= (−1 + cos(nπ)) − −3 2 − cos − 2 cos dx
(nπ)2 (nπ)2 3 3 0 0 3
.
nπx 3 n−7
1 1 6 (−1)
= (−1 + (−1)n ) − 6− sin = ;
(nπ)2 (nπ)2 nπ 3 0 (nπ)2
nπx 0 nπx 3 nπx
1 l 1 x x2
bn = f (x) sin dx = − sin dx + 2x − sin dx
l −l l 3 −3 3 3 0 3 3
0 3 2
1 x nπx 0 1 nπx 1 x nπx
= cos − cos dx + 2x − sin dx
nπ 3 3 −3 3 −3 3 3 0 3 3
1 1 nπx 0 x2 nπx 3 3 2x nπx
= (−1)n − sin + − 2x − cos + 2− cos dx
nπ nπ 3 −3 3 3 0 0 3 3
3 3
1 3 2x nπx 3 2 nπx
= (−1)n − 3 cos(nπ) + 2− sin + sin dx
nπ nπ 3 3 0 nπ 0 3 3
nπx 3
1 6 6(1 − (−1)n ) 2(−1)n
= −2(−1)n + − cos = − .
nπ (nπ)2 3 0 (nπ)3 nπ
∞
nπx nπx
5 (−1)n − 7 6(1 − (−1)n ) 2(−1)n
. + cos + − sin .
4 n=1 (nπ)2 3 (nπ)3 nπ 3
3.6. Solved Exercises 303
Since f is piecewise C 1 , Theorem 3.5.2 allows us to conclude that the series converges pointwise to the
function, S, periodic of period 4, illustrated in Fig. 3.54 and whose analytical expression is
f (x), if x = 6k + 3, k ∈ Z
.S(x) =
2, if x = 6k + 3, k ∈ Z.
45. In Fig. 3.55 we can see the graph of the periodic function of period 2π defined on the interval [0, 2π] in the
following way: ⎧a
⎪
⎪ x, if 0 ≤ x ≤ b
⎪
⎪
⎪
⎪
b
⎪
⎪
⎪
⎪ a, if b < x ≤ π − b
⎪
⎪
⎨a
.f (x) = (π − x), if π − b < x ≤ π + b
⎪
⎪ b
⎪
⎪
⎪
⎪ −a, if π + b < x ≤ 2π − b
⎪
⎪
⎪
⎪
⎪
⎪ a
⎩ (x − 2π), if 2π − b < x ≤ 2π
b
for the case a = b = 1.
It is easy to see that f is odd; therefore, an = 0, ∀n ∈ N0 . By Proposition 3.5.3 the coefficients bn can be
calculated by the formulas:
2π π π
1 1 2
.bn = f (x) sin(nx) dx = f (x) sin(nx) dx = f (x) sin(nx) dx
π 0 π −π π 0
b π−b π
2 a a
= x sin(nx) dx + a sin(nx) dx + (π − x) sin(nx) dx
π 0 b b π−b b
304 Series of Functions
2 a cos(nx) b a b cos(nx) cos(nx) π−b a π
= −x + dx − a + (π − x) sin(nx) dx
π b n 0 b 0 n n b b π−b
b π−b
2 a a sin(nx) cos(nx) a cos(nx) π a π cos(nx)
= − cos(nb) + −a − (π − x) − dx
π n b n2 0 n b b n π−b b π−b n
π−b
2 a a cos(nx) a a π cos(nx)
= − cos(nb) + 2 sin(nb) − a + cos(nπ − nb) − dx
π n n b n b n b π−b n
π
2 a a a a a sin(nx)
= − cos(nb) + 2 sin(nb) − (cos(nπ − nb) − cos(nb)) + cos(nπ − nb) −
π n n b n n b n2 π−b
2 a a 2a 2a
= sin(nb) + 2 sin(nπ − nb) = (sin(nb) + sin(nπ − nb)) = (1 − (−1)n ) sin(nb).
π n2 b n b b n2 π b n2 π
Since f is a continuous function and piecewise of class C 1 , Theorem 3.5.1 allows us to conclude that the series
converges pointwise to f .
∞
a0
. + (an cos(nx) + bn sin(nx))
2 n=1
f (x) + f (−x)
Let us calculate the Fourier coefficients of g(x) = .
2
π π π π
1 1 f (x) + f (−x) 1
.c0 = g(x) dx = dx = f (x) dx + f (−x) dx
π −π π −π 2 2π −π −π
π
1
= f (x) dx = a0 ;
π −π
3.6. Solved Exercises 305
π π
1 1 f (x) + f (−x)
cn = g(x) cos(nx) dx = cos(nx) dx
π −π π −π 2
π π
1
= f (x) cos(nx) dx + f (−x) cos(nx) dx
2π −π −π
π
1
= f (x) cos(nx) dx = an ;
π −π
π π
1 1 f (x) + f (−x)
dn = g(x) sin(nx) dx = sin(nx) dx
π −π π −π 2
π π
1
= f (x) sin(nx) dx + f (−x) sin(nx) dx = 0.
2π −π −π
f (x) − f (−x)
The coefficients of h(x) = are
2
π π
1 π 1 π f (x) − f (−x) 1
c∗0 = h(x) dx = dx = f (x) dx − f (−x) dx = 0;
π −π π −π 2 2π −π −π
1 π 1 π f (x) − f (−x)
c∗n = h(x) cos(nx) dx = cos(nx) dx
π −π π −π 2
π π
1
= f (x) cos(nx) dx − f (−x) cos(nx) dx = 0;
2π −π −π
.
1 π 1 π f (x) − f (−x)
d∗n = h(x) sin(nx) dx = sin(nx) dx
π −π π −π 2
π π
1
= f (x) sin(nx) dx − f (−x) sin(nx) dx
2π −π −π
1 π
= f (x) sin(nx) dx = bn .
π −π
47. Let a ∈ R. From f (x + π) = −f (x), ∀x ∈ R, and the fact that f is periodic with period 2π, we get
a+π a+2π
. f (x) cos(2nx) dx = − f (x) cos(2nx) dx (3.15)
a a+π
a+π a+2π
f (x) sin(2nx) dx = − f (x) sin(2nx) dx. (3.16)
a a+π
306 Series of Functions
Moreover, the result follows from the periodicity of the two trigonometric functions. Similarly, we can show that
a+π a+2π
. f (x) cos (2n − 1)x dx = f (x) cos (2n − 1)x dx (3.17)
a a+π
a+π a+2π
f (x) sin (2n − 1)x dx = f (x) sin (2n − 1)x dx. (3.18)
a a+π
Let us calculate the Fourier coefficients of the function f using equalities (3.15), (3.16), (3.17), and (3.18).
0 π π π
1 π 1 1
a0 = f (x) dx = f (x) dx + f (x) dx = − f (x) dx + f (x) dx = 0;
π −π π −π 0 π 0 0
0 π
1 π 1
an = f (x) cos(nx) dx = f (x) cos(nx) dx + f (x) cos(nx) dx
π −π π −π 0
⎧ π π
⎪ 1
⎪
⎪ − f (x) cos(nx) dx + f (x) cos(nx) dx = 0, if n is even
⎨π 0 0
= π π
⎪
⎪ 1 2 π
⎪
⎩ f (x) cos(nx) dx + f (x) cos(nx) dx = f (x) cos(nx) dx, if n is odd;
.
