Open access peer-reviewed chapter

Sequences and Series: An Introduction for Beginners

Written By

Luís Vieira

Reviewed: 30 October 2023 Published: 22 May 2024

DOI: 10.5772/intechopen.113863

From the Compact

Sequences and Series - An Introduction for Beginners

Authored by Luis Vieira

Chapter metrics overview

15 Chapter Downloads

View Full Metrics

Abstract

This mathematics book presents the main definitions and properties of sequences, series of real numbers, sequences, and series of functions. We combine a rigorous exposition with a set of intuitive examples to achieve an easily readable final text that is approachable by intermediate and more advanced calculus students from the secondary to the university level. This book is divided into three sections: an introductory one where we present the basic mathematical concepts and notation we will follow; a section devoted to sequences of real numbers; and a section dedicated to series of real numbers, sequences, and functions. These two last sections end with exercises with solutions for the readers to test their learning and knowledge. Our objective was to provide students and curious minds on the subject of the book with a tool for autonomous learning.

Keywords

  • calculus
  • real analysis
  • sequences of functions
  • sequences of real numbers
  • series of functions
  • series of real numbers
  • Taylor series

1. Preliminaries and notation

We decided to begin this book with a brief exposition of the main basic mathematical concepts we will use and assume the reader is familiar with. Also, the notation used in this book will be presented in the following paragraphs. For further reading on the topics addressed in this book, we recommend the detailed expositions presented in Refs. [1, 2, 3, 4, 5, 6, 7, 8, 9, 10]. Refs. [1, 2, 3, 4, 5, 6, 7] consist of theoretical fundamental texts while Refs. [8, 9, 10] present more practical approaches.

1.1 Basic set notation

In this book, we will deal with real numbers. Let us start by recalling the basic set notation.

We denote by N the set of all natural numbers excluding 0, that is,

N=123,

and the set of integers is denoted by Z,

Z=321,0,1,2,3.

Next, the set of rational numbers is denoted by Q:

Q=ab:abZb0.

Finally, we denote by R the set of real numbers, that is,

R=Qx:xis irrational,

where it is sufficient to say that being irrational is not being able to be written as the division of two integers.

Some additional symbols may be used to represent special subsets of the main sets presented above like, for example,

N0=N0;Np=nZ:npZ=xZ:x<0;R0+=xR:x0.

1.2 Upper and lower bounds of sets

Some sets of real numbers are limited as their elements will not be bigger or smaller than some values.

Definition 1.1 (Bounded sets). Let X be a non-empty subset of R.

  1. X is said to be bounded from above if there is a real number M such that xM,xX. In these conditions, M is called an upper bound of X.

  2. X is said to be bounded from below if there is a real number m such that xm,xX. In these conditions, m is called a lower bound of X.

  3. X is said to be bounded if it is bounded from above and below and unbounded if it is not bounded.

Definition 1.2 (Maximum, minimum, supremum, and infimum of sets). Let X be a non-empty subset of R. A real number is

  1. the maximum of X if it is an element of X and if it is an upper bound of X.

  2. the minimum of X if it is an element of X and if it is a lower bound of X.

  3. the supremum of X if it is the smaller of all upper bounds of X.

  4. the infimum of X if it is the bigger of all lower bounds of X.

Note that, for a given set, the existence of any of the numbers presented in Definition 1.2 is not guaranteed. Also, observe that, when the maximum and minimum of X do exist, they coincide with the supremum and infimum of X, respectively.

Example 1.1. Let X=]0,1]. For this set, the following statements are true.

  • This set is bounded from above, as [1,+[ is the set of all upper bounds of X.

  • The set X is bounded from below, as ],0] is the set of all lower bounds of X.

  • Hence, X is a bounded set.

  • The maximum of X is 1.

  • X does not have a minimum.

  • The supremum of X is 1.

  • The infimum of X is 0.

The concept of neighborhood of a number will also be used ahead. We define it in a simple way as follows.

Definition 1.3 (Neighborhood of a number). Let a be a real number. A neighborhood of a is a set that contains an interval ]aε,a+ε[, with ε>0.

1.3 Relations and functions

Given two sets X and Y, the Cartesian product X×Y is defined as

X×Y=xy:xXyY

and its elements are called ordered pairs.

A binary relation or correspondence, R, between X and Y is simply a subset of X×Y. That subset is often called the graph of R. If xyR, it means that x is related to y through R, we say that x is R-related to y, and we write xRy. The order is important, as xRy does not necessarily imply that yRx.

The sets X and Y are usually called the departure set and the destination set of R, respectively. The domain of R is the set of elements of X that are R-related to at least one element of Y, and the image of R is the set of elements of Y for which there is at least one element of x such that xRy.

There are different types of binary relations.

Definition 1.4 (Types of binary relations). Let R be a binary relation between the sets X and Y.

  1. If, for all xX and y1,y2Y we have xRy1xRy2y1=y2, then R is said to be a functional relation.

  2. If, for all xX, there exists at least one yY such that xRy, then R is said to be left-total.

  3. If R is functional and left-total, it is called a function.

  4. If, for all x1,x2X and yY we have x1Ryx2Ryx1=x2, then R is said to be injective.

  5. If, for all yY there exists at least one xX such that xRy, then R is said to be surjective or right-total.

  6. A function that is injective and surjective is said to be bijective or a bijection.

Example 1.2. Let X=1,2,3,Y=abc, and Z=αβ. Then:

  • R1=1a2b is a functional and injective relation between X and Y that is not a function;

  • R2=1α2β is a functional, injective, and surjective relation between X and Z that is not a function because it is not left-total;

  • R3=αaαbβc is a left-total, surjective, but not injective, relation between Z and Y that is not a function because it is not functional;

  • R4=1a2a3a is function between X and Y that is not injective or surjective;

  • R5=αaβb is an injective, but not surjective, function between Z and Y;

  • R6=1α2α3β is a surjective, but not injective, function between X and Z;

  • R7=1c2a3b is a bijective function between X and Y.

Usually, functions are denoted by the letter f and, if x and y are f-related, instead of writing xfy, we write y=fx.

1.4 Limits and derivatives of functions

In this section, we make a brief summary of two concepts that might be of interest for the reader to understand a few examples presented in this text.

The concept of limit of a function can actually be formally defined with the help of the limit of sequences, which will be surveyed in detail in Section 2.4 of Section 2. The idea behind both of them is, however, the same. Given a function f, we say that the limit of f, when x approximates to the value a, is b, and we write

limxafx=b,

if the images of the elements of the domain, which are converging to a, are converging to b. Note that a and b can be . In the special case of sequences, that we will study in the next section, the limits are always calculated when a=+ because of the nature of this special kind of functions.

The derivative of a function, f, of one variable, x, is essentially an instantaneous rate of change of the images of f with the values of x. It is denoted by f, Lagrange’s notation, or by dfdx, Leibniz’s notation, and can be defined by the limit (if the limit exists):

fx=limh0fx+hfxh.E1

If the limit in (1) is a real number, then f is said to be differentiable at x.

From (1), one can easily deduce formulae for the derivative of many functions. Two basic rules that can be deduced are very useful while differentiating are the product and the quotient rules: given two functions f and g, both differentiable, then

fgx=fxgx+fxgxandfgx=fxgxfxgxg2x,

provided, in the latter case, that gx0.

To differentiate other more complex analytic expressions of functions one makes use of the so-called chain rule. Indeed, if we have a composite function fg, then we have:

fgx=fgxgx,E2

provided that g is differentiable at x and f is differentiable at g(x).

For the reader to be able to follow some of the examples presented ahead, we will provide the most common derivatives in Table 1. All of the formulae presented can be obtained by making use of (1) and/or (2). For more insight on this derivative operator, including its operation properties, please consult any calculus book; as for the scope of this book, what was enlightened is enough, and we do not pretend to go deeper on this subject.

f(x)fxf(x)fx
a0uu
axaauau
axnnaxn1aunanun1u
x12xuu2u
xn1nxn1nununun1n
ln(x)1xln(u)uu
logbx1xlnblogbuuulnb
exexeuueu
bxbxlnbbuubulnb
sin(x)cos(x)sin(u)ucosu
cos(x)− sin(x)cos(u)usinu
tan(x)sec2xtan(u)usec2u
arcsin(x)11x2arcsin(u)u1u2
arccos(x)11x2arccos(u)u1u2
arctan(x)1x2+1arctan(u)uu2+1
sinh(x)cosh(x)sinh(u)ucoshu
cosh(x)sinh(x)cosh(u)usinhu
tanh(x)sech2xtanh(u)usech2u

Table 1.

Most common derivatives (aR,bR+\1 and u is a function with respect to x).

Finally, we will make use, at some point, of a technique, called l’Hôpital’s Rule, for computing limits of functions that are the division of two other functions. The technique has two cases of application that are presented next, but both work basically in the same way; that is, you just have to consider the derivatives of the numerator function and the denominator function and recalculate the limit. Usually, differentiating a function makes it simpler. This rule made it is first apparition in a book of the French mathematician Guillaume de l’Hôpital.

Theorem 1.1 (l’Hôpital’s Rule for 0/0). Let f and g be two differentiable functions on an open interval I containing aR and g is such that gx0 in I\a. If limxafx=limxagx=0, then

limxafxgx=limxafxgx,

so long as the limit is finite or ±.

Theorem 1.2 (l’Hôpital’s Rule for /). Let f and g be two differentiable functions on an open interval I containing aR and g is such that gx0 in I\a. If limxafx=± and limxagx=±, then

limxafxgx=limxafxgx,

so long as the limit is finite or ±.

Note that, in both cases, similar results hold for a=±.

In the following, we present some practical examples of the effectiveness of l’Hôpital’s Rule on calculating limits.

Example 1.3. Let’s use l’Hôpital’s Rule to compute

limx0ex+23x.

Indeed, we have

limx022ex3x=00limx0ddx22exddx3x=limx02ex3=2e03=23.

Example 1.4. Now we will compute

limx0sinx5x.

using l’Hôpital’s Rule. We have

limx0sinx5x=00limx0ddxsinxddx5x=limx0cosx5=15.

Example 1.5. In this example, we will apply l’Hôpital’s Rule to calculate

limx+ex12x3.

We have

limx+ex12x3=limx+ddxex1ddx2x3=limx+ex16x2=limx+ddxex1ddx6x2=limx+ex112x=limx+ddxex1ddx12x=limx+ex112=+12=+.

1.5 Mathematical induction

Since sequences are essentially, as we will shortly see, functions with domain N, it is important to recall the principle of mathematical induction, an important technique for proving properties among the naturals.

Proposition 1.1 (Principle of Mathematical Induction). Let P(n) be a statement involving the natural number n. The proposition

nN,Pn

is true if:

  1. the statement P(n) is true for n=1, that is, P(1) is true;

  2. if the statement P(n) is true for n=mN, then it is also true for n=m+1, that is,

mN,PmPm+1.

The principle of mathematical induction comes from the axiomatic definition of N which is defined as the smallest subset, S, of R such that, if xS, then x+1S. If the statement has this inductive property, that is, if it is valid for a number m, then it is also valid for it is successor, m+1, then verifying that the statement is valid for 1 originates a chain reaction that proves the statement for all-natural numbers.

Note that Proposition 1.1 can be generalized for sets Np, with pN. In this case, the first step is to verify if the statement is valid for n=p and not n=1. The rest is analogous.

Example 1.6. Using mathematical induction, let’s prove that, for each nN, we have

12+22+32++n2=nn+12n+16.

For n=1, we have only one element in the sum, therefore,

12=1×1+1×2×1+161=1×2×361=1,

and the formula is valid.

Now, suppose that, for a certain mN, we have

12+22++m2=mm+12m+16.

Under this assumption (usually called the hypothesis of induction), let’s prove that the formula is valid for m+1 (usually called the thesis of induction), that is, that is valid the equality

12+22++m+12=m+1m+22m+36=2m3+9m2+13m+66.

Indeed, using the hypothesis of induction, we have

12+22++m2+m+12=mm+12m+16+m+12=mm+12m+1+6m+126=2m3+3m2+m+6m2+12m+66=2m3+9m2+13m+66=m+1m+22m+36.

This means that, if the formula is valid for mN, then it is also valid for m+1. Since we already proved that the formula is valid for n=1, we can conclude, by mathematical induction, that the formula is valid for all nN.

Example 1.7. In this example, we will prove, using induction, that for each natural n, we have:

n+3!2n+3.E3

Indeed, for n=1, we have

1+3!=2416=21+3,

which means that (3) is valid.

Let’s suppose that, for a certain natural m, we have

m+3!2m+3.

Let’s prove that

m+1+3!2m+1+3m+4!2m+4.

We have:

m+4!=m+4m+3!m+42m+3>2×2m+3=2m+4,

using the induction hypothesis in the second step and, in the third step, the fact that m+3>2.Hence,

m+4!2m+4

and we can conclude, by induction, that (3) is valid for all n in N.

Advertisement

2. Sequences

In this section, we present a general survey on the topic of sequences. A sequence is just a function of arbitrary real numbers but whose domain is the set of natural numbers, N. The origin of such functions can be traced back farther until at least ancient Greece, particularly to the grounds of the mythical Pythagoras’ school, sixth century BC, and the practice of its followers.

This special kind of functions has many practical applications, since many real-life situations are mainly a discrete succession of outputs, which can be modeled by a sequence of real numbers. For example, consider the sequence of perfect squares, also known simply as the sequence of squares: 1, 4, 9, 16, 25, 36, and so on. This sequence is the function of domain N that takes each natural number and maps it into its square.

Additionally, sequences are able to open up a window toward some rather complex mathematical concepts, such as the different types of infinity, as well as the concepts of convergence and limit. In particular, sequences allow us to establish, in a rather easy and elegant manner, the concept of limit, which is a structural stone in the field of mathematical analysis.

This section is organized in the following way. First, in Section 2.1, we present some general properties of sequences such as monotony and bounds. In Sections 2.2 and 2.3, we present and characterize a couple of special kinds of sequences, which are the arithmetic and the geometric progressions, respectively. Then, in Section 2.4, we address the fundamental problem of sequence convergence, presenting the most relevant results in the subject and examples, and finally, in Section 2.5, we present useful techniques for computing the limits of sequences. We finish the section with some suggested exercises in Section 2.6.

2.1 General properties of sequences

We will start this section with a formal definition of sequence.

Definition 2.1 (Sequence). Let A be a non-empty set. A sequence, u, in A is any function that can be defined from N into A, that is,

u:NAnun.

The function values of the sequence u, that is, u(1), u(2), u(3), and so on, are called the terms of the sequence u, with u(1) being the first term, u(2) the second term, u(3) the third term, etc. The set of terms of a sequence u is precisely the image of the map u. The object of a term is also referred to as the order of the term. For instance, the fourth term of the sequence, u(4), is the term of order 4, and the nth term, u(n), is the term of order n.

Along this text, we will consider only real sequences, that is, sequences defined in R. We will also adopt the following standard notation for sequences. Instead of writing u(1), u(2), … , u(n),… , we will denote the terms of the sequence as u1,u2,,un,, and the sequence u, itself, will be represented as unnN, or simply by un.

In frequent cases, the term of order n of a sequence can be obtained by a formula involving n. Let us take another look at the example of the square numbers presented at the beginning of this section.

Example 2.1. Consider the sequence, un, of squares number (see Figure 1).

Figure 1.

Figure of squares.

The term of order n of this sequence is given by un=n2.

In cases similar to Example 2.1, where the term of order n is given by a certain formula involving n, un assumes a greater level of importance within the sequence un because it basically yields an algebraic expression that is capable of generating all the terms of un, given the value of n. In view of this, the term of order n of un is often referred to as the general term of the sequence.

The terms of a sequence un can also be plotted out in a Cartesian coordinate system. The graph of the sequence un is the set

1u1(2u2)(3u3)(nun).

Example 2.2. Consider the sequence vn defined by its general term

vn=3n+2n.

The graph of vn is the set

15(24)31134725175,

which can be graphed as in Figure 2:

Figure 2.

Graphically representation of a succession 2.

One can also define a sequence by recurrence, that is, after defining one, or more, of the terms of the sequence, all the other terms are obtained by solving a so-called recurrence equation involving terms already determined before. In general, each term is obtained by making use of the term or terms that precede it. A famous example of a sequence defined in such a manner is the sequence of the Fibonacci numbers.

Example 2.3. The Fibonacci numbers are the following: 1, 1, 2, 3, 5, 8, 13, 21, and so on. After the first two numbers, each new number is obtained by adding the previous two Fibonacci numbers. This sequence of numbers, FnnN, can be defined as follows:

F1=1F2=1Fn=Fn1+Fn2,nN\12

All the basic algebraic operations for real numbers are still valid for real sequences. Indeed, if annN and bnnN are two real sequences, then one can define the sequence of the sum, a+bnN, the sequence of the difference, abnN, the sequence of the product, abnN, and the sequence of the quotient, abnN, providing, in the latter case, that bn0,nN, whose general terms are defined as a+bn=an+bn,abn=anbn,abn=anbn and abn=anbn, respectively.

We proceed the exposition with some general properties of some sequences.

Definition 2.2 (Monotonic sequence). A sequence un is:

  1. increasing if, nN,unun+1;

  2. decreasing if, nN,unun+1;

  3. strictly increasing if, nN,un<un+1;

  4. strictly decreasing if, nN,un>un+1;

  5. monotonic if it is increasing or decreasing;

  6. strictly monotonic if it is strictly increasing or strictly decreasing.

We follow up Definition 2.2 with some examples.

Example 2.4. Consider the sequence wn with general term given by

wn=23n.

Since

wn+1wn=23n+123n=23n32+3n=3<0,

we conclude that wn>wn+1 holds for each nN, thus wn is strictly decreasing.

Example 2.5. Let un be the sequence whose general term is

un=22n.

In this case, we have

un+1un=22n+122n=21n22n=22n21n1n2n=21n2n>0,

since the formula in the denominator always yields positive numbers, regardless of the value of n. Therefore, we can conclude that un<un+1 holds for each nN, thus un is strictly increasing.

Example 2.6. In this example, consider vn to be the sequence defined by

vn=4n+21n.

For this sequence, we have

vn+1vn=4n+1+21n+14n+21n=4+21n+11n.

Observe that this expression yields the values 0 and 8, successively, when n is even or odd, respectively. This means that vn=vn+1 when n is even and vn<vn+1 when n is odd. Accordingly, vn is an increasing, but not strictly increasing, sequence.

Example 2.7. To finish this set of examples, consider the sequence wn, whose general term is given by

wn=n26n.

In this case, we have

wn+1wn=n+126n+1n26n=n2+2n+16n6n2+6n=2n5.

This expression yields negative values for n12 and positive values in the remaining cases. This means that wn>wn+1 when n=1,2 and wn<wn+1 when n>2. Accordingly, wn is a non-monotonic sequence.

In the following, we introduce the concept of bounded sequences.

Definition 2.3 (Bounded sequence). A sequence un is:

  1. bounded from above if there is a MR, such that unM,nN;

  2. bounded from below if there is a mR, such that unm,nN;

  3. bounded if it is bounded from above and below.

From Definition 2.3, we conclude that a bounded sequence, un, is one in which all terms are bounded by certain real numbers m and M, that is, munM. It is straightforward to see, though, that this assertion can be slightly modified. In fact, considering k=maxmM, we have that unk,nN.

There is a couple of immediate consequences relating to Definitions 2.2 and 2.3. An increasing sequence is clearly bounded from below by its first term, and a decreasing sequence is clearly bounded from above also by its first term.

One might be tempted to conclude that a strictly increasing sequence cannot be bounded from above, but such a conclusion is false. The same applies to strictly decreasing sequences bounded from below. The classical examples are the sequences of general terms un=1n and vn=1n. The sequence un is strictly increasing and vn is strictly decreasing, yet we have 1un0 and 0vn1, thus both sequences are bounded.

Example 2.8. Consider the sequence wn, whose general term is

wn=2n+5n+1.

By making use of Euclid’s division algorithm, we can rewrite the general term’s formula as

wn=2+3n+1.

Now, we observe that the sequence

1n+1

is strictly decreasing, bounded above by its first term and below by 0, and we proceed as it follows:

0<1n+1120<3n+1322<2+3n+12+32,

which allows us to conclude that 2<wn72, that is, wn is a bounded sequence.

2.2 Arithmetic progressions

In this section we will survey a class of sequences on which each term is obtained from its previous by adding a constant.

Definition 2.4 (Arithmetic progression). An arithmetic progression un is a sequence defined by recurrence as

u1=aun+1=un+d,nN,

where a and d are real constants.

The constant a in Definition 2.4 corresponds to the first term of the arithmetic progression. It is immediate to observe that the difference between any two consecutive terms of an arithmetic progression is constant and equal to d:

nN,un+1un=d

and, in view of this property, d is normally called the common difference of the arithmetic sequence un.

Example 2.9. Consider the sequence un whose first terms are:

2,5,8,11,14,

un is the arithmetic sequence of first term 2 and common difference 3.

The value of the common difference of an arithmetic sequence contains all the information we need to analyze the sequence’s monotony. In fact, if the common difference is positive, that means that the arithmetic sequence is strictly increasing, since each term increases relatively to its predecessor by adding a positive constant. Similarly, if the value of the common difference is negative, then, by an analogous argument, the sequence is strictly decreasing. The particular case in which we have a null common difference yields a trivial constant arithmetic sequence.

One can easily derive a formula for the general term of an arithmetic sequence. Consider the arithmetic sequence of the first term a and common difference d. Then, we have

u1=a,u2=u1+d=a+d,u3=u2+d=k+d+c=a+2d,u4=u3+d=k+2d+c=a+3d,un=a+n1d,E4

which is the general term of the arithmetic sequence of first term a and common difference d.

Example 2.10. Continuing Example 2.9, the general term of the arithmetic sequence un, of first term 2 and common difference 3, is

un=2+n1×3=2+3n3=3n1.

The general term formula (4) can be generalized for any given term of the sequence:

un=um+nmd.

This means that we do not need the first term of the sequence to compute the general term. Any other term can do the trick.

Example 2.11. Continuing Examples 2.9 and 2.10, the general term of the arithmetic sequence un, of first term 2 and common difference 3, can also be computed by using the third term u3=8:

un=8+n3×3=8+3n9=3n1.

We will finish this section with a formula for calculating the sum of a finite number of terms of an arithmetic sequence. Let un be an arithmetic sequence of common difference d. Then, we have

2i=1nui=2u1+u2+u3++un2+un1+un=u1+un+u2+un1+u3+un2+++un2+u3+un1+u2+un+u1E5

Observe that, for any natural m we have:

um+unm1=u1+m1d+u1+nm11d=u1+mdd+u1+ndmd=u1+u1+n1d=u1+un.E6

Then, by making use of property (6), one can rewrite equality (5) as:

2i=1nui=u1+un×n,E7

which, in turn, allows us to conclude that the sum, Sn, of the first n terms of an arithmetic sequence, un, regardless of its common difference, can be given by:

Sn=i=1nui=u1+un2×n.E8

This formula is usually associated with the German mathematician Carl Friedrich Gauss, who allegedly surprised the teacher by computing the sum of the first hundred natural numbers, 5050, by adding 1 to 100, multiplying the result by 100, and dividing by 2.

Formula (8) can be generalized in a straightforward way to calculate the sum of any consecutive terms of the sequence un between terms um and un as:

Sm,n=i=mnui=um+un2×nm+1.E9

Example 2.12. Continuing Examples 2.9, 2.10, and 2.11, the sum of the ten first terms of the arithmetic sequence un, of first term 2 and common difference 3, is:

S10=2+3×1012×10=3102=155.

The sum of the ten consecutive terms of un starting from u3 is:

S3:12=3×31+3×1212×10=432×10=215.

2.3 Geometric progressions

In this section, a different type of sequence is presented. Contrary to the arithmetic progressions presented in Section 2.2, which have a linear growth, the sequences presented next have an exponential growth, as each of its terms is obtained by multiplying, not adding, its predecessor by a constant number.

Definition 2.5 (Geometric progression). A geometric progression un is a sequence defined by recurrence as

u1=aun+1=un×r,nN,

where a and r are real constants.

The value a is the first term of the sequence. If a=0, then the geometric progression is the null sequence. Also, if a0 but r=0, then all the terms of the geometric progression from the second onwards will be null. In the remaining case, that is, with a0 and r0, all the terms of the geometric progression will not be null. In this latter case, as an immediate consequence, we observe that the ratio between any two consecutive terms of a geometric progression is constant and equal to r:

nN,un+1un=r,

and because of this r is usually referred to as the common ratio of the geometric progression.

Example 2.13. Consider the sequence un of the numbers

3,6,12,24,48,

The sequence un is the geometric progression of first term 3 and common ratio 2.

As we did for the arithmetic progressions, we can also derive a formula for the general term of a geometric progression. Consider the geometric sequence of the first term a and common ratio r. Then, we have

u1=a,u2=u1×r=a×r,u3=u2×r=a×r×r=a×r2,u4=u3×r=a×r2×r=a×r3,un=a×rn1,E10

which is the general term of the geometric sequence of the first term a and common ratio r.

Example 2.14. Continuing Example 2.13, we have that the general term of the geometric progression un of first term 3 and common ratio 2 is:

un=3×2n1.

One can obtain the general term’s formula by making use of any term of the geometric progression and not necessarily the first term. In fact, it is easy to obtain that:

un=um×rnm,E11

for any given term of order mN.

Example 2.15. Continuing Examples 2.13 and 2.14, we can also obtain the general term of the geometric progression un of first term 3 and common ratio 2, by using the term u3=12 and formula (11):

un=12×2n3=3×22×2n3=3×22+n3=3×2n1.

The analysis of the monotony of geometric progressions is not as straightforward as in the case of the arithmetic progressions. Let’s exclude the trivial cases with first term or common ratio equal to 0 which yield null sequences at least from the second term onwards.

Proposition 2.1. Let un be a geometric progression with u10 and common ration r0. The following hold:

  1. If r=1, then un is constant and all terms are equal to u1.

  2. If r>1, then

    1. if u1>0, then un is increasing.

    2. if u1<0, then un is decreasing.

  3. If 0<r<1, then

    1. if u1>0, then un is decreasing.

    2. if u1<0, then un is increasing.

  4. If r<0, then un is not monotonic, as it has alternating sign terms.

The proof of Proposition 2.1 is left as an exercise for the reader.

As in the previous section, there is also a formula for calculating the sum, Sn, of n consecutive terms of a geometric progression un.

The case where r=1 is trivial. Since, in this case, all terms are equal to u1, then Sn=n×u1.

Let’s consider the more general case with r1. We have that

Sn=i=1nui=u1+u2+u3++un1+un.E12

Multiplying both members of equality (12) by r, we obtain

Snr=u1r+u2r+u3r++un1r+unr=u2+u3+u4++un+un+1.E13

Now, subtracting equalities (12) and (13), we obtain:

SnSnr=u1+u2+u3++un1+unu2+u3+u4++un+un+1Sn1r=u1un+1Sn1r=u1unrSn1r=u1u1rn1rSn1r=u1u1rnSn1r=u11rnSn=u11rn1r,E14

which gives the sum of the n first terms of the geometric progression un with first term u1 and common ratio r. Note that formula (14) can be easily applied to calculate the sum of any consecutive terms of un between um and un:

Sm:n=um1rnm+11r.E15

Example 2.16. Continuing Examples 2.13, 2.14, and 2.15, we have that the sum the ten consecutive terms of the geometric progression un of first term 3 and common ratio 2, starting at term u4=24 is:

S4:13=24121012=24552.

2.4 Convergent sequences

Some sequences have an interesting property: their terms approximate to a certain value as the order increases. This kind of sequence is called convergent, and we will survey it along this section.

Definition 2.6 (Limit of a sequence, convergent sequence). Let un be a sequence. A real number a is said to be the limit of un if

ε>0,mN:nN,nmuna<ε,

and we write limn+un=a,limnun=a or simply limun=a. If a number a exists in these conditions, un is said to be convergent. Otherwise, un is divergent.

When we say that limun=a the idea is that we can always find an order from which all the terms are as close as we want, or as we previously defined, from the value a. Therefore, un gets closer and closer to that value a, that is, it converges to a.

Example 2.17. Consider the sequence un of general term

un=2n+1n.

Let’s prove that limun=2.

Let ε>0. We have

un2<ε2n+1n2<ε2n+12nn<ε1n<ε1n<εn>1ε.

Then, for all ε>0, there is a natural number m such that for all nN, we have nmuna<ε. That number m can be chosen as any natural number bigger that 1/ε. Hence, limun=2.

An expected property of the definition of limit of a sequence is that when the limit exists, it is unique; that is, a convergent sequence cannot have two distinct limits.

There are two particular cases of divergent sequences that should be highlighted.

Definition 2.7 (Infinite limit of a sequence). Let un be a sequence.

  1. It is said that un has limit +, and we write limun=+, if

    L>0,mN:nN,nmun>L.

  2. It is said that un has limit , and we write limun=, if

    L>0,mN:nN,nmun<L.

Example 2.18. Let un and vn be the sequences defined by un=3n+2 and vn=32n, respectively. Let’s prove that limun=+ and limvn=.

Let L>0.

Regarding un, we have

un>L3n+2>Ln>L23.

Then, there is a order mN such that, for all nN, we have

nmun>L.

We just have to choose m to be a natural number bigger that L2/3. Hence, limun=+.

Regarding vn, we have

vn<L32n<Ln>L+32.

Then, there is a order mN such that, for all nN, we have

nmvn<L.

In this case, it is enough to choose m to be a natural number bigger that L+3/2. Hence, limvn=.

There are some interesting and well-known results involving the convergence of sequences.

Theorem 2.1. Let X be a finite subset of N and un and vn be two sequences such that, for all nN\X, we have un=vn. Then, un is convergent if and only if vn is convergent. Also, limun=limvn.

Theorem 2.1 basically states that two different sequences that differ only in a finite number of terms converge or diverge in the same manner. We leave the proof as an exercise for the reader.

Example 2.19. Let un be the sequence whose general term is un=2n and vn the sequence defined by:

vn=1n+2,ifn<25un,ifn25.

Since un and vn only differ in a finite number of terms (the first 24 terms), then limvn=limun=2=.

As it is implied in the resolution of Example 2.19, let it be clear that the first step in evaluating a sequence’s limit is to replace every instance of n, in the sequence’s general term, with + and calculate. Operating with is, therefore, of great importance. We will not elaborate on this topic for now, as this section deals with a more theoretical part of convergence, but it will be one of the main topics surveyed in Section 2.5.

Theorem 2.2. Let un be a sequence. If un is convergent, then un is bounded.

Proof. Let un be a convergent sequence and suppose that limun=a. Then, by definition, for any fixed ε>0, there is an order mN such that una<ε, for all nm. This means that the set of terms of un with n>m belongs to Va, the neighborhood of a. We observe that a+ε and aε constitute an upper bound and a lower bound of that set of terms of un, respectively; thus, that set is bounded.

As for the remaining terms of un, that is, the terms with n<m they constitute a finite set that is naturally bounded. We can, therefore, conclude that all terms of un are bounded, hence un is a bounded sequence. □

Observe that the converse from Theorem 2.2 is not true. Consider the trivial alternating sequence, that is, a sequence whose terms alternate between positive and negative, with general term 1n. It is clearly a bounded sequence, but it does not converge as it keeps alternating from −1 to 1.

Theorem 2.3 (Squeeze Theorem for Convergent Sequences). Let un,vn and wn be sequences such that

mN:nN,n>munvnwn.

If limun=limwn=a, where aR, then limvn=a.

Proof. Let ε>0. Then, since limun=limwn=a, there are natural numbers m1 and m2 such that

nm1una<εaε<un<a+ε,nm2wna<εaε<wn<a+ε.

Then, considering m=maxm1m2, we have

nmaε<unvnwn<a+ε,

which means that

nmvna<ε.

Hence, limvn=a. □

Note that the result of Theorem 2.3 can be straightforwardly extended to the cases where a=±.

Example 2.20. Let un be the sequence whose general term is

un=sinnn.

Since the sine yields values in the interval [−1, 1], we can write

1nsinnn1n.

Now, observing that lim1n=lim1n=0, we can conclude, by Theorem 2.3, that limun=0.

From Theorem 2.3, one can prove the following result.

Theorem 2.4. Let un be a sequence such that limun=0. Then, limun=0.

Theorem 2.4 is especially useful for computing certain limits involving alternating sequences.

Example 2.21. Let un be the sequence defined by

un=1n2n+1.

Since

limun=lim12n+1=1+=0,

we can conclude, by Theorem 2.4, that limun=0.

Theorem 2.5. Let un and vn be two convergent sequences such that

limun<limvn.

Then, there is a mN such that

nmun<vn.

Proof. Let un and vn be two convergent sequences with limun=a and limvn=b such that a<b. Let ε=ba2. Note that . By the definition of limit, there are natural numbers m1 and m2 such that

nm1una<εaε<un<a+ε,nm2vnb<εbε<vn<b+ε.

Now, observing that a+ε=bε=a+b2, and taking m=maxm1m2, we have that

n>mun<a+ε=bε<vn.

Theorem 2.6. Let un be a sequence.

  1. If un is increasing and bounded from above, then it is convergent.

  2. If un is decreasing and bounded from below, then it is convergent.

  3. If un is monotonic and bounded, then it is convergent.

Proof. We will just present proof for assertion 1 from Theorem 2.6. Assertion 2 is analogous, and assertion 3 is a direct consequence from the previous two.

Let un be an increasing sequence bounded from above. We can immediately conclude that un is bounded, that is, that the set of terms of un is a bounded set. Since it is bounded and it is a non-empty set, then we can conclude that it has a supremum. Let a be the supremum of the set of terms of un. We will prove that limun=a.

Let ε>0. By definition of supremum, we have that, for all nN,una. We also have that there is at least one mN such that um>aε. Since un is an increasing sequence, we also must have that, for all n>m,unum>aε. Therefore, we can conclude that there is an order mN such that, for all nN, we have una<ε, if n>m. Thus, limun=a and un is a convergent sequence. □

The result presented above has direct application as it is shown in the following example.

Example 2.22. Consider the sequence un with general term

un=2n1n.

Let’s show that un is convergent.

Firstly, since

un+1un=2n+11n+12n1n=2n+1n+12n1n=2n+1n2n1n+1nn+1=1nn+1>0,

then un is increasing.

Now, observe that un=21/n and that

0<1n111n<0121n<2.

Hence, un is bounded. Then, from Theorem 2.6, we can conclude that un is convergent.

Note that being convergent does not imply being monotonic. For instance, the sequence with general term 1n/n is convergent to 0, but it is not monotonic as it is an alternating sequence.

Theorem 2.7. Let un and vn be two sequences such that un is bounded and limvn=0. Then limun×vn=0.

Proof. Let un be a bounded sequence and vn be a sequence such that limvn=0. Let wn=un×vn and ε>0.

Since un is bounded, then there is a kR such that unk, for all nN. Then, we can write

wn=un×vn=un×vnkvn.

Now, since limvn=0, there is an order mN such that, for all n>m,vn<ε/k. Then, we can conclude that, for all n>m, we have

wnkvn<ε,

that is, limwn=0.

Some potentially complicated limits can be computed using the result presented in Theorem 2.7.

Example 2.23. Let’s compute the value of

limcosn2n+1.

Observe that

cosn2n+1=cosn2×1n+1.

The sequence whose general term is cosn2 is clearly bounded, as the cosine yields values in the interval [−1, 1].

As for the sequence whose general term is 1/n+1, we have that

lim1n+1=1+=0.

Then, we can conclude, by Theorem 2.7, that

limcosn2n+1=0

In the following, we present a set of operating properties regarding limits of sequences.