π 0 0 π 0
0 π
1 π 1
bn = f (x) sin(nx) dx = f (x) sin(nx) dx + f (x) sin(nx) dx
π −π π −π 0
⎧ π π
⎪ 1
⎪
⎪ − f (x) sin(nx) dx + f (x) sin(nx) dx = 0, if n is even
⎨π 0 0
= π π
⎪
⎪ 1 2 π
⎪
⎩ f (x) sin(nx) dx + f (x) sin(nx) dx = f (x) sin(nx) dx, if n is odd.
π 0 0 π 0
b) Since f˜ is an odd function, the Fourier series expansion is, in fact, a Fourier sine series. We can consider f˜
periodic with period 2l = 8. The coefficients of the Fourier sine series are
4 nπx 4 nπx
2 1
bn = f (x) sin dx = (x − 2) sin dx
4 0 4 2 2 4
4
1 4(x − 2) nπx 4 nπx 4
= − cos + cos dx
2 nπ 42 2 nπ 4
2
1 8 4 nπx 4
. = − cos(nπ) + sin
2 nπ nπ 4 2
1 8(−1)n 4 2 nπ
= − − sin
2 nπ nπ 2
4(−1)n 8 nπ
= − − 2
sin ,
nπ (nπ) 2
By Theorem 3.5.2, the series converges pointwise to the periodic function with period 8, illustrated in
Fig. 3.57.
1, if 0 ≤ x < π
.f (x) =
0, if π < x ≤ 2π,
we must consider its extension as an even function, as illustrated in Fig. 3.58. By Proposition 3.5.7, the Fourier
series of f is
a0 ∞ nx
. + an cos ,
2 n=1
2
π nx
2
where an = f (x) cos dx, n ∈ N0 , that is,
2π 0 2
2π
2 1 2π 1 π 1 π
a0 = f (x) dx = f (x) dx = 1 dx = x = 1;
2π 0 π 0 π 0 π 0
. nx nx nx π nπ
1 2π 1 π 1 2 2
an = f (x) cos dx = cos dx = sin = sin .
π 0 2 π 0 2 π n 2 0 nπ 2
a) If we want the Fourier sine series of the function, we must consider its extension as an odd function, as
illustrated in Fig. 3.59a.
∞
2 π
By Proposition 3.5.8, the Fourier series of f is bn sin(nx), where bn = f (x) sin(nx) dx, n ∈ N,
n=1
π 0
that is,
2 π 2 π
bn = f (x) sin(nx) dx = (π − x) sin(nx) dx
π 0 π 0
π
2 cos(nx) π cos(nx)
. = −(π − x) − dx
π n 0 0 n
π
2 π sin(nx) 2
= − = .
π n n2 0 n
3.6. Solved Exercises 309
(a) (b)
Figure 3.59: (a) The odd extension of f (x) = π − x. (b) The even extension of f (x) = π − x
By Theorem 3.5.1, the series converges pointwise, on the interval ]0, 2π[, to the function f˜(x) = π − x.
From this fact, it follows that
∞
sin(nx) π−x
. = , 0 < x < 2π.
n=1
n 2
π−x
b) The periodic function of period 2π defined on ]0, 2π[ by g(x) = is piecewise continuous. By Theo-
2
rem 3.5.7 we have
∞
x ∞ x
x π−t sin(nt) sin(nt)
. dt = dt = dt, 0 < x < 2 π,
0 2 0 n=1
n n=1 0
n
or,
∞ ∞ ∞
πx x2 cos(nx) 1 cos(nx) 1
. − = − 2
+ 2
=− 2
+ .
2 4 n=1
n n n=1
n n=1
n2
∞
2 4π 2 1
.2π = +4 ;
3 n=1
n2
therefore,
∞
1 π2
.
2
= .
n=1
n 6
Then
∞
cos(nx) π2 πx x2 3x2 − 6πx + 2π 2
.
2
= − + = .
n=1
n 6 2 4 12
310 Series of Functions
c) If we want the Fourier cosine series of the function, we must consider its extension as an even function, as
illustrated in Fig. 3.59b. By Proposition 3.5.7,
π π π π
1 2 2 2 x2
a0 = f (x) dx = f (x) dx = (π − x) dx = πx − = π;
π −π π 0 π 0 π 2 0
π
2 2 π
an = f (x) cos(nx) dx = (π − x) cos(nx) dx
π 0 π 0
. π π
2 (π − x) 1
= sin(nx) + sin(nx) dx
π n 0 0 n
π
2 cos(nx) 2
= − = 2 (1 − (−1)n ) .
nπ n 0 n π
∞ ∞
π 2 1 − (−1)n π 4 1
. + cos(nx) = + cos (2n − 1)x .
2 π n=1 n2 2 π n=1 (2n − 1)2
d) As the even extension of the function f , f˜(x) = π − |x|, is a continuous function and piecewise C 1 on
] − π, π[, we have
∞
π 2 1 − (−1)n
.π − |x| = + cos(nx), −π < x < π.
2 π n=1 n2
In particular, if x = 0,
∞
π 2 1 − (−1)n
.π = + .
2 π n=1 n2
∞ ∞
1 (−1)n
Since the series 2
and are convergent, Theorem 2.2.2 allows us to write
n=1
n n=1
n2
∞ ∞
π 2 1 2 (−1)n
.π = + − .
2 π n=1 n2 π n=1 n2
By item b),
∞
π 2 π2 2 (−1)n
.π = + · − .
2 π 6 π n=1 n2
Finally,
∞
(−1)n π2
.
2
=− .
n=1
n 12
a) If we want the Fourier sine series of the function, we must consider its extension as an odd function, as illus-
∞
trated in Fig. 3.60a. By Proposition 3.5.8, the Fourier series of f is bn sin(nx), where
n=1
3.6. Solved Exercises 311
π
2
bn = f (x) sin(nx) dx, n ∈ N, that is,
π 0
π π
2 2
bn = f (x) sin(nx) dx = x sin(nx) dx
π 0 π 0
π π
2 cos(nx) cos(nx)
. = −x + dx
π n 0 0 n
2 π(−1)n sin(nx) π 2(−1)n+1
= − + = .
π n n2 0 n
(a) (b)
Figure 3.60: (a) The odd extension of f (x) = x. (b) The even extension of f (x) = x
By Theorem 3.5.1, the series converges pointwise, on the interval ] − π, π[, to the function f˜(x) = x. From
this fact, it follows that
∞
sin(nx) x
. (−1)n+1 = , −π < x < π.
n=1
n 2
b) If we want the Fourier cosine series of the function, we must consider its extension as an even function, as
illustrated in Fig. 3.60b. By Proposition 3.5.7,
π π π π
1 2 2 2 x2
a0 = f (x) dx = f (x) dx = x dx = = π;
π −π π 0 π 0 π 2 0
π π
2 2
an = f (x) cos(nx) dx = x cos(nx) dx
π 0 π 0
.