Proposition 2.2. Let un and vn be convergent sequences such that limun=a and limvn=b. Let αR and kN. We have:

  1. limun+vn=a+b;

  2. limunvn=ab;

  3. limun×vn=ab;

  4. limαun=αa;

  5. limunk=ak;

  6. limun=a;

  7. lim1un=1a, if un0,nN and a0;

  8. limunvn=ab, if vn0,nN and b0;

  9. limunk=ak, with if un0,nN, if k is an even number.

We finish this survey on sequences by presenting a very important class of sequences called the Cauchy sequences, named after the French mathematician Augustin-Louis Cauchy. Their relevance and their implications on convergence make them frequently referred to as fundamental sequences.

Definition 2.8 (Cauchy sequence). The sequence un is a Cauchy sequence if

ε>0,pN:m,nN,m,n>pumun<ϵ.

Observe that the above formal definition simply states that the terms of a Cauchy sequence become as close as we want as the sequence progresses.

Theorem 2.8. Let un be a Cauchy sequence. Then un is bounded.

Proof. Let un be a Cauchy sequence. By definition, for any fixed ε>0, there is an order pN such that umun<ε, for all natural numbers m,n>p. In particular, if n=p+1, we have that umup+1<ε, for all m>p, which means that the set of terms of un with n>p is bounded as ε+up+1 is a lower bound and ε+up+1 is an upper bound for this set. As for the set of terms of un with n<p, since it is a finite set, it is also bounded. Hence, un is a bounded sequence.

Using Theorem 2.8, one can prove the following result that states that being a convergent sequence and being a Cauchy sequence are the same thing.

Theorem 2.9. Let un be a sequence. Then, un is convergent if and only if un is a Cauchy sequence.

2.5 Techniques for evaluating limits

In this section, we will present a set of relevant tools for computing the limits of certain sequences. We will also analyze the main indeterminate forms that one can attain when computing limits, and we will discuss techniques to evaluate these indeterminate forms.

We will start this section with a result that can be of extreme importance in all domains of sequence analysis.

Theorem 2.10. Let un be a sequence and f a real function such that, nN,fn=un. Then we have:

  1. If limx+fx=a, for aR, then limun=a.

  2. If f is increasing (decreasing), then un is increasing (decreasing).

This result allows us to apply functional analysis to solve some problems involving sequences by considering a corresponding real function.

Example 2.24. Calculating the limit of the sequence un defined by

un=3nlnn+2n

may prove to be difficult at first glance, but if one considers the real function f defined by

fx=3xlnx+2x,

then we obtain, by applying the well-known L’Hôpital’s Rule,

limx+fx=limx+3xlnx+2x=limx+ddx3xddxlnx+2x=limx+31x+2=30+2=32.

By Theorem 2.10, we can conclude that limun=3/2.

Example 2.24 brings out again the need to clarify operations involving . The following proposition presents a list of relevant properties. Before, however, let us introduce the following notation.

Let un be a convergent sequence with limit equal to a. If un converges to the value a through values which are strictly smaller than a, then we write limun=a. If un converges to the value a through values which are strictly bigger than a, then we write limun=a+.

As a trivial example, consider the sequences un and vn whose general terms are un=1/n and vn=1/n. It is easy to verify that limun=0+ and limvn=0.

Proposition 2.3. While evaluating a limit, the following equalities hold.

  1. a±=±,aR;

  2. ++=+;

  3. =;

  4. a×±=±,aR+ or a=+;

  5. a×±=,aR or a=;

  6. a±=0,aR;

    1. a+=0+,aR+;

    2. a=0,aR+;

    3. a+=0,aR;

    4. a=0+,aR;

  7. a0=,aR\0;

    1. a0+=+,aR+ or a=+;

    2. a0+=,aR or a=;

    3. a0=,aR+ or a=+;

    4. a0=+,aR or a=;

  8. k=,kN;

    1. +k=+,kN;

    2. k=+,kN and k even;

    3. k=,kN and k odd;

  9. +k=+,kN\1;

  10. k=,kN\1 and k is odd.

The properties presented are invaluable to operate with and compute the values of limits by direct substitution of the variable n in the general term by +. The following example presents some cases.

Example 2.25. Let’s compute the value of the following limits.

  1. lim2n5

    lim2n5=2×+5=+5=+.

  2. lim233n

    lim233n=233×+=23=2=0.

  3. limn21n

    limn21n=+2×1+=+×1=+×=.

Left out from the list presented in Proposition 2.3 are the cases of +,,00,0×,0,00 and 1. These are called indeterminate forms as they can actually be equal to different real values or ± depending on the situation at hand. We will observe this particular behavior in the following paragraphs, on which we will survey some of the techniques one can apply to evaluate different types of indeterminate forms.

2.5.1 Indeterminate form

This type of indeterminate form, , usually arises from evaluating the limit of sequences constructed in the form of a polynomial of variable n, although it can also appear in other contexts. For example, if one attempts to evaluate the limit of the sequence un whose general term is un=n22n+1 by direct substitution, one will obtain this indeterminate form.

In the context of sequences defined by polynomials, this difficulty is very easy to surpass. Indeed, one needs only to collect the higher degree term in un and substitute again. The indeterminate form will disappear.

Example 2.26. Consider the sequence un with general term un=n22n+1. We have:

limun=limn22n+1=n212n+1n2=+12++1+=+10+0=+.

Example 2.26 leads to a more general property.

Proposition 2.4. Let p(x) be a non-null polynomial with leading coefficient a and un be a sequence such that, nN,un=pn. Then, the following statements hold.

  1. If a>0, then limun=+;

  2. If a<0, then limun=.

When defining sequences with the help of the concept of a polynomial, it is implied that the degree of the powers of n are greater or equal than 0. The next result lists clarifies all possibilities regarding the limits of rational powers of n.

Proposition 2.5. Let qQ and un be the sequence whose general term is un=nq. Then, the following hold.

  1. If q>0, then limun=+;

  2. If q=0, then limun=1;

  3. If q<0, then limun=0.

Proof. Let qQ.

  1. If q>0, let L>0. Then, we have

    un>Lnq>Ln>Lq.

    Therefore, choosing any natural m bigger than Lq, we have that

    nN,nmun>L,

    which is the same as saying that limun=+.

  2. If q=0, then un=1,nN. Hence, limun=1.

  3. If q<0, let ε>0. Then, we have

un0<εnq<ε1nq<ε1nq<εnq>1εn>εq.

Therefore, choosing any natural m bigger than εq, we have that

nN,nmun0>ε,

which is the same as saying that limun=0.

The next example presents another occasion where the indeterminate form appears.

Example 2.27. Consider the sequence un with general term un=n+1n. The limit of un can be obtained by making use of the well-known multiplication identity ABA+B=A2B2 in the following manner:

limun=limn+1n=limn+1nn+1+nn+1+n=limn+1nn+1+n=lim1n+1+n=1++=1+=0.

We will finish this subsection by presenting one type of sequence for which the limit evaluation might result in indeterminate forms, particularly the one that is under the scope in these paragraphs.

Proposition 2.6. Let aR+ and un be the sequence whose general term is un=an. Then, the following statements hold.

  1. If a>1, then limun=+;

  2. If a=1, then limun=1;

  3. If 0<a<1, then limun=0.

Proof. Let aR+.

  1. If a>1, let L>0. We have

    un>Lan>Ln>logaL.

    Then, choosing any natural number bigger than logaL, we have that

    nN,nmun>L,

    which means that limun=+.

  2. If a=1, then all terms of un are equal to 1. Hence, limun=1

  3. If 0<a<1, then a1>1, which allows us to write

    limun=liman=lim1a1n=1lima1n=1+=0.

In the following, we present an example of an indeterminate form that arises from subtracting two sequences of the form presented in Proposition 2.6.

Example 2.28. Let un be the sequence defined by un=5n2n.

Computing the limit of un gives us, by Proposition 2.6, the indeterminate form . We can evaluate this limit by rewriting the general term, collecting the biggest power, and applying Proposition 2.6:

limun=lim5n2n=lim5n12n5n=lim5n125n=+10=+.

2.5.2 Indeterminate form

The type of indeterminate form represented by can appear when working out limits of sequences whose general term is a division of polynomials in the variable n, but it is not restricted to these cases. Trivial examples, which can be solved just by simplifying the fraction, are presented next.

Example 2.29. Let un,vn and wn be the sequences whose general terms are

un=n23n,vn=n3nandwn=n3n2,

respectively.

By simplifying the general terms, we obtain

limun=limn23n=limn3=+3=+,
limvn=limn3n=lim13=13,
limwn=limn3n2=lim13n=13×+=1+=0.

Example 2.29 enlightens in a very simple way the concept of indeterminate form. Observe how the three cases of the indeterminate form , after simplification, yield three different outcomes.

Some cases, however, cannot be solved through direct simplification only. Similarly to what we did in the previous section, collecting the higher degree terms of polynomials can be useful before simplifying. In the following example, three cases are presented using this technique.

Example 2.30. Consider the sequence un,vn and wn with general terms defined by

un=2n3n2+1n3,vn=3n2+n2n23n+1andwn=32nn2+n+1,

respectively.

In each case, direct substitution ends with the indeterminate form . The idea is to collect the term with a higher degree in the fraction’s numerator and repeat the process with the denominator. Then, simplify and, finally, substitute. We have:

limun=lim2n3n2+1n3=lim2n31n22n3+12n3n13n=lim2n2112n+12n313n=+11++1+13+=+10+010=+,
limvn=lim3n2+n2n23n+1=lim3n21+n3n223n2n213nn2+1n2=lim31+13n23n21+3n1n2=31+1+2+1+3+1+=31+001+00=3,
limwn=lim32nn2+n+1=lim2n32n+1n21+nn2+1n2=lim232n+1n1+1n+1n2=23++1+1+1++1+=20+1+1+0+0=2+=0.

The three different cases presented in Example 2.30 allow us to straightforwardly establish the following more general result.

Proposition 2.7. Let p(x) and q(x) be two non-null polynomials with degrees k and l and leading coefficients a and b, respectively. Let un be the sequence defined by

un=pnqn,nN.

Then, the following statements hold.

  1. If k>l and ab>0, then limun=+;

  2. If k>l and ab<0, then limun=;

  3. If k=l, then limun=ab;

  4. If k<l, then limun=0.

Next, we present two examples where an indeterminate form appears when evaluating the limit of sequences defined with radicals.

Example 2.31. Let un and vn be the sequences defined by:

un=n22n+1andvn=2n+1n+2.

When evaluating the limit either of un or vn by direct substitution we come upon the indeterminate form . In each case, we will solve the indeterminate form by collecting terms and rewriting the fraction. Indeed, we have

limun=limn22n+1=limn12nn2+1n=lim12n2+1n=12+2+1+=102+0=12,

as well as

limvn=lim2n+1n+2=lim2n1+12nn2nn2+2n2=lim2n1+12nn1n+2n2=lim21+12n1n+2n2=21+1+1++2+=21+00+0=20+=+.

Note that the square root will always attain 0 through values bigger than 0 as it is not defined in R.

In the following, we present an example of an indeterminate form that arises from dividing sequences of the form presented in Proposition 2.6.

Example 2.32. Consider the sequence un with general term given by

un=4n2n+3n.

Clearly, evaluating the limit of un by simple substitution results in the indeterminate form . In a similar manner than that of Example 2.28, we will solve this problem by collecting the higher power of the numerator and denominator like it follows.

limun=lim4n2n+3n=lim4n3n2n3n+1=lim4n3n×123n+1=lim43n×123n+1=+10+1=+

Example 2.33. Let vn be the sequence defined by

vn=n7n.

Attempting to evaluate the limit of vn directly gives us the indeterminate form . Applying L’Hôpital’s Rule, to the function f, such that

fx=x7x,

allows us to obtain the limit of vn. Indeed,

limx+fx=limx+x7x=limx+ddxxddx7x=limx+17xln7=1+=0.

Hence, by Theorem 2.10, limvn=0.

2.5.3 Indeterminate form 00

While working out the limits of sequences, the indeterminate form 00 arises mainly when dividing two sequences that converge to 0 and can be solved by simplifying the expression at hand, eventually obtaining a indeterminate form and applying some technique surveyed in the previous subsection.

Example 2.34. Consider the sequences un and vn with general terms given by

un=1n+1andvn=3n22.

We have that

limunvn=lim1n+13n22=00limn223n+1=limn212n23n1+1n=limn12n231+1n=+12+31+1+=+1031+0=+.

Example 2.35. Let wn be the sequence defined by

wn=ln11n2ln12n2.

If we try to evaluate the limit of wn by direct substitution, we obtain the indeterminate form 00.

Consider the function f such that

fx=ln11x2ln12x2.

Applying L’Hôpital’s Rule, we have

limx+fx=limx+ln11x2ln12x2=00limx+ddxln11x2ddxln12x2=limx+2x311x24x312x2=limx+12x2211x2=10210=12.

Hence, by Theorem 2.10, limwn=12.

2.5.4 Indeterminate form 0×

The case of the indeterminate form is solved using the techniques presented in the two previous subsections as it can always be converted into either the indeterminate form or 00.

Example 2.36. Let un,vn and wn be the sequences defined by

un=3n,vn=12nandwn=n2.

We have:

limun×wn=lim3n×n2=0×lim3n6n=lim3n163nn=lim312n=312+=310=3.

Also,

limvn×wn=lim12n×n2=0×limn22n=limn2n22n=limn2nlim21n=limn2n0=limn2n.E16

Now consider the real function f defined by

fx=x2x,

applying L’Hôpital’s Rule, we have

limx+fx=limx+x2x=limx+12xln2=limx+1+=0.

Therefore, we can conclude, by Theorem 2.10, that (16) is equal to 0.

Example 2.37. Let un and vn be the sequences defined by

un=lnnandvn=1n.

We have that

limun×vn=limlnn×1/n=0×limlnnn.E17

The limit (17) results in the indeterminate form . As we did in previous examples, let us consider the function f such that

fx=lnxx.

Applying L’Hôpital’s Rule, we obtain

limx+fx=limx+lnxx=limx+ddxlnxddxx=limx+1x1=limx+1x=1+=0.

Then, by Theorem 2.10, we can conclude that limun×vn=0.

2.5.5 Indeterminate forms 00,0 and 1

In this subsection, we present some examples for each of the remaining indeterminate forms, 00,0, and 1, which were not addressed before. We apply the natural logarithm to the expressions to solve the indeterminate forms.

Example 2.38. In this example, we will calculate the value of

lim1n1n.

We have:

lim1n1n=00limeln1n1n=lime1nln1n=lime1nlnn=elimlnnn=e0=1

using Example 2.37.

Example 2.39. In this example, we will calculate the value of

limn1n.

We have:

limn1n=0limelnn1n=lime1nlnn=limelnnn=elimlnnn=e0=1

using Example 2.37.

Example 2.40. In this example, we will calculate the value of

lim1+1n2n.

We have:

lim1+1n2n=1limeln1+1n2n=limenln1+1n2.

Considering the real function f, defined by

fx=xln1+1x2,

we have

limx+fx=limx+xln1+1x2=×0limy01yln1+y2=00limy0ddyln1+y2ddyy=limy02y1+y21=0,

changing variable y=1/x and applying l’Hôpital’s Rule.

Hence, we conclude that the limit at hand is equal to e0=1.

We finish this subsection with two results containing two notable cases for solving indeterminations treated here, 1 and 0.

Proposition 2.8. Let un be a sequence of real numbers such that un0 for nm, with mN, converging to ± and aR. Then,

lim1+aunun=ea.

Proposition 2.9. Let aR+ and nN. Then,

  1. liman=1;

  2. limnn=1.

2.6 Suggested exercises

General properties of sequences

  1. Let un be the sequence of real numbers 0, 2, 6, 12, 20, 30, …

    1. Find the general term of un.

    2. Verify that 156 is a term of un. Find its order.

    3. Solve the condition 400<un<600.

  2. Let vn be the sequence whose general term is given by:

    vn=6n+83n+2.
    1. Find the first five terms of vn.

    2. Verify if 94 is a term of vn.

    3. Prove that vn is a decreasing sequence.

    4. Show that vn is bounded. Determine a lower bound and an upper bound of vn.

    5. Determine the order from which vn is not greater than 2,1.

  3. Consider the sequence wn whose first five terms are:

    14;89;2716;6425;12536
    1. Find the general term of wn.

    2. Determine w10.

    3. Show that wn is unbounded.

  4. Let un be the sequence defined by recurrence as

    u1=5un+1=3un+10,nN.
    1. Find the first four terms of un.

    2. Using mathematical induction, show that

      un=532×3n3,nN.

    3. Show that un is monotonic.

    4. Prove that un is not bounded from above.

  5. Let vn be the sequence defined by

    vn=sin4.
    1. Find the image of vn.

    2. Justify that vn is bounded.

    3. Justify that vn is not monotonic.

  6. Let wn be the sequence defined by

    wn=n2,ifn45n4n,ifn>4.
    1. Study the monotony of wn.

    2. Show that wn is bounded.

  7. Consider the sequence un of general term

    un=2n7n+1.

    Determine the number of positive terms of un that are smaller than 75.

  8. Let vn be the sequence whose general term is

    vn=12n2n+3.
    1. Obtain the term of order 5 of vn.

    2. Verify if 2529 is a term of vn

    3. Study the monotony of vn.

    4. Show that vn is bounded.

  9. Study the monotony of the sequence wn defined by wn=1n+3n.

    Arithmetic an geometric progressions

  10. Let un be the arithmetic sequence defined by:

    u1=2un+1=un+3,nN.
    1. Obtain the general term of un.

    2. Determine the value of u31u10.

    3. Determine the sum of the first 20 terms of the un.

  11. Let vn be the sequence whose general term is

    vn=15n3.
    1. Show that vn is an arithmetic progression.

    2. Justify that vn is a decreasing sequence.

    3. Determine the sum of the 15 consecutive terms of vn after v22.

  12. Let wn be an arithmetic progression, such that w5=12 and w13=44.

    1. Determine the common difference of wn and classify its monotony.

    2. Obtain the general term of wn.

    3. The sum of the first m terms of wn is 360. Determine the value of m.

  13. Consider the geometric progression un defined by

    u1=4un+1=un5,nN.
    1. Obtain the general term of un.

    2. Determine the monotony of un.

    3. Prove that un is bounded.

    4. Compute the exact value of the sum of the first 10 terms of un.

  14. Consider the sequence vn whose general term is

    vn=52n1.
    1. Show that vn is a geometric progression.

    2. Determine the monotony of vn.

    3. Show that vn is bounded.

    4. Obtain the exact value of v3+v4++v9.

  15. Consider the arithmetic progression wn defined by

    w1=50wn+1=wn+2,nN.

    For a certain pN, the sum of the first p terms of wn is equal to p. Determine the value of p.

    Limits and convergence

  16. Let un be the sequence defined by

    un=2nn+3.

    Using the definition of convergent sequence, show that limun=2.

  17. Consider the sequence vn such that

    vn=2cosn+1.

    Using the Squeeze Theorem for Convergent Sequences, show that

    limvn=0.

  18. Evaluate limwn where

    1. wn=4n2+n5

    2. wn=7n5100n2+500

    3. wn=3n52n2+3n72n3n2+2n3

    4. wn=3n3+n252n33n2n+6

    5. wn=n2+3n52n35n2+n+2

    6. wn=2n+5nn+5

    7. wn=2nn2+1+3n

    8. wn=2n33nn4+4n3

    9. wn=n21n+3

    10. wn=ln2n+5n21

    11. wn=sinn2n+1

    12. wn=7n2n

  19. Consider the sequences un and vn whose general terms are given by

    un=1nandvn=32n.

    Calculate limit of the sequence un×vn.

  20. Let wn be a monotonic geometric progression such that w5=125 and w11=1125.

    1. Show that w1=78125 and r=15, where r is the common ratio of wn.

    2. Let Sn be the sum of the first n terms of wn. Calculate limSn.

  21. Consider the sequence un whose general term is given by

    un=3n+1!5n.

    Show that un is a Cauchy sequence.

    Miscellaneous problems

  22. Determine the monotony of the sequence vn as defined below.

    1. vn=12n+1;

    2. vn=n+2n+1;

    3. vn=nen;

    4. vn=nn2+1.

  23. Consider the sequence wn such that

    wn=n!nn

    for all nN.

    1. Show that wn is decreasing.

    2. Justify that wn is convergent.

  24. Using the definition of limit, prove the following equalities.

    1. limn+12n+1=12;

    2. limn41n4+1=1;

    3. limsinnn2=0;

    4. limlnnn=0;

    5. lim1+nnn=e.

  25. Let un be a sequence of real numbers converging to 1. Consider the new sequence vn such that

    vn=un31un1.

    Analyze the convergence of vn.

  26. Show that

    lim12+22++n2n4=0.

  27. Calculate the following limits:

    1. limn+1n+2n;

    2. limn+1n;

    3. limnn;

    4. lim2nn;

    5. lim1n3+1n+13++12n3;

    6. lim3n;

    7. lim2n+3n4n+7n;

    8. lim7n+8n3n+4n;

    9. lim3n+9nn.

  28. Show that, if wn is a sequence of positive real numbers such that limwn+1wn=LR, then limwnn=L.

  29. Prove that, given un a convergent sequence of real numbers, if limun=L, then

    limu1+u2++unn=L.

  30. Consider the sequence vn such that v1=2 and vn+1=2+vn for n2.

    1. Show that vn is an increasing sequence.

    2. Show that vn is bounded from above.

    3. Show that limvn=2.

  31. Show that the sequence wn of positive terms, where w1=2 and wn+1=2wn for nN\1, is convergent. Find its limit.

  32. Show that

    lim1n+1+1n+2++12n+1=+.

  33. Show that the sequence un, where

    un=1n2+1+1n2+2++1n2+n,

    for all nN, is convergent.

  34. Show that

    limn+13n3=0.

  35. Use mathematical induction to show that, nN, we have:

    1. 1+3+5++2n1=n2;

    2. k=1nk3=k=1nk2;

    3. k=1n1kk+1k+2=nn+34n+1n+2.

  36. Consider a convergent sequence un such that un2, for all nN and limun=2. Calculate

    limun+un2+un314un2.

  37. Show that the sequences vn are decreasing, where:

    1. vn=lnn+2n+2;

    2. vn=n9n.

Advertisement

3. Series

This section is devoted to the subject of series, which are essentially generalized infinite sums of numbers. In view of this, it is easy to understand the relevance of having addressed the subject of sequences in the previous section. Indeed, we can say that the following survey on series will be made on the shoulders of all that was presented about sequences of real numbers.

The subject of series plays a major role in mathematical analysis, and it is widely applied to many different areas like economics and finance, statistics, physics, or computer science.

This section is organized as follows. We start by presenting, on Section 3.1, the basic definitions and results involving series, including convergence. Then, from Sections 3.2 to 3.5 we present some important types of series, including series of non-negative terms, alternating series, function and power series, along with the application to the representation of functions. As in the previous section, we propose, in Section 3.6, a set of exercises for the reader to grasp their acquired knowledge on this subject, with which we finish this section.

3.1 Basic definitions and convergence

We start this section by defining the concept of series by means of the concept of sequence. Before, however, note that along this text, we will only consider series of real numbers.

Definition 3.1 (Series). Given a sequence of real numbers, un, a series is an expression of the form

k=1+uk=u1+u2+,

where each uk,kN, is simultaneous a term of the series and a term of the sequence un.

The sequence un, of Definition 3.1, is sometimes called the generating sequence of the series, due to its nature.

One can also come to an equivalent definition of series if, from un, we build a new sequence sn, with general term

sn=u1+u2++un,

which is called the sequence of partial sums of un. Note that the value of the series k=1+uk of Definition 3.1, also referred to as the sum of the series, is the same as calculating limn+sn. The next definition is, therefore, natural.

Definition 3.2 (Convergent series). Let un be a sequence and sn be its corresponding partial sums sequence. Then, we say that the series k=1+uk is convergent (divergent) if and only if the sequence sn is convergent (divergent).

Next, we present some examples where Definition 3.2 is put to effective use.

Example 3.1. The series k=1+1 is divergent. Indeed, it is only necessary to observe that the sequence of partial sums of the constant sequence with all terms equal to 1 is given by the general term sn=n and, for this sequence, we have that limsn=limn=+. Since sn is divergent, then the corresponding series k=1+1 is divergent.

Example 3.2. Let’s study the convergence of the series

n=1+12n.

The generating sequence of this series is a geometric sequence with the first term and common ratio both equal to 12. Then, by applying the sum formula for a geometric sequence, the partial sums sequence is given by

sn=12112n112.

Then, we have

limsn=12×lim112n112=lim112n=1.

Therefore, sn is convergent and, consequently, the series n=1+12n is also convergent and we write

n=1+12n=1.

Example 3.3. Consider the series n=1+5n.

Like in Example 3.2, this series also has a generating sequence that is a geometric progression. In this case, the first term and the common ratio are both equal to 5. Thus, we have

limsn=lim515n15=54lim15n=541+=54.=+.

Since sn is divergent, the series n=1+5n is also divergent. Some authors write n=1+5n=+, though.

Next, we present an example of a series whose generating sequence is an alternating sequence.

Example 3.4. Consider the series n=1+1n+1.

For the corresponding sequence, the partial sums are:

s1=1s2=11=0s3=11+1=1s4=11+11=0s5=11+11+1=1s2n1=1s2n=1.

Since lims2n1=1 and lims2n=0,limsn does not exist and, therefore, the series n=1+1n+1 is divergent.

Example 3.5. Let’s analyze the convergence of the series of the form

n=1+rn1,

with rR, also known as a geometric series.

The corresponding sequence of partial sums, sn has its general term given by

sn=k=1nrk1.

We will break down this analysis into different situations.

  1. Let r1. Then, we have the following equalities:

    sn=1+r+r2++rn1E18

    and

    rsn=r+r2++rn,E19

    Subtracting (19) from (18) yields

    sn1r=1rnsn=1rn1r.

    Observe that the convergence of sn, and, therefore, of the series, depends now on the fact that r is smaller or bigger than 1.

    1. If r<1, we have that limrn=0. Then,

      limsn=lim1rn1r=11r.

      Then the series is convergent and its sum is equal to 11r.

    2. If r>1, there are two cases two consider, r>1 and r<1.

      1. If r>1, then limrn=+ and we have

        limsn=lim1rn1r=1+1r=11r=+,

        and we conclude that the series is divergent.

      2. If r<1, then limrn does not exist since limr2n1= and limr2n=+. Therefore,

        limsn=lim1rn1r=1limrn1r

        Does not exist and the series is divergent.

    3. We are missing to consider one single sub-case from the general case r1, that is, the case when r=1. If r=1, then

      limsn=lim11n11=lim11n2.

      Now, since

      lims2n1=lim1111=1

      and

      lims2n=lim111+1=0,

      we conclude that limsn does not exist and the series is divergent.

  2. The final case to consider is, of course, the case when r=1. If r=1, we have

limsn=limk=1n1k=limn=+.

Thus, the series diverges.

Hence, we conclude that the series n=1+rn1 is convergent if and only if r<1 and, in this case, we have

n=1+rn1=11r.

The general case of the geometric series is presented next.

Proposition 3.1. Let a,rR. If r<1, then the geometric series of first term a and common ratio r,

n=1+arn1,

is convergent and its sum is a1r.

In some cases, the properties of the generating sequence allow us to deduce easily an expression for the partial sums sequence.

Example 3.6. Consider the series n=1+nn+1.

Observe that

s1=12s2=12+23s3=12+23+34sn=12+23+34++nn+1=1n+1

Hence, limsn= and, therefore, the series n=1+nn+1 is divergent.

Example 3.7. Consider the series

n=1+1nn+1.

Firstly, note that

1nn+1=1n1n+1.

Then, we have

s1=112s2=112+1213s3=112+1213+1313sn=112+1213+1314+++1n11n+1n1n+1=11n+1.

Hence, limsn=lim11n+1=1 and the series n=1+1nn+1 is convergent with sum equal to 1.

We will now introduce a special kind of series.

Definition 3.3 (Mengoli’s series). A series n=1+un such that un=vnvn+m, with mN, is called Mengoli’s series.

Mengoli’s series is also sometimes referred to as Telescopian series. The convergence of Mengoli’s series is easy to analyze.

Proposition 3.2. A Mengoli’s series n=1+vnvn+m,mN, is convergent if and only if the sequence vn is convergent. Additionally,

n=1+vnvn+m=k=1mvkmlimvn.

Proof. For Mengoli’s series n=1+vnvn+m, we have that the partial sums sequence, sn, is given by

sn=k=1nvkvk+m=k=1nvkk=1nvk+m=k=1nvkk=m+1n+mvk=k=1mvk+k=m+1nvkk=m+1nvkk=n+1n+mvk=k=1mvkk=n+1n+mvk.

Now, let wn be the sequence whose general term is

wn=vn+1+vn+2++vn+m.

Observe that the series n=1+vnvn+m converges if and only if wn converges and, in this case, we have

limwn=limvn+1+vn+2++vn+m=limvn+1+limvn+2++limvn+m=mlimvn.

Hence,

n=1+vnvn+m=limsn=limk=1mvkk=n+1n+mvk=k=1mvklimk=n+1n+mvk=k=1mvkmlimvn.

The particular case of a Mengoli’s series with m=1 is worth considering.

Corollary 3.1. A Mengoli’s series n=1+vnvn+1 is convergent if and only if the sequence vn is convergent. Additionally,

n=1+vnvn+1=v1limvn.

We proceed with some relevant properties involving series convergence.

Theorem 3.1. Let α be a real non null number. Then n=1+un converges if and only if n=1+αun converges.

Proof. Let sn1 and sn2 denote the sequences of partial sums of the series n=1+un and n=1+αun, respectively.

Since

sn2=k=1nαuk=αk=1nuk=αsn1,

then we have that

limsn2=αlimsn1

and, therefore,

n=1+αun=αn=1+un.

Theorem 3.2. Let α and β be non null real numbers. If n=1+un and n=1+vn are convergent series, then the series n=1+αun+βvn is also a convergent series.

Proof. Let sn1 and sn2 denote the partial sums sequences of the series n=1+un and n=1+vn, respectively. Also, since both series are convergent, let limsn1=S1 and limsn2=S2, for some S1,S2R.

We have:

k=1nαuk+βvk=k=1nαuk+k=1nβvk=αk=1nuk+βk=1nvk=αsn1+βsn2.

Hence,

limαsn1+βsn2=αlimsn1+βlimsn2=αS1+βS2.

We conclude that the series n=1+αun+βvn is convergent and that the sum of this series is αS1+βS2 where S1 and S2 are the sum of the series n=1+un and n=1+vn, respectively. □

Note that if we consider two divergent series, n=1+un and n=1+vn, we cannot conclude that the series n=1+αun+βvn is also divergent, for any real numbers α and β. Indeed, we have the series n=1+1n and n=1+1n are both divergent and yet the series n=1+1n1n is clearly convergent to 0.

Theorem 3.3. Let l be a natural number. Then, the series n=1+un and n=l+1+un are either both convergent or both divergent.

Proof. Let sn1 and sn2 denote the partial sums sequences: sn1=k=1nuk and sn2=k=l+1nuk and let S1=limsn1 and S2=limsn2.

If n is a natural number greater than l, we have

sn1=k=1nuk=k=1luk+k=l+1nuk=k=1luk+sn2.

Applying limits, we obtain

limsn1=k=1luk+limsn2E20

From equality (20), we straightforwardly conclude that the series n=1+un and n=l+1+un are both convergent or both divergent, and, in the first case, we have that

S1=k=1lak+S2.

It is a rather common situation to find ourselves with a series at hand for which there is not easy way to obtain an expression for the respective partial sums sequence. In these situations it is important to establish a criterium that allows us to identify convergent series without having to resort to the partial sums sequence. Such a criterium, that is, a necessary condition for the convergence of series, is presented next.

Theorem 3.4. If the series n=1+un is convergent, then limun=0.

Proof. Let n=1+un be a convergent series, such that n=1+un=l,lR, and let sn be its sequence of partial sums.

Since un=snsn1 and limsn=limsn1=l, we have

limun=limsnlimsn1=ll=0.

Example 3.8. Let’s prove that the series

n=1+2n2n+1

is divergent.

Let un be the sequence whose general term is given by

un=2n2n+1.

We have that

limun=lim2n2n+1=lim2n2n1+12n=lim11+12n=11+0=10.

Then, by Theorem 3.4, we conclude that the series n=1+2n2n+1 is divergent.

3.2 Series of non-negative terms

In this section we focus on series n=1+un, such that un0,nN, which are called series of non-negative terms, and we survey the most relevant techniques for studying their convergence.

Theorem 3.5. Let n=1+un be a sequence of non-negative terms. Then n=1+un is convergent if and only if its sequence of partial sums is bounded from above.

Proof. Let n=1+un be a series of non-negative terms with sn being the respective partial sums sequence.

Let n=1+un be convergent with limsn=L. Note that sn+1=sn+un+1 and un+10 by hypothesis. Then, clearly, sn is an increasing sequence and snL, that is, sn is bounded from above.

Conversely, let sn be bounded from above. Since sn is an increasing sequence bounded from above, then it is convergent to the supremum of its range, LR. Then, limsn=L and n=1+un is a convergent series. □

Example 3.9. Let’s show that the Harmonic series, n=1+1n, is divergent by proving that the corresponding partial sums sequence sn is divergent.

We have that

s2n=k=12n1k=k=1n1k+k=n+12n1k=sn+1n+1+1n+2++12nsn+12n+12n++12n=sn+n2n=sn+12

and, therefore,

s2nsn12

Now, suppose that limsn=lims2n=L, with LR. Then, we would have

lims2nlimsn=LL=012,

which is absurd. Therefore, we conclude that such an L does not exist and, accordingly, the sequence sn is divergent. Hence, the harmonic series is also divergent.

Theorem 3.6 (Comparison test). Let n=1+un and n=1+vn be two series of non-negative terms such that unvn,nN. Then we have:

  1. If n=1+vn converges then n=1+un converges.

  2. If n=1+un diverges then n=1+vn diverges.

Proof. Let n=1+un and n=1+vn be two series of non-negative terms such that unvn,nN. Denote by sn1 and sn2 denote the partial sums sequences of the series n=1+un and n=1+vn, respectively.

  1. If n=1+vn is convergent, let limsn2=S, with SR. Then, we have the following inequalities:

    sn1sn2S.

    Since sn2 is an increasing sequence that is bounded from above, we have that sn1 is convergent and, consequently, the series n=1+un is convergent.

  2. If n=1+un is divergent, since it is a series of non-negative terms, then limsn1=+. Therefore, limsn2=+ and the series n=1+vn is divergent.

Theorem 3.6 is usually referred to as the comparison test. Note that it is enough that there exists a mN such that unvn,nm, for Theorem 3.6’s assertions to be valid.

We will now present a couple of examples where we apply Theorem 3.6.

Example 3.10. In this example, we will study the convergence of the series:

n=1+lnn+2n+2.

Firstly, note that the following inequality holds for every natural n:

lnn+2n+2ln3n+2.

Secondly, we have that

n=1+ln3n+2=ln3n=1+1n+2=ln3n=3+1n.

Since n=3+1n is just the Harmonic series starting at the third term, which we already saw on Example 3.9, that it is not convergent, we can conclude that the series n=1+ln3n+2 is also divergent.

Finally, using the comparison test, Theorem 3.6-(b), we can conclude that the series

n=1+lnn+2n+2

is divergent.

Example 3.11. Let’s analyze the convergence of the series:

n=1+3n32n+1.

Observe that

3n32n+13n32n3n32n=3n32n=13n.

Now, since the series n=1+13n converges because it is a geometric series with ratio 13 and absolute value smaller than 1, then, using the comparison test, Theorem 3.6-(a), we conclude that the series

n=1+3n32n+1

is also convergent.

Theorem 3.7 (Cauchy condensation test). Let un be a decreasing sequence of non-negative terms. Then the series n=1+un converges if and only if the series n=0+2nu2n converges.

Proof. Let un be a decreasing sequence of non-negative terms.