π π 1
2x
= sin(nx) − sin(nx) dx
πn 0 0 n
π
2 cos(nx) 2
= = 2 (−1)n − 1 .
nπ n 0 n π
312 Series of Functions
52. The function f (x) = | sin(x)| is even (see Fig. 3.61); therefore, bn = 0, ∀n ∈ N. In addition, f is periodic with
period π, so it also has period 2π. Let us calculate the Fourier coefficients.
2 π 2 π 2" #π 4
a0 = f (x) dx = sin(x) dx = − cos(x) 0 = ;
π 0 π 0 π π
2 π 1 2 π
an = sin(x) cos(nx) dx = sin (n + 1)x + sin (1 − n)x dx
π 0 0 2 π
⎧ $ %π
⎪
⎪ 1 cos (n + 1)x cos (1 − n)x
⎪
⎪ − − , if n = 1
.
⎨ π n+1 1−n
= 0
⎪
⎪
⎪
⎪ 1 cos(2x) π
⎩ − , if n = 1
π 2 0
⎧
⎪ 1 (−1)n + 1 (−1)n + 1 2 (−1)n + 1
⎨ − =− · , if n = 1
= π n+1 n−1 π n2 − 1
⎪
⎩
0, if n = 1.
Since f is a continuous function and piecewise C 1 , Theorem 3.5.1 allows us to conclude that the series converges
pointwise to f on R.
Let x = 0. Then
∞
2 4 1
.f (0) = − .
π π n=1 4n2 − 1
π
Let x = . Then
2
π 2 4
∞
1 2
∞
4 (−1)n
.f = − cos(nπ) = − .
2 π π n=1 4n − 1
2 π π n=1 4n2 − 1
3.6. Solved Exercises 313
π
Since f = 1, we have
2
∞
2 4 (−1)n
. −1= ,
π π n=1 4n2 − 1
that is,
∞
(−1)n 1 π
.
2−1
= − .
n=1
4n 2 4
53. Since the function f (x) = x2 (π − x)2 is defined on ]0, π[ and we want to obtain a Fourier cosine series, we consider
an even extension of f to the interval ] − π, π[ (see Fig. 3.62). Since the function is even, the coefficients bn are
zero. Let us calculate a0 and an .
π π π
2 2 2
a0 = f (x) dx = x2 (π − x)2 dx = (x4 − 2π x3 + π 2 x2 ) dx
π 0 π 0 π 0
π
2 x5 π x4 π 2 x3 π4
= − + = ;
π 5 2 3 0 15
π π
2 2
an = x2 (π − x)2 cos(nx) dx = (x4 − 2π x3 + π 2 x2 ) cos(nx) dx
π 0 π 0
π π
2 sin(nx) sin(nx)
= (x4 − 2π x3 + π 2 x2 ) − (4x3 − 6π x2 + 2π 2 x) dx
π n 0 0 n
. π
2 cos(nx) π cos(nx)
= (4x3 − 6π x2 + 2π 2 x) − (12x2 − 12π x + 2π 2 ) dx
πn n 0 0 n
π π
2 2 sin(nx) sin(nx)
= − (12x 2
− 12π x + 2π ) + (24x − 12π) dx
π n2 n 0 0 n
π
2 cos(nx) π cos(nx)
= − (24x − 12π) + 24 dx
π n3 n 0 0 n
π
2 sin(nx) 24
= −12π cos(nπ) − 12π + 24 = − 4 ((−1)n + 1)
π n4 n 0 n
314 Series of Functions
∞ ∞
π4 (−1)n + 1 π4 1
. − 24 4
cos(nx) = −3 4
cos(2nx).
30 n=2
n 30 n=1
n
The extension, f˜, of f is a continuous function on [−π, π], and Theorem 3.5.1 allows us to conclude that the
series converges pointwise to f˜ on [−π, π], that is,
∞
π4 1
.f˜(x) = −3 cos(2nx), ∀x ∈ [−π, π].
30 n 4
n=1
x2 (π + x)2 , if − π ≤ x ≤ 0
.f˜(x) =
x2 (π − x)2 , if 0 ≤ x ≤ π.
Let x = 0. Then
∞
π4 1
.f˜(0) = −3 .
30 n 4
n=1
π
Let x = . Then
2
π π4 ∞
1 π4 ∞
(−1)n
.f˜ = −3 cos(nπ) = −3 .
2 30 n 4 30 n4
n=1 n=1
π π4
Since f˜ = , we have
2 16
π π4 π4 ∞
(−1)n π4 ∞
1 ∞
1 − (−1)n
.f˜ − f˜(0) = = −3 − + 3 = 3 ,
2 16 30 n 4 30 n 4 n4
n=1 n=1 n=1
∞ ∞
π4 1 π4 1
that is, =6 , which is equivalent to = .
16 n=1
(2n − 1) 4 96 n=0
(2n + 1)4
∞
54. a) The series n an sin(nx) results from term by term differentiation of the original series. In fact, according
n=1
to Theorem 3.5.8, by term by term differentiation of the series of f˜, we obtain the Fourier series of f˜ , that
is,
∞
∞ ∞
π 2 (−1)n − 1 2 (−1)n − 1 2 (−1)n − 1
. + cos(nx) = cos(nx) = − n sin(nx).
2 π n=1 n2 π n=1 n 2 π n=1 n2
3.6. Solved Exercises 315
By the same theorem, this series converges to f˜ at the points of continuity of f˜ . At the points of
discontinuity the sum of the series is equal to the arithmetic mean of the left- and right-hand derivatives of
the function f˜ at these points. Thus,
∞
2 (−1)n − 1 f˜ (x), if x = kπ, k ∈ Z
. − n sin(nx) =
π n=1 n2 0, if x = kπ, k ∈ Z.
∞
Therefore, the series n an sin(nx) converges pointwise on R.
n=1
∞
b) Let g(x) = − n an sin(nx). By the previous item,
n=1
. ⎧
⎪
⎪ 1, if x ∈ ]2kπ, (2k + 1)π[, k ∈ Z
⎨
= −1, if x ∈ ](2k + 1)π, (2k + 2)π[, k ∈ Z
⎪
⎪
⎩
0, if x = kπ, k ∈ Z.
that is,
∞
(−1)n π3
. =− .
n=1
(2n − 1) 3 32
∞
55. Suppose that the function f has a Fourier sine series development, that is, f (x) = Bn sin(nπx) (note that,
n=1
in this case, the two conditions f (0) = f (1) = 0 are satisfied).
Assuming that f and its derivatives meet the conditions of Theorem 3.5.8, we have
∞
∞
∞
∞
.f (x) + 3f (x) = Bn sin(nπx) +3 Bn sin(nπx) = − n2 π 2 Bn sin(nπx) + 3 Bn sin(nπx).
n=1 n=1 n=1 n=1
Let g(x) = 3x, 0 < x < 1. By extending g to an odd function, we obtain the Fourier sine series of g on the
interval ] − 1, 1[
∞
.g(x) ∼ bn sin(nπx),
n=1
where
1
1 3x 1 3 6 (−1)n
.bn =2 3x sin(nπx) dx = 2 − cos(nπx) + cos(nπx) dx =− .