  1. Suppose that the series n=0+2nu2n converges.

    Since the sequence un is decreasing, then the following inequalities hold:

    u2+u3u2+u2=21u21u4+u5+u6+u7u4+u4+u4+u4=22u22u8+u9++u15u8+u8++u8=23u23u16+u17++u31216u216u2n++u2n+112nu2n.

    Then, denoting by sn the sequence of partial sums of n=1+un, we can write the following:

    sn=k=1nuk=u1+u2+u3++unu1+u2+u3+u4+u5+u6+u7+u8++u2n+u2n+1++u2n+11u1+2u2+4u4+8u8++2nu2n=n=0+2nu2n.

    We conclude that sn is bounded from above and, since it is an increasing sequence, we can conclude that it converges. Hence, the series n=1+un is also convergent.

  2. Suppose that the series n=1+un is convergent.

    Let sn denote the sequence of partial sums of the series n=0+2nu2n. We have that

    12sn=12k=0n2ku2k=12u1+2u2+4u4+8u8++2nu2n=12u1+u2+2u4+4u8++2n1u2nu1+u2+u3+u4+u5+u6+u7+u8++u2n1+1+u2n1+2++u2n1+u2nn=1+un.

    Therefore, we have that sn2n=1+un, which means that sn is bounded from above and, since it is an increasing sequence, it is convergent. Hence, n=0+2nu2n is also a convergent series.

We will now present a special kind of series.

Definition 3.4 (p-series). For any real number p, a p-series is a series defined as

n=1+1np.

The convergence of a p-series is determined by the value of p as it is shown next.

Theorem 3.8. Let p be a real number. The series n=1+1np is convergent if p>1 and divergent if p1.

Proof. Let p be a real number and consider the p-series n=1+1np. The proof will be separated into different cases.

  1. If p=1, then the p-series corresponds to the Harmonic series which we already know, from Example 3.9, that is divergent.

  2. If 0<p<1, then we consider the following trivial inequalities:

    npn1np1n.

    Therefore, since n=1+1n is divergent, see (1.), by the comparison test, we have that the series n=1+1np is also divergent.

  3. If p is a natural number greater or equal to 2, consider the trivial inequalities

    np>nn11np<1nn1.

    Now, observe that

    n=2+1nn1=n=2+1n+1n1=n=2+1n11n=n=1+1n1n+1,

    which is a Mengoli’s series with sum 1. Then, since n=2+1nn1 is convergent, by the comparison test, we have that n=1+1np is a convergent series.

  4. If p is a real number greater or equal than 2 and not natural, choose k to be the greatest natural number smaller or equal than p. We have that nk<np and, therefore,

    1nk>1np.

    Since the series n=1+1nk is convergent, see 4., applying the comparison test, we conclude that the series n=1+1np is also convergent.

  5. If p is a real number such that p]1,2[, consider the sequence un whose general term is un=1np. Then, we have

    2na2n=2n12np=12npn=12np1=12p1n.

    Since p]1,2[, we have that

    0<12p1<1

    and, therefore, the series 12p1n is convergent, since it is a geometric series with ratio 12p1<1. Finally, by the Cauchy condensation test, we have that the series n=1+1np is convergent

  6. If p=0, then n=1+1np=n=1+1 which is a divergent series because lim1=10 (Theorem 3.4).

  7. If p<0, then n=1+1np=n=1+np, which is a divergent series because limnp=+0 (Theorem 3.4).

Example 3.12. In this example, we will show that the series

n=1+n7n

converges using the comparison test.

As a preliminary observation, since 7n is much bigger than n, as n grows, it is natural to expect that the series will converge. Let us see that, indeed, this happens.

Observe that:

7n=1+6n=j=0nnj1nj6j=n01n60+n11n161+n21n262++nn106n>n31n363=n!3!n3!216=216nn1n23!=36nn1n2>36n22.

Then, we also have that

17n<136n22,

and

n7n<n36nn22,

for all n3.

Now, since the p-series n=1+1n2 converges, by Theorem 3.8, then the series n=3+1n22 and 136n=3+1n22 also converge.

Finally, by the comparison test, we conclude that the series n=1+n7n is convergent.

In view of the fact that we have a good understanding of their natures, the geometric series and the p-series are of great importance in the context of the usage of the comparison test to study the convergence of many series.

We will now present another test that is called the limit comparison test.

Theorem 3.9 (Limit comparison test). Let n=1+un be a series of non-negative terms and n=1+vn be a series of positive terms. If limunvn exists, then the following statements hold.

  1. If limunvn=r, with rR+, then n=1+un converges (diverges) if and only if n=1+vn converges (diverges).

  2. If limunvn=0, then:

    1. if n=1+vn converges, then n=1+un converges.

    2. if n=1+un diverges, then n=1+vn diverges.

  3. If limunvn=+, then:

    1. if n=1+vn diverges, then n=1+un diverges.

    2. if n=1+un converges, then n=1+vn converges.

Proof. Let n=1+un be a series of non-negative terms and n=1+vn be a series of positive terms.

  1. Suppose that limunvn=r, with rR+. Then, there is a mN such that:

    n>mrr2<unvn<r+r2.

    Therefore, for n>m we have the following inequalities

    unvn>r2un>r2vnE21

    and

    unvn<3r2un<3r2vn.E22

    If n=1+un converges, so does the series n=m+1+un. Then, considering inequality (21) and applying the comparison test, we conclude that the series n=m+1+r2vn is convergent and, therefore, n=1+vn is also convergent.

    If n=1+vn is convergent, n=m+1+3r2vn is also convergent. Then, taking into consideration the inequality (22) and applying the comparison test, we conclude that n=m+1+un converges and so does the series n=1+un.

  2. Suppose that limunvn=0. Then, there is a mN such that:

    n>m1<unvn<1.

    This means that, for n>m we have that un<vn.

    If n=1+vn converges, so does the series n=m+1+vn. Then, applying the comparison test, we conclude that n=m+1+un is convergent and, therefore, n=1+un is also convergent. This proves assertion (i.). Assertion (ii.) is just the contrapositive implication of (i.), which is logically equivalent.

  3. Suppose that limunvn=+. Then, there is a mN such that:

n>mun>vn.

If n=1+vn is divergent, n=m+1+vn is also divergent. Then, by the comparison test, we conclude that n=m+1+un is divergent. Therefore, the series n=1+un is also divergent. This proves assertion (i.). Assertion (ii.) is just the contrapositive implication of (i.), which is logically equivalent.

Much of the analysis of a series convergence is done through the medium of the comparison test or the limit comparison test together with the support of series whose nature we are familiar with, such as the geometric series or the p-series. Next, we present some examples.

Example 3.13. In this example, we will show that the series

n=1+n+1n3+n2+n+1

is convergent.

As a preliminary observation, note that if n is sufficiently big, we have that

n+1n3+n2+n+1nn3=1n2.

It is, therefore, natural to compare the series n=1+n+1n3+n2+n+1 to the series n=1+1n2 which is a convergent p-series, since p=2>1.

Considering un=n+1n3+n2+n+1 and vn=1n2, we have:

limunvn=limn+n+1n3+n2+n+11n2=limn3+n2n3+n2+n+1=limn31+1nn31+1n+1n2+1n3=lim1+1n1+1n+1n2+1n3=1+01+0+0+0=1>0.

Then, applying the limit comparison test, we conclude that the series n=1+1n2 and n=1+n+1n3+n2+n+1 have the same nature of convergence and, therefore, the series n=1+n+1n3+n2+n+1 is convergent.

Example 3.14. Let’s show that the series

n=1+nn2+n+1

is divergent.

Consider the sequence un whose general term is

un=nn2+n+1

and the sequence vn defined by vn=1n. Then, we have:

limunvn=limnn2+n+11n=limn+n2n2+n+1=limn+n2n21+1n+1n2=limn+11+1n+1n2=11+0+0=1>0.

Now, since the series n=1+1n diverges, it’s the harmonic series, then, by the limit comparison test, we conclude that n=1+nn2+n+1 is a divergent series.

Example 3.15. In this example, we will again apply the limit comparison test to show that the series

n=1+n2nn5+1

is convergent.

In this case, we consider the sequences un defined by

un=n2nn5+1

and vn defined by vn=1n4. Then, we have:

limunvn=limn2nn5+11n4=limn52nn5+1=limn5n52n1+1n5=lim12n1+1n5=1+=0.

Since n=1+1n4 is a convergent series, because it is a p-series with p=4>1, then, by the limit comparison test, we conclude that the series n=1+n2nn5+1 is also convergent.

Note that we could have also used the comparison test to prove the convergence of n=1+n2nn5+1. Indeed, since we have, for all nN, that

2nn5+1>n512nn5+1<1n5n2nn5+1<nn5n2nn5+1<1n4

and since n=1+1n4 is convergent, as we already observed, then, by the comparison test, we have that n=1+n2nn5+1 is convergent.

We will now present another test for verifying the convergence of series of positive numbers.

Theorem 3.10 (Ratio test). Let n=1+un be a series of positive terms. If limun+1un exists, then the following statements hold.

  1. If limun+1un is a real number smaller than 1, then n=1+un is a convergent series.

  2. If limun+1un is a real number greater than 1, then n=1+un is a divergent series.

  3. If limun+1un=1+, then n=1+un is a divergent series.

  4. If limun+1un=1, then no conclusion can be drawn about the convergence of n=1+un.

  5. If limun+1un=+, then n=1+un is a divergent series.

Proof. Let n=1+un be a series of positive terms.

  1. Let limun+1un=r, with r being a real number smaller than 1. Then, there is a mN such that

    n>mun+1un<rr12=r+12.E23

    Since 0<r<1, we have that r+12<1. Let s=r+12. We can, then, rewrite (23) as

    n>mun+1un<s,

    which means that, for n>m, we have:

    um+2um+1<sum+2<um+1s.

    We can, therefore, deduce the following inequalities:

    um+2<um+1sum+3<um+2s<um+1s2um+4<um+3s<um+1s3um+k+1<um+ks<um+1sk

    Hence, we can conclude that:

    un<um+1snm1=um+1sm+1sn,

    for all nm+2.

    The series n=m+2+um+1sm+1sn is a convergent series since it is a geometric series with ratio s and such that s<1. Then, by the comparison test, we conclude that the series n=m+1+un is convergent and, therefore, the series n=1+un is also convergent.

  2. Let limun+1un=r, with r being a real number bigger than 1. Then, there is a mN such that

    n>mun+1un>rr12=r+12.E24

    Since r>1, we have that r+12>1. Let s=r+12>1. Then, we can rewrite (24) as

    n>mun+1un>s,

    which means that, for n>m, we have

    un+1un>sun+1>sun.

    We can, therefore, conclude that, for n>m, we have un+1>un. But then, the following statements are true:

    um+2>um+1um+3>um+2>um+1um+k+1>um+k>>um+1.

    Hence, um+k+1>um+1,kN. Then un>um+1>0 for all nm+2. But then limun>0 and, therefore, the series n=1+un diverges.

  3. Let limun+1un=1+. Then, there exists a mN such that for n>m we have

    un+1un>1.

    But that means that un>um,n>m. Then, limun0 and, therefore, the series n=1+un is divergent.

  4. In the case that limun+1un=1 nothing can be concluded about the convergence of the series n=1+un. To show this, we will present two examples.

    On the one hand, if we consider the divergent harmonic series n=1+1n, we have that un=1n and

    limun+1un==lim1n+11n=limnn+1=limnn1+1n=lim11+1n=1.

    On another hand, if we consider the convergent series n=1+1n2, we have un=1n2 and

    limun+1un=lim1n+121n2=limn2n+12=limn2n21+2n+1n2=lim11+2n+1n2=1.

    Thus, nothing can be concluded if limun+1un=1.

  5. Let limun+1un=+. Then, there is a natural number m such that

    n>pun+1un>1.

    Therefore, we have that un+1>un for n>m. But this means that um+2>um+1,um+3> um+2,,um+k+1>um+k,kN, and, therefore, we conclude that un>um+1 for n>m+1. Thus,

    limunum+1>0.

    Hence, limun0 and n=1+un is divergent.

The ratio test is often referred to as the d’Alembert criterion in honor of the French mathematician Jean-Baptiste le Rond d’Alembert, who is credited for publishing it first. Some applications are presented next.

Example 3.16. In this example, we will study the convergence of the series

n=1+n35n.

We start by considering un to be the sequence whose general term is

un=n35n.

Now, applying the ratio test, we have

limun+1un==limn+135n+1n35n=limn+135n3=limn31+1n35n3=lim1+1n35=15<1.

Therefore, we conclude that the series n=1+n35n converges.

Example 3.17. In this example, we will again apply the ratio test in order to prove that the series

n=1+n!n2

is divergent.

Let un be the sequence defined by

un=n!n2.

Then, we have

limun+1un==limn+1!n+12n!n2=limn+1!n!n2n+12=limn+1n2n21+1n2=limn+111+1n2=+×1=+.

Then, by the ratio test, we conclude that the series n=1+n!n2 diverges.

Example 3.18. In this example, we will use the ratio test to show that

limn2n=0.

Let un be the sequence defined by

un=n2n.

Then, we have

limun+1un=limn+12n+1n2n=limn+1n2n2n+1=limn+1n12=12limn1+1nn=12lim1+1n1=12lim1+1n=12<1.

Therefore, by the ratio test, we conclude that the series n=1nn2n converges and, therefore, we have that limn2n=0.

We will now present another test for the convergence of sequences that is known as Cauchy’s root test or simply the root test.

Theorem 3.11 (Root test). Let n=1+un be a series of non-negative terms such that there is an order from which all terms of the series are positive. If limunn exists, then the following statements are true.

  1. If limunn is a real number smaller than 1, then n=1+un is a convergent series.

  2. If limunn is a real number greater than 1, then n=1+un is a divergent series.

  3. If limunn=1+, then n=1+un is a divergent series.

  4. If limunn=1, then nothing can be concluded about the convergence of n=1+un.

  5. If limunn=+, then n=1+un is a divergent series.

Proof. Let n=1+un be a series of non-negative terms such that there is an order from which all terms of the series are positive.

  1. Let limunn=r, where r is a real number smaller than 1. Then, there is a natural number m such that

    n>munn<rr12=r+12.

    Let s=r+12. note that s is such that 0<s<1. Then, for n>m, we have

    unn<sun<sn.

    Hence, we have the following inequalities:

    um+1<sm+1um+2<sm+2um+k<sm+k,kNun<sn,nm+1.

    Now, since the series n=1+sn converges, as it is a geometric series with a ratio smaller than 1 in absolute value, then the series n=m+1+sn also converges. Then, by the comparison test, we conclude that the series n=m+1+un is convergent and, therefore, the series n=1+un is also convergent.

  2. Let limunn=r, where r is a real number bigger than 1. Then, there is a natural number m such that

    n>munn>rr12=r+12.

    Let s=r+12. Note that s is such that 1<s<unn. Therefore, we have, for n>m, that un>sn. But then we have un>1 for n>m and, therefore, limun1, that is, limun0, which, in turn, implies that the series n=1+un is divergent.

  3. Let limunn=1+. Then, there is a natural number m such that

    n>munn1.

    Hence, for n>m, we have that un>1. Then, we can conclude that limun0 and, therefore, the series n=1+un is divergent.

  4. In the case that limunn=1 nothing can be concluded about the convergence of the series n=1+un. To show this, we will present two examples.

    We first consider the series

    n=1+nn+1n.

    In this case, un is the sequence defined by

    un=nn+1n.

    The series n=1+nn+1n is divergent since we have

    limun=limnn+1n=limnnn+1n=limnnnn1+1nn=lim11+1nn=1lim1+1nn=1e0.

    Yet, the root test yields:

    limunn=limnn+1nn=limnn+1=limnn1+1n=lim11+1n=1.

    Now, we consider the series

    n=1+n+1n3.

    In this case, un is the sequence whose general term is

    un=n+1n3.

    In the case of this series, the root test also yields 1:

    limunn=limn+1n3n=limn+1nn3n=limn+1nnn3=limnn1+1nnnnnn2=lim1+1nnnn2=1.

    But the series n=1+n+1n3 is a convergent series. Indeed, considering vn to be the sequence defined by vn=1n2, we have

    limunvn=limn+1n31n2=limn2n+1n3=limn31+1nn3=lim1+1n=10.

    Since the series n=1+1n2 is convergent, it is a p-series with p=2>1, then, by the limit comparison test, we conclude that n=1+n+1n3 is also a convergent series.

    Thus, nothing can be concluded if limunn=1.

  5. The proof of this assertion is analogous to the one produced for (b).

Next, we present a couple of examples where we apply the root test.

Example 3.19. Consider the series

n=1+1+2nnn.

Let un be the sequence defined by

un=1+2nnn.

Then, we have

limunn=lim1+2nnnn=lim1+2nn=limn1n+2n=lim1n+21=lim1n+2=2>1.

Then, by the root test, we conclude that n=1+1+2nnn is a divergent series.

Example 3.20. Consider the series

n=3+1lnnn.

Let un be the sequence whose general term is

un=1lnnn.

Then, we have

limunn=lim1lnnnn=lim1lnn=0<1.

Then, by the root test, we conclude that the series n=1+1lnnn is convergent.

It can be proved that the root test is stronger than the previously presented ratio test in the sense that whenever the ratio test is conclusive, so is the root test, but the converse is not true.

Proposition 3.3. Let n=1+un be a series of positive terms. If the ratio test proves that n=1+un is convergent, then the root test also proves that n=1+un is convergent.

Proof. Let n=1+un be a series of positive terms.

If the ratio test proves that n=1+un is convergent, then we have that limun+1un=r, where r is a real number smaller than 1. Then, there is a natural m such that:

n>mun+1un<rr12=r+12.

Let s=r+12. Note that 0<s<1. Then, for n>m, we have:

un+1un<sun+1<uns.

We can, therefore, write the following inequalities

um+2<um+1sum+3<um+2s<um+1s2um+k+1<um+1sk,kN.

This means that, for n>m, we have

un<um+1snm1.E25

From (25), we can deduce the following:

unn<um+1snm1n=um+1n×snm1n=um+1n×s1m+1n.