0 nπ 0 0 nπ nπ
∞
6 (−1)n+1
.3x = sin(nπx), ∀x ∈ ] − 1, 1[ .
n=1
nπ
Thus,
∞
∞
∞
6 (−1)n+1
f (x) + 3f (x) = 3x ⇔ − n2 π 2 Bn sin(nπx) + 3 Bn sin(nπx) = sin(nπx)
n=1 n=1 n=1
nπ
.
∞
∞
6 (−1)n+1
⇔ −n2 π 2 Bn + 3Bn sin(nπx) = sin(nπx).
n=1 n=1
nπ
3.6. Solved Exercises 317
6 (−1)n+1
. − n2 π 2 Bn + 3Bn = , ∀n ∈ N,
nπ
that is,
6 (−1)n+1
.Bn = , ∀n ∈ N.
nπ(3 − n2 π 2 )
∞
6 (−1)n+1
The desired function is f (x) = sin(nπx).
n=1
nπ(3 − n2 π 2 )
2 2 2
x2 2x
bn = x2 sin(nπx) dx = − cos(nπx) − − cos(nπx) dx
0 nπ 0 0 nπ
2 2 2 sin(nπx)
4 2 4 2 x
= − + x cos(nπx) dx = − + sin(nπx) − dx
nπ nπ 0 nπ nπ nπ 0 0 nπ
4 2 cos(nπx) 2 4
= − + =− .
nπ (nπ)2 nπ 0 nπ
b) Term by term differentiation of the series obtained in item a) results in the series
∞
∞
4 4 4
. − n π sin(nπx) − n π cos(nπx) = − sin(nπx) − 4 cos(nπx) ,
n=1
(nπ)2 nπ n=1
nπ
which is divergent because if x ∈ ]0, 2[, the general term of the series does not have a limit. Therefore, the
sum of the series does not exist and cannot define the function f .
318 Series of Functions
Note that the function f does not verify the conditions of Theorem 3.5.8. In fact, the periodic extension of
f to the interval ] − 1, 1[
(x + 2)2 , if − 1 < x ≤ 0,
.g(x) =
x2 , if 0 < x < 1
is not continuous at 0.
3 3 3
1 1 1 x2 x3
a0 = f (x) dx = (x − x2 ) dx = − = −6;
3 −3 3 −3 3 2 3 −3
1
3 nπx
3(x − x2 ) nπx 3 1
3
3(1 − 2x) nπx
an = (x − x2 ) cos sin dx = − sin dx
3 −3 3 nπ 3 3 −3 −3 nπ 3
3
1 nπx 1 3(2x − 1) nπx 3 3 6 nπx
= (2x − 1) sin dx = − cos + cos dx
nπ −3 3 nπ nπ 3 −3 −3 nπ 3
n n
1 36(−1) 18 nπx 3 36(−1)
. = − + sin =− ;
nπ nπ (nπ)2 3 −3 (nπ)2
1 3 nπx 1 3(x2 − x) nπx 3 3
3(1 − 2x) nπx
bn = (x − x2 ) sin dx = cos + cos dx
3 −3 3 3 nπ 3 −3 −3 nπ 3
1 3(1 − 2x) nπx 3 3 6 nπx
= − 6(−1)n − sin − sin dx
nπ nπ 3 −3 −3 nπ 3
nπx 3
1 18 6(−1)n
= − 6(−1)n + cos =− .
nπ (nπ)2 3 −3 nπ
∞ nπx nπx
36(−1)n 6(−1)n
. −3+ − 2
cos − sin
n=1
(nπ) 3 nπ 3
and, on the interval ] − 3, 3[, its sum is equal to f because f is continuous and of class C 1 .
b) Since
x nπt 3 nπt x 3 nπx
. sin dt = − cos = 1 − cos
0 3 nπ 3 0 nπ 3
and
nπx
x nπt 3 nπt x 3
. cos dt = sin = sin ,
0 3 nπ 3 0 nπ 3
the series
3 nπx nπx
∞ ∞
a0 x 3
. + bn + an sin − bn cos ,
2 n=1
nπ n=1
nπ 3 3
3.6. Solved Exercises 319
where a0 , an , and bn are the Fourier coefficients of f , results from the term by term integration between
0 and x of the series of f . By Theorem 3.5.7, with a = 0 and b = x, the sum of the series is equal to
x
f (t) dt, that is,
0
∞ x x 2 x
3 t t3 x2 x3
.S(x) + bn = f (t) dt = (t − t2 ) dt = − = − , x ∈ ] − 3, 3[.
n=1
nπ 0 0 2 3 0 2 3
∞
x2 x3 18(−1)n x2 x3 3
.S(x) = − + 2 2
= − − , x ∈ ] − 3, 3[.
2 3 n=1
n π 2 3 2
320 Series of Functions
∞
∞
(−1)n
a) n n xn r) √
3
(x − 2)n
n=1 n=0 4n n2 + 1
∞
n ∞
3n
1 2n x s) (x − 2)n
b) 1+ · 1 + 9n
n=1
n x+1 n=0
∞
∞ (−1)n
(−1)n 1 − x n t) (x − 2)n
c) 2n (n2
+ 3)
n=1
2n − 1 1 + x n=0
b) Using the series of derivatives from the result 23. Obtain the Maclaurin series of the function
of the previous item, calculate the sum of the
series .f (x) = (1 + x)−2
∞
n+1
. .
n=0
n! by two different processes. What is the radius of
convergence of the series?
19. Consider the function f : R → R defined by
24. Determine two power series that represent the func-
.f (x) = x sin(2x). tion
1
.f (x) =
2 − x
1 3
a) Write Maclaurin series of f and indicate on the interval , . Justify the answer.
2 2
for which values of x the development
is valid. 25. Develop the function
b) Using the result of the previous item and by
term by term differentiation, compute the sum .f (x) = log(x)
of the series
∞ into a power series of x − 2 and indicate an open in-
(2n + 2)π 2n+1
. (−1)n . terval on which the function coincides with the sum
n=0
(2n + 1)! of the series.
Justify the answer.
26. Expand the function
20. Consider the function f : R → R defined by
.f (x) = e−x+4 + 3x
.f (x) = sin2 (x) + log(1 − x2 ).
into a power series of x − 1 and indicate the largest
open interval on which the development represents
a) Using term by term integration, develop f
the function.
into a Maclaurin series and indicate the
real values of x for which the development
27. Let f be the function defined by f (x) = x2 log(x2 )
is valid.
on R \ {0}. Develop f into a power series of x − 1
Note: 2 sin(x) cos(x) = sin(2x).
and indicate the largest open interval where this ex-
b) Using the result from the previous item, indi- pansion represents the function.
cate the value of f (15) (0) and f (16) (0).
28. Determine whether the following functions are even
21. Develop the function
or odd and sketch their graphs:
2
.f (x) = a) x2 − x g) x|x| + 3
4x + 5
into a power series of x + 3 and determine the radius b) x2 sin(3x) h) sin2 (x)
of convergence of the series.
i) e−|x|
c) tan(6x)
22. Develop the function
2
x3 e−x
1 d) x4 + 20 cos(5x) j)
.f (x) = cos(x)
x2 − 6x + 5
e) x5 − 2x k) x + 5 cos(x)
into a power series of x−3 and determine the interval
of convergence of the series.
f) sin(x) + cos(2x) l) log(x4 )
3.7. Proposed Exercises 323
⎧ π
29. Sketch the graph of the following functions, check if ⎨0, <x≤0
if −
d) 2
they are periodic, and, if so, indicate a period:
x ⎩π, if 0 < x < 3π
2
a) sin(7x) e) sin + cos(2x) ⎧ 3π
3 ⎨0, if − <x≤0
b) sinh(2x) 2
f) sin (x) e) 2
x ⎩sin(x), if 0 < x < π
c) tan(π x) g) sin cos(x) 2
π x π
d) sin h) x + cos(8x) 33. The following functions are periodic and defined on
4
the interval corresponding to one period. Determine
0, if 2n − 1 ≤ x < 2n, n ∈ Z
i) the Fourier series and the respective sum for each
1, if 2n ≤ x < 2n + 1, n ∈ Z
function.
30. Determine the Fourier series related to the func- a) x, 0 < x < 3
tions periodic with period 2π, defined on the interval
] − π, π[ as follows: b) 2 − x, 0 < x < 2
⎧
⎧ ⎪
⎨−1, if −π <x≤0 ⎨1, if − 2 ≤ x ≤ −1
⎪
a) c) 0, if − 1 < x < 1
⎩1, ⎪
⎪
if 0 < x < π ⎩1, if 1 ≤ x ≤ 2
2
1, if − π < x ≤ 0 x + 1, if − 1 ≤ x ≤ 0
b) d)
3, if 0 < x < π 1 − x, if 0 < x ≤ 1
x2 , if − π < x ≤ 0
c) 0, if − 3 ≤ x ≤ 0
−x2 , if 0 < x < π e)
x2 (3 − x), if 0 < x ≤ 3
d) x + |x|
34. Let L > 0. Develop the following functions into
e) e−|x| Fourier series on the indicated intervals:
f) cos2 (2x) a) |x|, ] − L, L[
−2x, if 0 < x ≤ π 36. Determine the Fourier cosine series and its respective
c)
2x, if π < x < 2π sum for each function.
324 Series of Functions
46. Using the equality 47. Determine, using Fourier series, the function f such
that
∞
(−1)n+1 ⎧
.x =2 sin(nx), −π < x < π, ⎪
n ⎨f (x) + 10f (x) = 5x, −2 < x < 2
n=1
.
⎪
⎩f (−2) = f (2) = 0.
obtain the Fourier series development of f (x) = x4 .
Answers to
Proposed Exercises
Chapter 1
1
3. a) .− 15 c) 1 e) .− 16. a) True b) False c) False
4
b) .+∞ d) 0
17. a) .lim un = +∞; .lim un = 0
4. a) 0 c) 0 e) 1
b) .lim un = 1; .lim un = −1
1
b) .−1 d) . c) .lim un = +∞; .lim un = −1
2
√
5. c) .nn , .n!, .en , .2n , .n3 , .2 n, . 10 n, .log(n) d) .lim un = 1; .lim un = −1
1 1 1 1 1 1 1 1 e) .lim un = +∞; .lim un = 0
6. . ,. ,. ,. ,. ,. , .√ ,.
nn n! en 2n n3 2 n 10 n log(n)
f) .lim un = 1; .lim un = 0
7. a) .e4 c) .+∞ g) .lim un = +∞; .lim un = −∞
b) 0 d) 1 π 3π
h) If .a ∈ + 2kπ, + 2kπ , .k ∈ Z,
1 4 4 4
8. a) . b) . .lim un = sin(a); .lim un = − sin(a)
2e e
5π 7π
If .a ∈ + 2kπ, + 2kπ , .k ∈ Z,
1 4 4
9. .p = .lim un = − sin(a); .lim un = sin(a)
3e π π
If .a ∈ − + 2kπ, + 2kπ , .k ∈ Z,
10. a) 1 if .x = 2kπ, .k ∈ Z 4 4
limit does not exist if .x = (2k + 1)π, .k ∈ Z .lim un = cos(a); .lim un = − cos(a)
0 if .x ∈ R \ {kπ, k ∈ Z} 3π 5π
If .a ∈ + 2kπ, + 2kπ , .k ∈ Z,
4 4
b) 0 e) 2 h) .+∞
.lim un = − cos(a); .lim un = cos(a)
c) .+∞ 1
f) . i) 0
2 e i) .lim un = +∞; .lim un = 0
d) . g) .+∞ j) 1
e j) .lim un = +∞; .lim un = −∞
13. .−1
n
√ 8
15. c) . 2 22. a) .
9
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 327
A. Alves de Sá, B. Louro, Sequences and Series,
https://doi.org/10.1007/978-3-031-67202-6
328 Answers to Proposed Exercises
Chapter 2
1 1 3
1. a) . = ;. h) Convergent or divergent, depending on the se-
(n + 1)2 − 1 n(n + 2) 4
ries
1 1
b) . ;.
n(n + 1)(n + 2) 4 9. a) Convergent e) Divergent
1 1
c) . ;. b) Divergent f) Convergent
n(n + 1)(n + 2)(n + 3) 18
98
(−1)n b) Convergent d) Divergent
7. b) .S98 =
n=1
(n + 1)2 14. a) Divergent
Chapter 3
x +x3 −1
1. Convergent on .] − ∞, 0[, .S = e 2 1 − ex
3
, e) Absolutely convergent on .] − 5, −1[
divergent on .[0, +∞[ Conditionally convergent at .x = −5
Divergent on .] − ∞, −5[ ∪ [−1, +∞[
1
2. Convergent on . , 1 , f) Absolutely convergent on .[−2, 8]
3 Divergent on .] − ∞, −2[ ∪ ]8, +∞[
1 √ √
divergent on . − ∞, ∪ [1, +∞[ g) Absolutely convergent on .] − 2, 2[\{0}
3 √
Conditionally convergent at .x = − 2 and at
√
4. Absolutely convergent on .] − ∞, −1[ ∪ ]1, +∞[ and .x = 2
√ √
divergent on .] − 1, 1[, .∀α ∈ R Divergent on .] − ∞, − 2[ ∪ ] 2, +∞[ ∪ {0}
if .x = 1: absolutely convergent if .α < −1 and diver-
1 1
gent if .α ≥ −1 h) Absolutely convergent on . 2− , 2+
π π
if .x = −1: absolutely convergent if .α < −1, condi-
1 1
tionally convergent if .−1 ≤ α < 0 and divergent if Divergent on . −∞, 2 − ∪ 2 + , +∞
π π
.α ≥ 0
5 1
i) Absolutely convergent on . − , −
6. a) Convergent b) Is continuous 6 6
5 1
Divergent on . −∞, − ∪ − , +∞
6 6
7. a) .]1, +∞[ √ √
j) Absolutely convergent on .] − 2 − 1, 2 − 1[
√ √
∞ Divergent on .] − ∞, − 2 − 1] ∪ [ 2 − 1, +∞[
1 − cos(n)
9. b) .