Applying limits to both sides of the inequality unn<um+1n×s1m+1n, we obtain

limunn<limum+1n×s1m+1n=1×s10=s.

We can, therefore, conclude that limunn<1 which means that the root test also proves that n=1+un is a convergent series. □

Proposition 3.3 shows that whenever the ratio test provides information on a series convergence, the root test will also provide that information. Additionally, there are some examples of series for which we can determine its convergence with the root test but not with the ratio test. We present one of those next.

Example 3.21. Consider the series n=1+un, where un is the sequence defined by

un=13n+1,ifnisodd13n,ifnis  even.

If we apply the ratio test, we obtain

limun+1un=lim13n+113n=lim3n3n+1=13

if n is even and

limun+1un=lim13n13n+1=lim3n+13n=3

if n is odd, which means that the limit does not exist rendering the test inconclusive.

However, if we apply the root test, we obtain

limunn=lim13nn=13

if n is even and

limunn=lim13n+1n=lim13n+1n=lim13n+1n=lim131+1n=lim13×31n=lim13×1=13

if n is odd. Thus, limunn=13<1 and we conclude that the series n=1+un is convergent.

Note that it is not our intention, with Proposition 3.3 and Example 3.21, to suggest that the ratio test is useless. Despite the root test being a stronger test, the ratio test, as some of our previous examples show, proves to be, in many situations, easier to apply than the root test. Thus, its pertinence is justified.

We will now present the last test in this section that can be applied to series of positive terms.

Theorem 3.12 (Integral test). Let f:[1,+[R+ be a monotonically decreasing and integrable function and un be a sequence such that un=fn,nN. Then, the series n=1+un is convergent if and only if the improper integral 1+fxdx converges.

Proof. Let f:[1,+[R+ be a monotonically decreasing function and un be a sequence such that un=fn,nN. Also, let sn denote the sequence of partial sums of n=1+un and vn denote the sequence whose general term is

vn=1nfxdx.

Before we start the proof itself, we need to deduce some preliminary relations between the terms of the sequences sn and vn. Since f is a decreasing function, we have, for any natural m, that

fm+1mm+1fxdxfm

which can be rewritten as:

um+1mm+1fxdxum.E26

Applying the principle of (26) recursively, we obtain:

u2+u3++unvnu1+u2++un1.E27

From (27), we conclude two things: that

vnsn1E28

and that

u2+u3++unvnu1+u2+u3++unu1+vnsnu1+vn.E29

We are now in conditions to prove Theorem 3.12. Suppose that n=1+un is a convergent series. Then, sn is a convergent sequence. Since sn converges, then, by (28), the sequence vn is bounded from above. Note that vn is an increasing sequence because f is a positive function. Since vn is increasing and is bounded from above we can conclude that vn is also a convergent sequence. Thus, the improper integral 1+fxdx converges.

Conversely, suppose that 1+fxdx converges. Then, vn is a convergent sequence. Then, by (29), we conclude that sn is bounded from above and, therefore, since it is an increasing sequence, it is also convergent. □

Note that the integral test can be straightforwardly generalized for functions defined in the interval [a,+[,aN. The integral test proves to be very useful for series whose general term converts immediately into a decreasing and positive function defined in the interval [1,+[ whose primitive is known. For instance, the integral test can be used to easily determine the nature of convergence of the series n=1+1np, with pR,1nlnn or n=1+1n2+1, as we will see in the following examples, with which we finish this section.

Example 3.22. Let us show, using the integral test, that the p-series n=1+1np, with pR, is convergent if p>1 and divergent if 0<p1.

Let p be a real number greater than 0 and f the function defined by

fx=1xp.

We start by observing that f is a continuous, positive, and decreasing function in the interval [1,+[. Indeed, since

fx=xp=pxp1=pxp+1

then we conclude that fx<0,x[1,+[, and, therefore, f is decreasing on the interval [1,+[.

Suppose that p>1. Then, we have

1+fxdx=limb+1b1xpdx=limb+1bxpdx=limb+xp+1p+11b=limb+1p+1bp11p+1=0+1p1=1p1.

Therefore, we conclude that the improper integral 1+fxdx is convergent and, by the integral test, the series n=1+1np is convergent when p>1.

Now, suppose that p is such that 0<p<1. In this case, we have

1+fxdx=limb+1b1xpdx=limb+1bxpdx=limb+xp+1p+11b=limb+1p+1bp11p+1=++1p1=+.

So, in this case, the improper integral 1+fxdx diverges and, therefore, the series n=1+1np, with p]0,1[, is divergent by the integral test.

Finally, let’s suppose that p=1. In this case we have

1+fxdx=limb+1bfxdx=limb+1b1xdx=limb+lnx1b=limb+lnbln1=limb+lnb=+.

Similarly to the last case studied, the improper integral 1+fxdx diverges and, therefore, the integral test allows us to conclude that the series n=1+1np, with p=1, is divergent.

Example 3.23. In this last example, we will analyze the convergence of the series

n=1+1n+2lnn+2

using the integral test.

We start by considering the function f defined by

fx=1x+2lnx+2,

for x[1,+[.

Note that fx>0,x[1,+[. Additionally, we have

fx=x+2lnx+21=1(x+2lnx+22lnx+2+x+2x+2=1(x+2lnx+22lnx+2+1

and, therefore, fx<0,x[1,+[ and f is a decreasing function in the interval [1,+[.

Now, since

1+1x+2lnx+2dx=limb+1b1x+2lnx+2dx=lnlnx+21b=limb+lnlnb+2lnln3=+lnln3=+,

we have that the improper integral 1+1x+2lnx+2dx diverges and, therefore, by the integral test, we conclude that the series n=1+1n+2lnn+2 is divergent.

3.3 Alternating series

In this section, we will focus on the study of the convergence of alternating series. Informally, an alternating series is just a series whose consecutive terms have opposite signs. The formal definition is as follows.

Definition 3.5 (Alternating series). An alternating series is a series of the form

n=1+1n1un,

where un is a sequence of positive terms.

The study of the convergence of alternating series often starts with a test most commonly known as the alternating series test.

Theorem 3.13 (Alternating series test). Let n=1+1n1un, with un being a sequence of positive terms, be an alternating series. If un is a decreasing sequence and limun=0, then the alternating series n=1+1n1un is convergent.

Proof. Let un be a decreasing sequence of positive terms such that:

limun=0.

Since un is decreasing, we have that un>un+1,nN. Let sn denote the sequence of partial sums of the series n=1+1n1un. Now, consider the sequence s2n. We have:

s2=u1u2;s4=u1u2+u3u4s6=u1u2+u3u4+u5u6s2n=u1u2+u3u4+u5u6++u2n1u2n

and, therefore, we conclude that

s2<s4<s6<<s2n2<s2n,

that is, s2n is an increasing sequence.

Additionally, we have

s2=u1u2<u1s4=u1u2u3u4<u1s6=u1u2u3u4u5u6<u1s2n=u1u2u3u4u5u2n2u2n1u2n)<u1.

This means that s2n<u1, for all nN, and, therefore, we conclude that s2n is a convergent sequence since it is an increasing sequence bounded from above by u1.

Now observe that, since

s2n=u1u2+u3u4+u5u6++u2n1u2n

and

s2n1=u1u2+u3u4+u5u6++u2n3u2n2+u2n1,

we have that s2n=s2n1u2n. Then, we have

lims2n=lims2n1limu2n

and since limun=0 we conclude that

lims2n=lims2n1.

We can, therefore, conclude that the sequence of partial sums sn converges and, thus, the series n=1+1n1un is convergent. □

The alternating series test is normally credited to the philosopher and mathematician Gottfried Leibniz and, therefore, it is frequently referred to as Leibniz’s test, Leibniz’s rule, or even the Leibniz criterion.

When dealing with an alternating series, it is a good strategy to study the series of the absolute values of the terms as there is a relationship between the convergence of these two series. That relationship is observed in the next result.

Theorem 3.14. If the series n=1+un is convergent, then the series n=1+un is convergent.

Proof. Let n=1+un be a convergent series.

Firstly, observe that

ununun0un+un2un,

for all nN.

Now, note that n=1+2un is a convergent series due to our hypothesis that the series n=1+un converges.

Then, by the comparison test, we have that the series n=1+un+un is convergent.

Finally, observe that the series n=1+un can be written as

n=1+un=n=1+un+unun=n=1+un+unn=1+un.

Then, since n=1+un is equal to the difference between two convergent series, we conclude that it is also convergent. □

Given a certain series, Theorem 3.14 exhibits the importance of the convergence of the series of the absolute values of the terms to conclude on the convergence of the original series. Also, since the result contains only one implication and not an equivalence, it points out to a certain division in the series’ universe between the convergent series whose series of absolute values also converge and those convergent series whose series of absolute values do not converge. Such a division is clarified with the next definition.

Definition 3.6 (Absolute and conditional convergence). A series n=1+un is said to be:

  1. absolutely convergent if n=1+un converges.

  2. conditionally convergent if n=1+un converges but n=1+un diverges.

Next, we present a couple of examples.

Example 3.24. Consider the series

n=1+1n1n.

Let un be the sequence whose general term is un=1n. We have that un is a decreasing sequence of positive terms and that limun=0. Then, by the alternating series test, the series n=1+1nun converges.

The corresponding series of absolute values of n=1+1n1n is the series n=1+1n which is the already studied divergent harmonic series.

Therefore, the series n=1+1n1n is conditionally convergent.

Example 3.25. Consider the series

n=1+sinn4.

Note that

sinn41n4,

for all nN, and that the series

n=1+1n4

is convergent because it is a p-series with p=4>1. Then, by the comparison test, we conclude that the series

n=1+sinn4

is convergent and, therefore, we can conclude that the series

n=1+sinn4

is absolutely convergent.

The ratio and root tests presented in Section 3.2 for series of non-negative terms can be straightforwardly adapted for series with terms with variable sign, including series of alternating terms, as we present next.

Theorem 3.15 (Generalized ratio test). Let n=1+un be a series such that there is a mN for which we have un0 for all n>m. If limun+1un exists, then the following statements are true.

  1. If limun+1un is a real number smaller than 1, then the series n=1+un is absolutely convergent.

  2. If limun+1un is a real number greater than 1, then the series n=1+un and n=1+un are divergent.

  3. If limun+1un=1+, then the series n=1+un is divergent.

  4. If limun+1un=1, then nothing can be concluded about the convergence of the series n=1+un and n=1+un.

  5. If limun+1un=+, then the series n=1+un and n=1+un are divergent.

Theorem 3.16 (Generalized root test). Let n=1+un be a series of real numbers. If limunn exists, then the following statements are true.

  1. If limunn is a real number smaller than 1, then n=1+un is absolutely convergent.

  2. If limunn is a real number greater than 1, then n=1+un is divergent.

  3. If limunn=1+, then n=1+un is divergent.

  4. If limunn=1, nothing can be concluded about the convergence of the series n=1+un.

  5. If limunn=+, then n=1+un is divergent.

In the following paragraphs we present two examples with which we finish this section.

Example 3.26. In this example we will study the convergence of the series

n=1+n!2nxn,

where x is a real number.

Let unx be the sequence whose general term is given by

unx=n!2nxn.

Applying the ratio test, we have, for x0,

limun+1xunx=limn+1!2n+1xn+1n!2nxn=limn+1!2n+1xn+12nn!xn=limxn+12=x2limn+1=x2×+=+

and, therefore, we conclude that the series n=1+n!2nxn is divergent for x0.

If x=0, then clearly we have that the series n=1+n!2nxn is convergent with sum equal to 0, since it is an infinite sum of null terms.

Example 3.27. In this example we will study the convergence of the series

n=1+nx1n,

where x is a real number.

Let unx be the sequence defined by unx=nx1n. Then, applying the root test, we have

limunxn=limnx1nn=limnnx1nn=limnnx1=1×x1=x1.

Hence, we have:

  • if x1<1, that is, if x]0,2[, then n=1+nx1n is absolutely convergent;

  • if x1>1, that is, if x],02,+[, then n=1+nx1n diverges.

For x=0 then, we have

n=1+nx1n=n=1+n1n.

In this case, let un be the sequence defined by un=n1n. Now, if we consider the sequences of even and odd terms of un,u2n and u2n1, respectively, then their general terms are u2n=2n and u2n1=12n and we have that limu2n=+ and limu2n1=. Therefore, limn+un does not exist and, consequently, the series n=1+n1n is divergent.

For x=2 we have that

n=1+nx1n=n=1+n

which is divergent since limn+n=+.

Therefore, the series n=1+nx1n is convergent if and only if x]0,2[.

3.4 Function sequences and series

In this section, we aim to study the convergence of a special type of series whose terms are functions themselves. These series have a great importance in several branches of Mathematics such as differential equations or approximation theory. Before, however, we must also study briefly the corresponding sequences of functions which are the generating sequences of those series.

3.4.1 Sequences of functions

A sequence of functions is just a sequence whose terms are functions instead of numbers.

Definition 3.7 (Sequence of functions). A set fnnN, or simply fn, is called a sequence of functions if, for each natural n, fn is a function fn:AR, with A, the common domain of all terms of fn, being a subset of R.

Unlike sequences of real numbers, a sequence of functions can converge in different manners. The first type of convergence, somewhat the most natural one, is called pointwise convergence, and it is presented next.

Definition 3.8 (Pointwise convergence). Let fn be a sequence of functions with common domain AR. Then, fn converges pointwise to the function f:AR if, for all xA, we have:

limn+fnx=fx.

In the study of sequences of functions, we deal with two variables: n, the natural variable of the sequence, and x, the real variable of the functions, which are the terms of the sequences. In view of this and to avoid confusion, along this section, we will write limn+fnx and not limfnx, although all limits will be evaluated when n tends to +.

An example of pointwise convergence is presented next.

Example 3.28. Let fn be the sequence of functions defined as

fn:01Rxxn.

The sequence of functions fn converges pointwise to the function

fx=0ifx11ifx=1.

Indeed, we have

limn+fnx=limn+xn=0=fx

if x[0,1[ and

limn+fn1=1=f1.

Saying that a sequence of functions fn converges pointwise to a function f, that is, that limn+fnx=fx,xA, is equivalent to say that

xA,ε>0,mN:nN:nmfnxfx<ε.

This means that, for each x and ε, we can find an order m from which fnx differs from f(x) no more than ε.

A stronger form of convergence is presented next.

Definition 3.9 (Uniform convergence). Let fn be a sequence of functions with common domain AR. Then, fn converges uniformly to the function f:AR if:

ε>0,mN:nN,xA,n>mfnxfx<ε.

Example 3.29. Let fn be the sequence of functions whose general term is

fn:ARxxn,

where AR.

The sequence fn converges pointwise to the function fx=0, since, for each xR, we have:

limn+fnx=limn+xn=x+=0.

However, uniform convergence of fn is not guaranteed as it depends on the domain A.

If A=01, for instance, then the series of functions fn converges uniformly to the function

f:01Rx0.

Indeed, let ε>0. Then

fnxfx<εxn0<εxn<εn>xε.

Since x01, then

xε1ε.

Therefore, choosing m=1ε+1, for example, we have that

ε>0,mN:nN,n>mfnxfx<ε,x01,

which means that fn is uniformly convergent.

If A=R, nonetheless, it is easy to see that, given an ε>0, we cannot find an m such that fnxfx<ε, for each xR. In fact, if ε=1, we would have

fnxfx<1xn0<1xn<1n>x,

which means that there is not any order m such that fnxfx<1 for each xR. Hence, fn does not converge uniformly, although it is pointwise convergent.

As it follows from the previous definitions, we have the following relationship between these two types of convergence of sequences of functions.

Proposition 3.4. Let fn be a sequence of functions. If fn is uniformly convergent, then fn is pointwise convergent.

Given a sequence of functions fn and a function f, all of them bounded functions on a domain A, consider the sequence of real numbers Mn defined by:

Mn=supxAfnxfx.

Then, the sequence of functions fn converges uniformly to f if and only if limMn=0.

In the cases that, for all nN,fn and f are continuous functions and A=ab,Mn can be defined by

Mn=maxxAfnxfx

and the sequence fn converges uniformly to the function f if and only if limMn=0.

Note that in Example 3.29, when A=01, we have

Mn=supx01fnxfx=maxx01fnxfx=maxx01xn=1n

and, therefore, since limMn=0, the sequence fn converges uniformly to the function fx=0.

Example 3.30. Let fn be the sequence of functions defined by:

fn:[1,+[Rxen2x.

We will show that fn is uniformly convergent to the function f defined by

f:[1,+[Rx0.

Indeed, we have that

Mn=supx[1,+[fnxfx=supx[1,+[en2x0=1en2.

Since

limMn=lim1en2=1+=0,

then the sequence of functions fn converges uniformly to the function f.

Note, however, that if the domain of the functions considered was the set ]0,+[, instead of [1,+[, we would have

Mn=supx]0,+[fnxfx=supx]0,+[en2x0=1

and, therefore, limMn=10. Hence, fn would not be uniformly convergent.

The next result relates the convergence of a sequence of functions with the continuity of the functions involved.

Theorem 3.17. Let A be a subset of R, a be an interior point of A and fn be a sequence of functions with domain A that is uniformly convergent to the function f also of domain A. Then, if for all nN,fn is a continuous function at a, then f is continuous at a.

Proof. Let A be a subset of R, a be an interior point of A and fn be a sequence of functions with domain A that is uniformly convergent to the function f also of domain A.

Suppose that, for each natural n, fn is a continuous function at a. Since the sequence fn converges uniformly to f let m be the natural number such that for n>m we have

fnxfx<ε3,

xA. Let n=m+1. The function fm+1 is a continuous function at a by hypothesis. Then, let δ be a real positive number such that ]aδ,a+δ[A and

fm+1xfm+1a<ε3,

x]aδ,a+δ[. Then, for x]aδ,a+δ[, we have the following:

fxfa<fxfm+1x+fm+1xfm+1a+fm+1afafxfm+1x+fm+1xfm+1a++fm+1afaε3+ε3+ε3ε.