n=1
n3 k) Absolutely convergent on .] − 1, 3[
Divergent on .] − ∞, −1] ∪ [3, +∞[
10. a) Converges uniformly
l) Absolutely convergent on .]1, 7[
∞ √ √
2 4n5 + 1 − 2n5 + 1 Conditionally convergent at .x = 7
b) .
n4 Divergent on .] − ∞, 1] ∪ ]7, +∞[
n=1
m) Absolutely convergent on
∞
1 n+1 .] − ∞, −1[ ∪ ]3, +∞[
11. b) . − log
n=1
n n Conditionally convergent at .x = −1
Divergent on .] − 1, 3] \ {1}
1
∞ sin
1 5
13. a) .R b) .
n2 n) Absolutely convergent on . ,
1 + (nx)2 3 3
n=1
1 5
Divergent on . −∞, ∪ , +∞
14. a) Divergent on .R \ {0} 3 3
o) Absolutely convergent on .] − 2, 0[
1
b) Absolutely convergent on . − , +∞ Conditionally convergent at .x = −2
2
1 Divergent on .] − ∞, −2[ ∪ [0, +∞[
Divergent on . −∞, − \ {−1}
2 p) Absolutely convergent on .[−4, 0]
c) Absolutely convergent on .]0, +∞[ Divergent on .] − ∞, −4[ ∪ ]0, +∞[
Conditionally convergent at .x = 0 q) Absolutely convergent on .]0, 2e[
Divergent on .] − ∞, 0[ \{−1} Divergent on .] − ∞, 0] ∪ [2e, +∞[
d) Absolutely convergent on .] − 2, 2[ r) Absolutely convergent on .] − 2, 6[
Conditionally convergent at .x = 2 Conditionally convergent at .x = 6
Divergent on .] − ∞, −2] ∪ ]2, +∞[ Divergent on .] − ∞, −2] ∪ ]6, +∞[
Answers to Proposed Exercises 331
∞
s) Absolutely convergent on .] − 1, 5[ n−3 n
o) . x , .|x| < 5
Divergent on .] − ∞, −1] ∪ [5, +∞[ n=0
5n+1
t) Absolutely convergent on .[0, 4]
∞
Divergent on .] − ∞, 0[ ∪ ]4, +∞[ (−1)n 3n+4
17. . x , .|x| < 1
n=0
n+1
15. a) Absolutely convergent on .R
b) Absolutely convergent on ∞
xn+1
.] − a, a[, if .a ≥ b
18. a) . , .∀x ∈ R
n=0
n!
.] − b, b[, if .b > a
b) .2 e
∞ n
log(a) ∞
16. a) . xn , for all .x ∈ R (−1)n 22n+1 2n+2
n! 19. a) . x , .∀x ∈ R
n=0 (2n + 1)!
n=0
∞
x2n+1
b) . , for all .x ∈ R b) .−π
n=0
n!
∞ ∞
(−1)n 2n (−1)n 22n+1 1
c) . x , .|x| < |a| 20. a) . − x2n+2 , .|x| < 1
n=0
a2n+2 n=0
(2n + 2)! n+1
∞
(−1)n 2n+1 b) .f (15) (0) = 0; .f (16) (0) = −215 − 2 × 15!
d) . x , .|x| < 1
n=0
2n +1
∞
∞
−22n+1
1 21. . (x + 3)n , .− 19 < x < − 45 , .r = 7
e) . x2n+1 , for all .x ∈ R 7n+1 4 4
n=0
(2n + 1)! n=0
∞
1 ∞
f) x2n , for all .x ∈ R −1
.
(2n)! 22. . (x − 3)2n , .1 < x < 5
n=0
n=0
22n+2
∞ n
log(2) (−1)n
g) . + n+1 xn , .|x| < 2 ∞
n! 2
n=0 23. . (−1)n (n + 1) xn , .|x| < 1, .r = 1
∞
(−1)n n=0
h) . x2n+1 , for all .x ∈ R
n!(2n + 1) ∞
n=0
∞
24. . (x − 1)n , .0 < x < 2
1 (−1)n+1
i) . − 1 xn , .|x| < 1 n=0
n=0
3 2n+1
∞
∞ 1
3n n . xn , .|x| < 2
j) . x , for all .x ∈ R n=0
2n+1
n=1
n! n
∞
∞
1 (−1)n
k) . 1 − n+1 xn , .|x| < 1 25. .log(2) + (x − 2)n+1 , .0 < x < 4
n=0
3
n=0
(n + 1)2n+1
∞
l) . (2n + 1) x , .|x| < 1
n
∞
e3 (−1)n + 3 logn (3)
n=0 26. . (x − 1)n , .∀x ∈ R
∞
n!
(−1)n+1
n=0
m) . x2n−1 , .∀x ∈ R
(2n)!(2n − 1) ∞
n=1 (−1)n (x − 1)n+3
∞ 27. .2(x − 1) + 3(x − 1)2 + 4 ,
1 (n + 1)(n + 2)(n + 3)
n) . 1 − n+1 xn , .|x| < 1 n=0
2 .0 <x<2
n=0
332 Answers to Proposed Exercises
b) Odd
c) Odd
h) Even
d) Even
i) Even
e) Odd
Answers to Proposed Exercises 333
l) Even
e) Periodic with period .6π
g) Not periodic
b) Not periodic
334 Answers to Proposed Exercises
∞
h) Not periodic cos (2n − 1)x sin (2n − 1)x
c) .π + 8 −
n=1
(2n − 1)2 π 2n − 1
⎧
⎪
⎪ −2(x − 2kπ),if x ∈ ]2kπ, (2k + 1)π[
⎪
⎪
⎨0, if x = (2k + 1)π
.