Hence, we conclude that f is continuous at a. □

Example 3.31. Consider the sequence of functions fn such that

fn:01Rxxn,

nN, and the real function f also of domain [0, 1] defined by

fx=0,ifx[0,1[1,ifx=1.

The sequence fn converges pointwise to the function f. Indeed, for x[0,1[, we have that

limn+fnx=limn+xn=0

and, for x=1, we have

limn+fn1=limn+1n=1.

However, by Theorem 3.17, fn does not converge uniformly to the function f since all the functions fn are continuous and f is a discontinuous function.

Theorem 3.17 and Example 3.31 provide a powerful criterium for a sequence of continuous functions not being uniformly convergent to a function. If that function is not continuous, then uniform convergence does not occur.

Example 3.32. Consider the sequence of functions fn such that

fn:[0,+[Rxx1+nx,

nN. Let us show that fn converges uniformly to the function f of domain [0,+[, such that fx=0.

Firstly, we note that

limn+fnx=limn+x1+nx=x+=0.

Now, since, for x0, we have

x1+nxxnx,

then, for each nN, we have

0fnx1n.

We can, therefore, write that

0fnxfx1n,

x[0,+[, which implies that

ε>0,m=1ε+1,n>mfnxfx<ε,x[0,+[,

that is, fn converges uniformly to f.

Next, we present a necessary and sufficient condition for the uniform convergence of a sequence of functions fn known as the Cauchy criterion.

Theorem 3.18 (Cauchy criterion – version 1). The sequence of functions fn, all of domain A, converges uniformly if and only if:

ε>0,pN:npmpfnxfmx<ε,xA.

Sometimes Theorem 3.18 is presented in the following manner.

Theorem 3.19 (Cauchy criterion – version 2). The sequence of functions fn, all of domain A, converges uniformly if and only if:

ε>0,mN:kN,n>mfn+kxfnx<ε,xA.

Now, we present a result that deals with sequences of Riemann integrable functions. Informally, a Riemann integrable function is just a function for which we can approximate the area under its curve using step functions, that is, rectangles with an equal given length, and that approximation can be as tight as we want. The Riemann integral, named after the German mathematician Bernhard Riemann, is credited to be the first rigorous definition of the integral of a function on a given interval.

Theorem 3.20. Let A be a subset of R and gn a sequence of Riemann integral functions in abA converging uniformly to a function g of domain [a, b]. Then, g is integrable on [a, b] and we have the following equality:

abgxdx=limn+abgnxdx.

Proof. Since it is out of the scope of our book and it requires some more advanced calculus tools, we will skip the proof that g is integrable.

Let A be a subset of R and gn a sequence of Riemann integral functions in abA converging uniformly to a function g of domain [a, b].

Let ε be a positive real number. Since gn converges uniformly to g, then let m be a natural number such that, for n>m we have

gnxgx<εba,

xab. But then we have

abgnxdxabgxdx=abgnxgxdxabgnxgxdxabεbadxεbaabdxεbaxabdxεbabaε.

Since we proved that

mN:n>mabgndxabgxdx<ε,

we conclude that

limn+abgnxdx=abgxdx.

We now present an example of an application of Theorem 3.13.

Example 3.33. In this example, we will show that the sequence of functions fn of domain [0, 1] defined by

fnx=4n2x,ifx014n4n2x+2n,ifx14n12n0,ifx12n1,

converges pointwise to the null function, f, defined on [0, 1] but does not converge uniformly to f.

Let x]0,1]. Next, we choose mN such that m=12x+1. Then, for n>m we have n>12x, which means that, for n>m, we have 12n<x. Therefore, for n>m, we have that fnx=0 and, consequently,

limn+fnx=0=fx.

Now, let x=0. Then, since fn0=0, we have

limn+fn0=0=fx

and, from all of the above, we conclude that fn converges pointwise to f.

As for uniform convergence, observe that

01fnxdx=14

and

01fxdx=010dx=0.

Then, by Theorem 3.13, since

limn+01fnx=140=01fxdx,

we conclude that fn does not converge uniformly to the null function f.

We finish this section with a result on sequences of differentiable functions.

Theorem 3.21. Let fn be a sequence of differentiable functions with domain [a, b]. If the sequence fn, of the derivatives of fn, converges uniformly to the function g of domain [a, b] and if there is a cab such that fnc converges, then the sequence fn converges uniformly to a function f of domain [a, b] such that f is differentiable and, for all xab, we have that

fx=limn+fnx=gx.

Proof. Let xab and cab such that the sequence fnc converges. By Theorem 3.13 we can write that

cxgudu=limn+cxfnudu.

Since g is a continuous function, then we have

cxgudu=limn+fnxfc.

Therefore, we conclude that fn converges to the function f defined by

fx=cxgudu+fc,

xab. Since g is a continuous function, we can, therefore, conclude that fx=gx.

Now, since fn converges uniformly to g, let m1 be a natural number such that, for n>m1, we have

gufnu<εba.

Since fnc converges pointwise to f(c), then let m2 be a natural number such that, for n>m2, we have

fncfc<ε2.

Then, for n>m=maxm1m2, we have

fxfnx=cxgudu+fcaxfnudu+fnc<cxguducxfnudu+fcfnccxguducxfnudu+fcfnccxgufnudu+fcfnccxε2badu+ε2ε2baxc+ε2ε2baba+ε2ε2+ε2ε2.

Therefore, for n>m, we have that fxfnx<ε, for all xab. Hence, fn converges uniformly to f. □

3.4.2 Function series

In the following, we define a function series and present some important results.

Definition 3.10 (Function series). A function series is an infinite sum n=1+fn, where fn is a sequence of functions defined in a set AR.

As we did with series of real numbers, for each function series, we consider the correspondent sequence of partial sums of order n, Sn, defined by: Sn=k=1nfk.

Now, we extend the concepts of pointwise and uniform convergence from sequences of functions to function series.

Definition 3.11 (Pointwise convergence). Let fn be a sequence of functions defined in a set AR and f:AR. The function series n=1+fn is said to be pointwise convergent to the sum function f in a set A, and we write n=1+fn=f, if the corresponding sequence of partial sums Sn is pointwise convergent to f in A, that is, if limSnx=fx,xA.

Definition 3.12 (Uniform convergence). Let fn be a sequence of functions defined in a set AR and f:AR. The function series n=1+fn is said to be uniformly convergent to the function f in A if the corresponding sequence of partial sums Sn is uniformly convergent to f.

An example is presented next.

Example 3.34. Let fn be the sequence of real functions with common domain ]1,1[ such that, for each nN,fnx=xn.

The corresponding sequence of partial sums is:

Snx=k=1nxk1=1+x+x2++xn1=1xn1x.

Hence, for x]1,1[, we have

Sx=limSnx=lim1xn1x=1limxn1x=101x=11x.

Therefore, we conclude that the series n=1+fn is pointwise convergent to the function:

f:]1,1[Rx11x.

Definition 3.13 (Domain of convergence). Let fn be a sequence of functions of domain AR. The set of all elements xA for which the series of real numbers n=1+fnx converges is called the domain of convergence.

Example 3.35. Consider the series of functions defined by

n=1+3nsinx5n.

We start by observing that:

3nsinx5n3nsinx5n3nx5nx3n5nx35n,

for all xR and for all nN.

Now, since the series n=1+x35n converges for all xR, then the series n=1+3nsinx5n is absolutely convergent for any xR. Therefore, the domain of convergence of the function series n=1+3nsinx5n is R.

We continue this survey by presenting two relevant results for determining uniform convergence of function series.

Theorem 3.22 (Cauchy’s uniform convergence criterion for series of functions). Let fn be a sequence of functions with common domain AR. Then, the series of functions n=1+fn is uniformly convergent in A if

ε>0,pN:mN,xA,k>pi=k+1k+mfix<ε.

Theorem 3.23 (Weierstrass M-Test). Let fn be a sequence of functions with common domain AR. Then, if there is a sequence Mn of real non-negative numbers such that

  1. n=1+Mn is convergent and

  2. nN,xA,fnxMn,

the series of functions n=1+fn is absolutely and uniformly convergent in A.

Proof. Let n=1+Mn be a convergent series of real numbers and fn be a sequence of functions with common domain A such that fnxMn, for all xA and all nN.

Let x be an element of A. Since fnxMn, for each nN, then the series n=1+fnx is convergent in A. Thus, n=1+fnx is absolutely convergent for any xA.

Regarding uniform convergence, since n=1+Mn is convergent, then the sequence of partial sums

Sn=k=1nMk

is also convergent and, therefore, it is a Cauchy sequence. Then, for all ε>0, there exists a mN such that, for all n>m and for all kN, we have

Mn+1+Mn+2++Mn+k<ε.

We can, therefore, write:

fn+1x+fn+2x++fn+kxfn+1x+fn+2x++fn+kxMn+1+Mn+2++Mn+k<ε,

for all xA. Hence, by Cauchy’s uniform convergence criterion for series of functions, we can conclude that the series n=1+fn is uniformly convergent. □

Example 3.36. In this example, we will show that the series of functions defined as

n=1+cosnxn3

converges uniformly in R.

Indeed, we have that

cosnxn31n3,

xN.

Now, since n=1+1n3 is convergent, then, by the Weierstrass M-Test, we can conclude that the series n=1+cosnxn3 is uniformly convergent in R.

We finish this subsection by presenting three results that we consider to be important to refer to in the context of this survey on function series.

Theorem 3.24. Let fn be a sequence of continuous functions in the domain abR. If n=1+fn is uniformly convergent to a function f in [a, b], then f is continuous in [a, b].

Theorem 3.25. Let fn be a sequence of integrable functions in the domain abR. If n=1+fn is uniformly convergent to a function f in [a, b], then f is integrable in [a, b] and we have that

abfxdx=n=1+abfnxdx.

Theorem 3.26. Let fn be a sequence of class C1 functions in the domain abR such that:

  1. there is a cab such that the series n=1+fnc is convergent and

  2. the series n=1+fn is uniformly convergent to a function g in [a, b].

Then, the function series n=1+fn is uniformly convergent to a differentiable function f in [a, b] such that f=g on [a, b] and

fx=n=1+fnx,xab.

3.5 Power series

In this section, we will study a particular class of function series that hold a great deal of importance for their applications, specifically regarding function analysis.

3.5.1 Definition and properties

In a very simple way, we can define a power series as an infinite sum like the following:

n=0+anxn=a0+a1x+a2x2++anxn+,

where each ai,iN, is a real number.

The type of power series presented above is normally referred to as a power series centered at 0. A more general definition of a power series centered at any real number x0 is given next.

Definition 3.14 (Power series). A power series centered at x0 is a function series n=0+fn, where each function fn can be defined as

fnx=anxx0n,

with anR,nN0.

Given a power series centered at x0, it is easy to observe that x0 belongs to its convergence domain. However, what can we conclude on the convergence of a power series that we know that is convergent at a certain point x=b? The following result sheds some light on this matter.

Theorem 3.27. Let n=0+anxn be a power series centered at 0, with anR,nN. Then:

  1. if n=0+anxn converges at a non-zero value c, then n=0+anxn converges absolutely at any value x such that x<c;

  2. if n=0+anxn diverges at a certain non-zero value c, then n=0+anxn diverges for any x such that x>c.

Proof. Let n=0+anxn be a power series centered at 0, with anR,nN. Let cR\0.

  1. If the series n=0+ancn is convergent, then we have that

    limn+ancn=0.

    But since the sequence whose general term is ancn is convergent, then it is bounded. Therefore, there exists a real positive number L such that

    ancn<L,nN.

    Now, let bR such that b<c and denote by Snb the sequence of partial sums of the series n=0+anbn. Then, we have:

    Snb=k=0nakbk=k=0nakbk=k=0nakckbkck=k=0nakckbckk=0nLbck=Lk=0nbck

    Hence, we have

    SnbLbc1bcn1bc

    and, consequently,

    SnbLbc11bc.

    Now, since Snb is an increasing sequence of positive terms that is bounded from above, then we conclude that Snb is convergent and, therefore, the series k=0+akbk is convergent which means that the series n=0+anxn is absolutely convergent for any xR, such that x<c.

  2. If n=0+ancn is a divergent series, let dR be such that d>c.

    If n=0+andn is convergent, then, by assertion (1.) proved before, the series n=0+anxn would be absolutely convergent on all values x such that x<d. But, since c<d, this would mean that the series n=0+ancn would be convergent, which is absurd because we are supposing precisely the contrary. Therefore, we conclude that n=0+anxn is divergent for any xR such that x>c.

The assertions of Theorem 3.27 allow us to conclude that:

  • if the power series n=0+anxn is convergent at cR\0, then it is also absolutely convergent in the interval ]c,c[ and

  • if the power series n=0+anxn diverges at a point cR\0, then it is divergent in the union of intervals ],cc,+[.

If there exists a real positive number R such that the power series n=0+anxn converges absolutely in the interval ] − R, R[ and is divergent in the interval ],RR,+[ (where at the points –R and R the convergence of the series needs to be checked), then this interval ] − R, R[ is called the interval of convergence while the number R is called the radius of convergence both regarding the power series n=0+anxn in x. If there is not such a real number, then we define the radius of convergence, R, as +.

Note that, although we studied the particular case of a power series centered at 0, analogous conclusions can be drawn in the general case of a power series centered at any given point x0R.

Next we will show how we can obtain the radius of convergence of power series.

Theorem 3.28. Let an be a sequence of real numbers such that:

mN:nN,n>man0

and n=0+anxx0n be a power series centered at x0R. If limanan+1 exists, then the radius of convergence of this power series, R, is given by

R=limanan+1.

Proof. Let n=0+anxx0n be a power series satisfying the conditions of Theorem 3.28 and suppose that limanan+1 exists.

We will divide this proof in three cases.

  1. If liman+1an=0+, then the power series n=0+anxx0n is absolutely convergent for any x. Indeed, in this case we have:

    liman+1xx0n+1anxx0n=liman+1anxx0n+1xx0n=liman+1anxx0=0xx0=0<1.

    Therefore, by the ratio test, the series n=1+anxx0n is absolutely convergent for any xR and, consequently, the radius of convergence of the power series is R=+. Observe that

    limanan+1=lim1an+1an=10+=+.

    Hence, R=limanan+1.

  2. If liman+1an=+, then we have:

    liman+1xx0n+1anxx0n=liman+1anxx0n+1xx0n=liman+1anxx0=+×xx0=+,

    if xx0. If x=x0, then

    liman+1xn+1anxn=0.

    Therefore, the power series converges only at x=x0. Hence,

    R=limanan+1=1+=0.

  3. If liman+1an=LR, then we have:

liman+1xx0n+1anxx0n=liman+1anxx0n+1xx0n=liman+1anxx0=Lxx0.

Therefore, if

Lxx0<1xx0<1L,

then the power series n=0+anxx0n is absolutely convergent.

Hence, the radius of convergence of the power series is

R=1L=limn+anan+1.

If R is the radius of convergence of a power series n=0+anxx0n in x, then the interval of convergence of the power series is the interval ]x0R,x0+R[. Note that, in general, the interval of convergence and the domain of convergence of the power series are not the same, as the latter may include the extreme points of the former in the case that the power series converges in such points.

Example 3.37. In this example, we will find the domain of convergence of the power series defined as:

n=0+5nxn.

Applying Theorem 3.28, we have:

R=limanan+1=lim5n5n+1=lim15=15.

Hence, the interval of convergence of the given power series is: 1515.

Now let’s analyze the power series for x=15 and x=15.

If x=15, the power series becomes the series n=01 which is divergent since liman=lim1=10.

If x=15, the power series becomes the series n=01n which is also divergent since liman=lim1n does not exist.

Hence, the domain of convergence of the power series n=0+5nxn is the interval 1515, that is, in this case, the domain of convergence corresponds to the interval of convergence.

We can also compute the radius of convergence through the root test presented before.

Theorem 3.29. Let an be a sequence of real numbers such that:

mN:nN,n>man0

and n=0+anxx0n be a power series centered at x0R. If lim1ann exists, then the radius of convergence of this power series, R, is given by

R=lim1ann.

Proof. The proof is similar to the one given for Theorem 3.28. We leave it as an exercise for the reader. □

Example 3.38. In this example, we will determine the domain of convergence of the power series defined by

n=0+x1nnn.

By Theorem 3.29, we have:

R=lim1ann=lim11nnn=lim11n=10+=+.

Therefore, the domain of convergence of the given power series is the interval of convergence ],+[ or simply the set R.

The following two results relate power series convergence with the concepts of differentiation and integration. The proofs are left out because they require some mathematical tools outside the scope of this book.

Theorem 3.30. Let n=0+anxx0n be a power series whose radius of convergence is R>0, convergent to the function f in ]x0R,x0+R[. Then, the power series n=1+nanxx0n1 has radius of convergence equal to R and converges to the derivative of f, f, in ]x0R,x0+R[. Additionally, this series converges uniformly in any closed interval ab]x0R,x0+R[.

A couple of examples are presented next.

Example 3.39. The series n=0+xn converge to the function fx=11x and the radius of this series is R=1. Hence, the series n=1+nxn1 is convergent to the function fx=11x2 in ]1,1[. Therefore, we can write

fx=n=1+nxn1,x]1,1[.

As for the extreme points of the interval ]1,1[, it is easy to conclude that the series n=1+nxn1 is divergent if x=1 or x=1.

Example 3.40. In this example, we will consider the power series defined by:

n=2+xnnn1.

We have that the radius of convergence of this series is R=1 and the domain of convergence is [–1, 1].

However, the series

n=2+nxn1nn1=n=2+xn1n1

with radius of convergence 1 is convergent at x=1 but diverges at x=1. Thus, its domain of convergence is [1,1[.

Theorem 3.31. Let n=0+anxx0n be a power series whose radius of convergence is R>0, convergent to the function f in ]x0R,x0+R[. Then, the power series n=0+anxn+1n+1 has radius of convergence equal to R and converges to the primitive of f, such that its value at x0 is 0, in ]x0R,x0+R[. Additionally, this series is uniformly convergent in any closed interval ab]x0R,x0+R[ and we have

fxdx=n=1+ann+1xx0n+1,x]x0R,x0+R[.