⎪2(x − 2kπ), if
⎪ x ∈ ](2k + 1)π, (2k + 2)π[
⎪
⎪
⎩
2π, if x = 2kπ
.k ∈Z
∞
3π sin 3nπ
2
i) Periodic with period 2 d) . + cos(nx)
4 n=1
n
3nπ
2 sin2 4
.+ sin(nx)
n
⎧
⎪
⎪ if x ∈ ](4k − 1) π2 , 2kπ[
⎨0,
. π, if x = (4k − 1) π2 ∨ x = 2kπ
⎪⎪2
⎩π, if x ∈ ]2kπ, (4k + 3) π2 [
∞
1 3 1
30. a) .− + sin (2n − 1)x .k∈Z
4 π n=1 2n − 1
1 1 1
∞ e) . + cos(x) + sin(x)
4 1 2π 2π 4
b) .2 + sin (2n − 1)x
π n=1 2n − 1 ∞
1 n sin 2 − 1
nπ
∞ .+ cos(nx)
2 2 + (−1)n (π 2 n2 − 2) π n=2 n2 − 1
c) . sin(nx)
π n=1 n3
n cos nπ
.−
2
∞ sin(nx)
π (−1)n − 1 (−1)n n2 − 1
d) . +2 2
cos(nx)− sin(nx) ⎧
2 πn n
⎪
⎨sin(x), if x ∈ [2kπ, (4k + 1) 2 [
n=1 π
⎪
∞
1 − e−π 2 1 − (−1)n e−π . 1, if x = (4k + 1) 2π
e) . + cos(nx) ⎪
⎪ 2
π π n=1 n2 + 1 ⎩0, if x ∈ ](4k + 1) π , (2k + 2)π[
2
1 1 .k ∈Z
f) . + cos(4x)
2 2
∞
3 3 1 2nπx
33. a) . − sin
∞ 2 π n=1 n 3
2 4 (−1)n
32. a) . − cos(2nx)
π π n=1 4n2 − 1 x − 3k, if x ∈]3k, 3k + 3[
.
3
.| cos(x)|, .∀x
∈R 2
, if x = 3k
∞
2π
n2 π 2 − 2 .k ∈Z
b) .3π 3 +18 cos(nx)+ sin(nx) ∞
n=1
n2 n3 2 1
b) .1 + sin(nπx)
⎧ π n=1 n
⎪
⎪3(x − 2kπ) π 2 − (x − 2kπ)2 ,
⎨
. if x ∈ ](2k − 2)π, 2kπ[ 2 − (x − 2k), if x ∈ ]2k, 2k + 2[
⎪
⎪
⎩9π 3 , if x = 2kπ
.
1, if x = 2k
.k ∈Z .k ∈Z
Answers to Proposed Exercises 335
∞ ⎧
2 (−1)n (2n − 1)πx ⎪
⎪−2, if x ∈ ] − 3π + 6kπ, −2π + 6kπ[
1 ⎪
c) . + cos ⎪
2 π n=1 2n − 1 2 ⎪
⎪
⎪
⎪ −1, if x ∈ ] − 2π + 6kπ, −π + 6kπ[
⎧ ⎪
⎪
⎪
⎪ if x ∈ ] − π + 6kπ, π + 6kπ[
⎪ ⎪
⎪ 0,
⎨1, if x ∈ ]1 + 4k, 3 + 4k[
⎪ ⎪
⎪
⎪
⎪
⎪ 1, if x ∈ ]π + 6kπ, 2π + 6kπ[
. 0, if x ∈ ]3 + 4k, 5 + 4k[ ⎪
⎪
⎪
⎪ ⎨2,
⎩ 1 , if x = 2k + 1 if x ∈ ]2π + 6kπ, 3π + 6kπ[
2 .
⎪1,
⎪ if x = π + 6kπ
⎪
⎪ 2
.k ∈Z ⎪
⎪− 1 ,
⎪
⎪ if x = −π + 6kπ
∞ ⎪
⎪ 2
1 4 1 ⎪3
⎪
d) . + 2 cos (2n − 1)πx ⎪
⎪ , if x = 2π + 6kπ
2 π n=1 (2n − 1) 2 ⎪
⎪
2
⎪
⎪ − 23 , if x = −2π + 6kπ
⎪
⎪
⎪
⎩
x − 2k + 1, if x ∈ [2k − 1, 2k] 0, if x = 3π + 6kπ
.
−x + 2k + 1, if x ∈ [2k, 2k + 1] .k ∈Z
∞
.k ∈Z 4 (−1)n + 1 4((−1)n −1) nπx
c) . − sin
∞ π n=1 n n3 π 2 2
9 27 6 + (−1)n (n2 π 2 − 6) nπx
e) .− 4 cos ⎧
8 π n=1 n4 3 ⎪
⎪2 − (x − 4k)2 , if x ∈ ]4k, 4k + 2[
⎨
2 π(1 + 2(−1)n ) nπx . (x − 4k)2 − 2, if x ∈ ]4k − 2, 4k[
.+ sin ⎪
⎪
n3 3 ⎩0, if x = 2k
(x − 6k)2 (3 + 6k − x),if x ∈ [6k, 3 + 6k] .k ∈Z
.
∞
0, if x ∈ [3 + 6k, 6 + 6k] 8 (−1)n+1 (2n − 1)πx
d) . 2 sin
.k ∈Z π n=1 (2n − 1)2 2
∞ x − 4k, if x ∈ [−1 + 4k, 1 + 4k]
L 4L 1 (2n − 1)πx
34. a) . − 2 cos .
2 π n=1 (2n − 1)2 L 2 − (x − 4k), if x ∈ [1 + 4k, 3 + 4k]
∞
.k ∈Z
1 L(−1)n nπx
∞
b) .sinh(L)
L
+2
L 2 + n2 π 2
cos
L 2 6 − n2 π 2
n=1 e) . 3 (−1)n sin(nπx)
n π n=1 n3
nπ(−1) nπx
.− sin
L2 + n2 π 2 L (x − 2k)3 , if x ∈ ]2k − 1, 2k + 1[
.
∞
1 2L 1 nx 0, if x = 2k + 1
c) .sinh(πL) + cos
πL π L2 + n2 L
n=1 .k ∈Z
∞
4
∞
1 2 π 4 4n2 + 1
35. a) sin (2n − 1)x 36. a) .− + cos(x) − cos(2nx)
.
π n=1 2n − 1 π 2 π n=1 (4n2 − 1)2
⎧
⎪ if x ∈ ]2kπ, (2k + 1)π[ −(x − 2kπ) cos(x), if x ∈ [(2k − 1)π, 2kπ]
⎪
⎨1, .
. −1, if x ∈ ](2k + 1)π, (2k + 2)π[ (x − 2kπ) cos(x), if x ∈ [2kπ, (2k + 1)π]
⎪
⎪
⎩0, if x = kπ .k ∈Z
∞
.k ∈Z 4 16 1 + 3(−1)n nπx
b) . + 2 cos
∞ 3 π n=1 n2 4
4 cos − (−1)n
nπ nπ
2
cos 6 nx
b) . sin
π n=1 n 3 .(x − 8k)2 − 2|x − 8k|, if .x ∈ [8k − 4, 8k + 4]
.k ∈ Z
336 Answers to Proposed Exercises
∞ ∞
1 4 1 + (−1)n − 2 cos 4
nπ
nπx 2 1
c) . − 2 2
cos 41. a) . + cos(2nx)
2 π n=1 n2 2 π π n=1 1 − 4n2
∞
1
∞
4 1 8 n
.= − 2 cos (2n − 1)πx b) . sin(2nx)
2 π n=1 (2n − 1)2 π n=1 4n2 − 1
c) No
x − 2k, if x ∈ [2k, 2k + 1]
.
2 − x + 2k, if x ∈ [2k + 1, 2k + 2] ∞
3π 1 − (−1)n (−1)n
42. a) . + 2
cos(nx)+ sin(nx)
.k ∈Z 4 n=1 πn n
∞
h 2 sin(nh)
37. . + cos(nx)
π π n=1 n
∞
h 2 sin2 (nh)
38. . + cos(nx)
π πh n=1 n2
43. a) .S(612345 ) = 0, .S(65) = 1
∞
sin(2nx) ∞
39. a) . 2 1 2 nπ nπx
2n b) . + 2 sin sin
n=1 π n=1 n n π 2 2
∞
2 cos (2n − 1)x
b) . ∞
π n=1 (2n − 1)2 1 − (−1)n (−1)n
44. a) . cos(nx)+ sin(nx)
∞ πn2 n
4 (−1)n+1 n=1
40. a) . cos (2n − 1)πx
π n=1 2n − 1 ⎧
⎪
⎪ π π2
⎧ ⎨ x+ , if − π < x < 0
⎪
⎪ 1, if x ∈ ] − 1
+ 2k, 1
+ 2k[ b) .S(x) = 4 12
⎪
⎨ 2 2 ⎪
⎪ 2 π2
⎩πx − x
+ , if 0 ≤ x < π
. −1, if x ∈ ] 12 + 2k, 3
+ 2k[ 4 2 12
⎪
⎪ 2
⎪
⎩
0, if x = 2k+1 ∞
2 4 (2n − 1)πx
45. .f ∼ sin
.k ∈Z n=1
(2n − 1)π 3
∞
3 12 (2n − 1)πx
.F (x) =− − cos
b) 2 n=1 (2n − 1) π
2 2 3
∞
π4 (−1)n 2 2
46. .x4 = +8 (π n − 6) cos(nx)
5 n=1
n4
∞
80 (−1)n nπx
47. . sin
π n=1 n (nπ)2 − 40 2
Bibliography
1. Anton, H., Bivens, I., Davis, S.: Calculus, 12th edn. Wiley, Hoboken, New Jersey (2022)
3. Bourchtein, L., Bourchtein, A.: Theory of Infinite Sequences and Series. Birkhäuser, Cham, Switzer-
land (2022)
4. Ellis, R., Gulick, D.: Calculus with Analytic Geometry, 5th edn. Saunders College Publishing,
Harcourt Brace College Publishers, Fort Worth (1994)
5. Hunt, R.A.: Calculus, 2nd edn. HarperCollins College Publishers, New York (1994)
6. Kreyszig, E.: Advanced Engineering Mathematics, 10th edn. Wiley, New York (2011)
7. Laczkovich, M., Sós, V.T.: Real Analysis, Series, Functions of Several Variables, and Applications.
Springer, New York (2017)
8. Larson, R., Hostetler, R., Edwards, B.H.: Calculus with Analytic Geometry, 7th edn. Houghton
Mifflin, Boston (2001)
9. Little, C.H.C., Teo, K.L., Van Brunt, B.: Real Analysis via Sequences and Series. Springer, New
York (2015)
10. Pinkus, A., Zafrany, S.: Fourier Series and Integral Transforms. Cambridge University Press, Cam-
bridge (1997)
11. Mcquarrie, D.A.: Mathematical Methods for Scientists and Engineers, University Science Books,
Sausalito, California (2003)
12. Salas, S., Hille, E., Etgen, G.: Calculus, One and Several Variables, 10th edn. Wiley, Hoboken,
New Jersey (2007)
13. Spivak, M.: Calculus, 4th edn. Publish or Perish, Inc, Houston, Texas (2008)
14. Stewart, J.: Calculus, 9th edn. Cengage Learning, Boston, MA (2021)
15. Stewart, J.: Calculus, Concepts and Contexts, 5th edn. Cengage Learning, Boston, MA (2023)
16. Taylor, A., Mann, R.: Advanced Calculus, 3rd edn. Wiley, New York (1983)
17. Tolstov, G.P.: Fourier Series. Dover Publications, New York (1962)
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 337
A. Alves de Sá, B. Louro, Sequences and Series,
https://doi.org/10.1007/978-3-031-67202-6
Index
Symbols E
.Int(x), integer part of x, viii Euler’s constant, 30
.R, extended real line, 15 Euler, Leonhard, 29
extended real line, 15
A
Alternating series, 82 F
arithmetic progression, see arithmetic sequence Fibonacci
arithmetic sequence, 6 sequence, 2
common difference, 6 Fibonacci, Leonardo, 2
definition, 6 Fourier series
general term, 7 coefficients, 205, 223
sum of the first n terms, 7 convergence, 225
cosine series, 216
definition, 205
B
differentiation term by term, 226
bounded, viii
even function, ix
bounded from above, viii
integration term by term, 226
bounded from below, viii
odd function, ix
period, 201
C periodic extension, 206
Cauchy periodic function, 201
Cauchy product, 113 piecewise continuous, 211
Cauchy’s Criterion, 78 piecewise .C 1 , 212
Cauchy’s Root Test, 106 pointwise convergence, 212
Condensation Test, 90 sine series, 217
sequence, 35 trigonometric polynomial, 202
Cauchy, Augustin Louis, 36 trigonometric series, 202
coefficients, 202
D uniform convergence, 225
D’Alembert’s Test, 102 Fourier, Joseph, 205
D’Alembert, Jean Le Rond, 102 function
difference of sequences, 4 even, ix, 214
Dirichlet’s Test, 83 odd, ix, 214
Dirichlet, Peter Gustave Lejeune, 83 of class .C 1 , ix
© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 339
A. Alves de Sá, B. Louro, Sequences and Series,
https://doi.org/10.1007/978-3-031-67202-6
340 Index
G
N
Gauss, Carl Friedrich, 7
neighborhood
General Comparison Test, 95
of a, vii, 15
general term of a sequence, 1
of .+∞, 15
geometric progression, see geometric sequence
geometric sequence, 8 of .−∞, 15
common ratio, 8 Neper’s number, 30
definition, 8 Neper, Jhone, 30
general term, 9 Newton Binomial, viii
sum of the first n terms, 9 Numerical series
absolutely convergent, 87
I alternating harmonic, 86
indeterminate forms, ix alternating series, 82
infimum, viii associative property, 80
integer part, viii
Cauchy Condensation Test, 90
Integral Test, 93
Cauchy product, 113
interval, vii
Cauchy’s Criterion, 78
K Cauchy’s Root Test, 106
Kummer’s Test, 109 conditionally convergent, 87
Kummer, Ernst Eduard, 109 convergent, 69
definition, 69
L divergent, 69
L’Hôpital’s Rule, ix General Comparison Test, 95
Leibniz’s Test, 82 general term, 69
Leibniz, Gottfried Wilhelm, 82 geometric, 70
lower bound, viii
harmonic, 71
Integral Test, 93
M
Maclaurin Kummer’s Test, 109
binomial series, 194 p-series, 91
Maclaurin series of .sin(x), 196 partial sums, 69
Maclaurin series of .ex , 197 product of series, 113
Maclaurin, Collin, 194 Ratio Test, 101
maximum, viii rearrangement, 87
Index 341