3.5.2 Representation of a function as a power series

Arguably, the most notable application of power series is the representation of functions that are infinitely differentiable at a real number of its domain. By representing a function by its power series, we can easily obtain approximations of that function at a given point with an associated error as small as we want. It was, therefore, natural not to finish this section on power series without summarizing the key aspects on this important application.

Indeed, in some situations, given a function f, it is important to obtain a power series that converges to f at least at a certain domain. Such a power series is said to represent the function f in the domain of convergence.

Before we proceed any further, let’s establish that this representation of a function by a power series is unique. In fact, if we have simultaneously that fx=n=0+anxn and fx=n=0+bnxn for all x in a certain domain D, then we have

n=0+anxn=n=0+bnxnn=0+anbnxn=0E30
a0b0+a1b1x+a2b2x2+E31

Since 0 naturally belongs to the domain of convergence D, if we make x=0 in (31), we obtain:

a0b0=0a0=b0.

Now, differentiating equality (30) established above, with respect to x, we obtain

n=1+nanbnxn1=0a1b1+2a2b2x+3a3b3x2+E32

Again, since 0D, substituting x=0 in (32) we obtain

a1b1=0a1=b1.

Applying this reasoning successively, we can conclude that an=bn,nN, that is, the two developments of f as a power series are equal. A similar conclusion would be drawn if the series were centered in a given non-zero point.

It is known that, if f admits derivatives until the order n+1 in an open interval I=]a,b[, with a<b, centered at x=x0, then f can be written in the following way:

fx=fx0+fx01!xx0+fx02!xx02+fx03!xx03++fnx0n!xx0n+rnx0x=k=0nfkx0k!xx0k+rnx0x,E33

for xI, where rnx0x is a remainder of this development of f given by

rnx0x=fn+1cn+1!xx0n+1,

for some c]x0,x[ if x>x0, or c]x,x0[ if x<x0.

In the cases that we have that

limn+rnax=0,E34

then, applying limits to equality (33), we deduce

limn+fx=limn+k=0nfkak!xak+limn+rnaxfx=n=0+fnx0n!xx0n.E35

The development of the function f presented in (35) is called the Taylor series of the function f at the point x=x0. In the particular case where the series is centered at the point x0=0, it is called a Maclaurin series instead of a Taylor series centered at 0. The names of these series were given after the British mathematicians who developed and applied them in their respective works: the English mathematician Brook Taylor (1685–1731) and the Scottish mathematician Colin Maclaurin (1698–1746).

From what was presented above, it should be clear that the Taylor series represents the function f if and only if equality (34) holds. However, by reading this very brief survey on the Taylor series, one might assume that it is enough for the function f to be infinitely differentiable to be able to be represented by a Taylor series. In general, this is not true, that is, although we can always construct the Taylor series of f if f is infinitely differentiable, the convergence

fx=n=0+fnx0n!xx0n

is not guaranteed. A classic example that enlightens this situation is the function f such that f0=0 and fx=e1x2, if x0. It is easy to see that

fxn=0+fn0n!xn,

for all x0, since that in this case the Taylor series is the null series because all the derivatives of f at 0 are null.

The functions, f, that can be represented by a Taylor series at a certain domain D are, therefore, required to have a “very good behavior” that goes beyond being infinitely differentiable. These functions are usually called analytic which means simply that they can be written, for any point x=x0D, as

fx=n=0+anxx0n

for some coefficients an and with a positive radius of convergence.

As we suggested in the beginning of this subsection, given a Taylor series at x0 that represents a function f on a certain interval I containing x0, we are in conditions to obtain an approximation of f at a point xI by considering a finite partial sum of the Taylor series at x0 up to the order that will provide the desired approximation significance. This partial sum up to the order, let’s say, m is called the Taylor polynomial of order m of f at the point x0, which we denote by Tmx0x, and we can write, for xI:

fxTmx0x=k=0mfkx0k!xx0k=fx0+fx01!xx01++fmx0m!xx0m.

Besides adding summands to the Taylor polynomial by increasing the value of m, we can also improve the quality of the approximation by considering a center of the power series, x0, closer to the x in I on which we are evaluating the function f. By doing so, we make the successive powers xx0k smaller and, consequently, the Taylor polynomial will provide a more accurate approximation to f(x).

Example 3.41. In this example, we will illustrate the approximation technique presented before with the function fx=ex.

For this function, we have the following Taylor series centered at 0:

fx=e0+e0x+e02!x2+e03!x2+=1+x+x22+x36+,

with the first Taylor polynomials of orders 1 to 3 being:

T10x=1+x;T20x=1+x+x22;T30x=1+x+x22+x36.

In Figure 3, we show graphically how these approximations improve as they come closer to the function f with the increase of the order of the polynomial.

Figure 3.

Approximations with Taylor polynomials to the function ex.

Let n=0+anxx0n be a power series with radius of convergence R. Using Theorem 3.30, we know that, if fx=n=0+anxx0n, then the series n=1+nanxx0n1 as also radius of convergence equal to R and

fx=n=1+annxn1,x]x0R,x0+R[.

If we keep applying Theorem 3.30 recursively, we conclude that, for x]x0R,x0+R[, we have

fx=nn1anxbn2;fx=nn1n2anxbn3;fpx=nn1np+1anxbnp,pN.

Therefore, a function fx=n=0+anxx0n, for x]x0R,x0+R[, is infinitely differentiable at x0 and its derivatives are such that fpb=p!ap,pN. We can, therefore, conclude that

ap=fpx0p!,pN.

Hence, we conclude that if fx=n=0+anxx0n, for x]x0R,x0+R[, then the development of f as a Taylor series at the point x0 is precisely fx=n=0+anxx0n and this development is valid in ]x0R,x0+R[, while for x<x0R or x>x0+R the Taylor series is divergent. The points x=x0R and x=x0+R must be objective of a particular analysis.

Next, we present an example where we deduce the Maclaurin series of a function.

Example 3.42. Consider the function f defined by

fx=11x2

for any x belonging to the interval ],1[. Let’s determine the Maclaurin series of f and its domain of convergence.

We know that

11x=1+x+x2++=n=0+xn,

for x]1,1[. The radius of convergence of this series is

R=lim1n=1.

Now, applying Theorem 3.30, we have that

11x=n=1+nxn111x2=n=1+nxn1,

for x]1,1[.

Therefore, the Maclaurin series of f is n=1+nxn1 and this development is valid for x]1,1[. Additionally, since the series n=1+nxn1 is divergent at x=1 and x=1, then the function f can be developed as a Maclaurin series only in the interval ]1,1[.

We finish this section with a list of the most commonly used Maclaurin series presented in Table 2.

f(x)Maclaurin seriesConv. domain
ex1+x+x22!+x33!+=n=0+xnn!R
ln(1 + x)xx22+x33=n=1+1n+1xnn] − 1, 1]
ln(1 − x)xx22x33=n=1+xnn[−1, 1[
11x1+x+x2+x3+=n=0+xn] − 1, 1[
11x21+2x+3x2+=n=1+nxn1] − 1, 1[
1+xn1+nx+nn12!+=k=0+nkxk] − 1, 1[
sin(x)xx33!+x55!=n=0+1n2n+1!x2n+1R
cos(x)1x22!+x44!=n=0+1n2n!x2nR
tan(x)x+x33+2x515+17x7315+π2π2
arcsin(x)x+x36+3x540+=n=0+2n!4nn!22n+1x2n+1[−1, 1]
arccos(x)π2xx363x540=π2n=0+2n!4nn!22n+1x2n+1[−1, 1]
arctan(x)xx33+x55=n=0+1n2n+1x2n+1[−1, 1]
sinh(x)x+x33!+x55!+=n=0+12n+1!x2n+1R
cosh(x)1+x22!+x44!+=n=0+12n!x2nR
tanh(x)xx33+2x51517x7315+π2π2

Table 2.

Most common Maclaurin series.

3.6 Suggested exercises

Convergence of series

  1. Classify, using the definition of convergent series, the following series.

    1. n=1+15n;

    2. n=1+n+2n+1;

    3. n=1+1n21n+12;

    4. n=2+lnn+1n;

    5. n=1+1n+2!1n!.

  2. Classify the following series and, if possible, obtain their sum.

    1. n=1+13n;

    2. n=1+1nn+1;

    3. n=1+12n+1n+1n+2;

    4. n=1+1nn+3;

    5. n=1+n+1n2n+22;

    6. n=1+32n;

    7. n=1+1n+13n+15n;

    8. n=1+1n+1n+2n+3;

    9. n=1+3n+4n+5n5n.

  3. Show that

    1. n=1+12n12n+5=2390;

    2. n=1+n+1nn2+n=1;

    3. n=1+lnn2+2nn+12=ln12.

  4. Obtain the general term of the series n=1+un knowing that

    Sn=nn+1.

  5. Find the general term of the following series (suppose that the terms follow the same rule of formation) and analyze their convergence. Also, in each case where convergence occurs, calculate the series sum.

    1. 2+4+8+16+32+;

    2. 213+19127+;

    3. 113+135+157+;

    4. 1+13+132+133+.

    Series of non-negative terms

  6. Prove Theorem 3.8 using the Integral Test.

  7. Use the comparison test to analyze the convergence of the following series.

    1. n=1+1n2+n+1;

    2. n=1+1n3+1;

    3. n=2+lnn+1n;

    4. n=1+1n+1n+2n+3;

    5. n=1+3n+2n7n+5n;

    6. n=1+1n+1.

  8. Show that the series

    n=1+ln3ln4ln5lnnn+2!

    is convergent.

  9. Use the limit comparison test to analyze the convergence of the following series of real numbers.

    1. n=1+n2+1n4+n2+1;

    2. n=1+n+1nn+n+1;

    3. n=1+3n+4n7n+6n;

    4. n=1+1n2+n+13;

    5. n=1+n2n2+1;

    6. n=1+n3n+2n;

    7. n=1+11n+55n;

    8. n=1+1nn+1n+2(n+3.

  10. Analyze the convergence of the following series by applying the ratio test.

    1. n=1+1n!;

    2. n=1+n5n;

    3. n=1+n+16n;

    4. n=1+5nn2;

    5. n=2+nnn!;

    6. n4ln5n;

    7. n=1+n45nn!;

    8. n=1+5n7n+8n;

    9. n=1+n+35nn+12.

  11. Use the root test to analyze the convergence of the following series.

    1. n=1+nn52n+1;

    2. n=1+nn;

    3. n=1+nn+1n2;

    4. n=1+n22n2+12n;

    5. n=1+n+1nn;

    6. n=1+1lnnn;

    7. n=1+1n+1n.

  12. Analyze the convergence of the following series by applying the integral test.

    1. 3+1nlnn;

    2. n=1+n4n34n1;

    3. n=1+1n2+1;

    4. n=1+n1+n2;

    5. n=1+12n+1;

    6. n=1+1nln4n;

    7. n=1+1n+1lnn+1;

    8. n=1+n3en4;

    9. n=1+n+1en+1;

  13. Find the nature of convergence of the following series analyzing its general term.

    1. n=1+n2+2n+1n2+2n+3;

    2. n=1+n+1n2n;

    3. n=1+n2n2+1;

    4. n=1+nn;

    5. n=1+2nn.

  14. Show that, if the series of positive terms n=1+un converges, then the series

    n=1+n+12n+1un

    also converges.

    Series of alternating terms

  15. Show that the following series are convergent using the alternating series test and study their absolute convergence.

    1. n=1+1nn2;

    2. n=1+1n1n5;

    3. n=1+1nn5n;

    4. n=1+1nnn+12n+1;

    5. n=1+1n3n!;

  16. Study the convergence of each of the following series stating if the convergence is absolute or conditional.

    1. n=1+sin2n+1n;

    2. n=1+1ncos;

    3. n=1+1nn+13n;

    4. n=1+1n+22n3n+1;

    5. N=1+1nn2n4+n2+1;

    6. n=1+sinn4;

    7. n=1+cosnen;

    8. n=1+cosn1n3.

    Function series

  17. In each of the following cases, study if the given sequence of functions fn is pointwise or uniformly convergent on the given domain D.

    1. fnx=x2n,D=],+[;

    2. fnx=11+n4x2,D=23;

    3. fnx=x2+nx,D=[2,+[;

    4. fnx=sinnxn4,D=R;

    5. fnx=x4+1n2,D=R;

  18. Show that the sequence of functions fn, where fnx=xn1xn, is pointwise convergent but not uniformly convergent to the null function on the interval [0, 1].

  19. Show that the sequence of functions fn, where, for each nN we have fn:01R and

    fnx=nn1x,ifx011nnx1,ifx]11n,1]

    is pointwise convergent but not uniformly convergent to the function f such that fx=x,x[0,1[ and f1=0.

  20. Consider the sequence of functions fn such that

    fn:01Rx1xxn.
    1. Show that, for all x01, the sequence of functions fn is pointwise convergent to the function f such that fx=0.

    2. Show that

      maxx01fnxfx=nn+1n1n+1.

    3. Show that the sequence fn is uniformly convergent to the function f presented in (a).

  21. Consider the sequence of functions fn such that

    fn:01Rxn2x21+n4x4.
    1. Show that fn is pointwise convergent to the null function f on [0, 1].

    2. Verify that fn1n=12 and justify why the sequence of functions fn does not converge uniformly to the null function f.

  22. Using the Weierstrass M-test, show that the following series of functions are uniformly convergent on the domain presented in each item.

    1. n=1+1n4+x8,D=R;

    2. n=1+sinnxn4,D=R;

    3. n=1+3nsinxn5n,D=01;

    4. n=1+1nnx+33,D=[1,+[.

  23. Consider the sequence of functions fn defined as fn:01R such that

    fnx=2n2x,ifx112n2nnx1,ifx12n1n0,ifx1n1.

    Show that fn is pointwise convergent but not uniformly convergent to the null function on [0, 1].

    [Suggestion: start by showing that the area between the null function and fn is equal to 12.]

  24. Show that the series

    n=1+2nx21+n4x4

    is uniformly convergent on the interval [0, 1].

  25. Show that if n=1+an is absolutely convergent, then the series of functions n=1+anx4n1+x4n is uniformly convergent on the interval [0, 1].

    Power series

  26. Find the radius of convergence and the interval of convergence of the following power series.

    1. n=1+x1nn!;

    2. n=1+5nxn;

    3. n=1+xnn;

    4. n=1+xnn2n;

    5. n=1+x2nn;

    6. n=1+xnn2.

  27. Show that

    11x2=n=0+n+1xn

    for x<1.

  28. Show that

    6x5x24x1=n=1+5n1xn

    for x<15.

  29. Show that

    ln1+x=n=1+1n1xnn

    for all x]1,1].

  30. Obtain the interval of convergence of the following power series.

    1. n=0+x3n2n;

    2. n=0+4n2n!x1n;

    3. n=1+1nxn;

    4. n=1+n2nx2n;

    5. n=1+nn+1xn.

  31. Find the domain of convergence of the following series.

    1. n=1+1nxn;

    2. n=1+1nxnn;

    3. n=1+xnn!;

    4. n=1+1nn2xn3n;

    5. n=1+n!x1n1352n1!.

  32. Consider the function fx=ex2.

    1. Determine the Maclaurin series of f.

    2. Using the Taylor polynomial of order 4 of the series determined in (a), obtain an approximation of f(2).

  33. Consider the function gx=sin2x8.

    1. Determine the Taylor series of g at x0=4.

    2. Using the Taylor polynomial of order 5 of the series determined in (a), obtain an approximation of g(5).

Advertisement

Solutions to suggested exercises

Sequences

  1. 1.

    1. (a) un=n2n

    2. (b) 156=u13

    3. (c) n21,22,23,24

  2. 2.

    1. (a) v1=145,v2=52,v3=2611,v4=167,v5=3817

    2. (b) No, it is not a term of vn since n=143

    3. (c) vn+1vn=123n+53n+2<0

    4. (d) 2<vn145

    5. (e) n=13

  3. 3.

    1. (a) wn=1nn3n+12

    2. (b) w10=1000121

    3. (c) limwn=+, for even n and limwn=, for odd n

  4. 4.

    1. (a) u1=5,u2=25,u3=85,u4=265

    2. (c) un is increasing un+1un=103×3n+13n>0

    3. (d) limun=+

  5. 5.

    1. (a) 1220221

    2. (b) 1vn1

    3. (c) v1<v2>v3

  6. 6.

    1. (a) wn is increasing

    2. (b) 1wn<5

  7. 7. 10

  8. 8.

    1. (a) v5=913

    2. (b) Yes, it is the term v13

    3. (c) vn is decreasing vn+1vn=82n+52n+3<0

    4. (d) 1<vn15

  9. 9. wn is increasing, since wn+1wn=1n+11n+3>1>0

  10. 10.

    1. (a) un=3n1

    2. (b) 63

    3. (c) 610

  11. 11.

    1. (a) vn+1vn=53

    2. (b) vn+1vn<0

    3. (c) –745

  12. 12.

    1. (a) d=4,wn is increasing

    2. (b) wn=4n8

    3. (c) m = 15

  13. 13.

    1. (a) un=4×51n

    2. (b) un is decreasing

    3. (c) 0<un4

    4. (d) 5159

  14. 14.

    1. (a) vn+1vn=12

    2. (b) vn is decreasing

    3. (c) 0<vn5

    4. (d) 635256

  15. 15. p=52

  16. 18.

    1. (a)

    2. (b) +

    3. (c) +

    4. (d) 32

    5. (e) 0

    6. (f) 0

    7. (g) 12

    8. (h) 0

    9. (i) 3

    10. (j) 0

    11. (k) 0

    12. (l) +

  17. 19. 0

  18. 20.

    1. (b) 3906254

  19. 22. They are all decreasing sequences.

  20. 23.

    1. (a) Show that wn+1wn<0

    2. (b) Show that 0<wn<1n and apply the Squeeze Theorem

  21. 25. Converges to 3

  22. 26. Use the fact that k=1nk2=nn+12n+16

  23. 27.

    1. (a) e1

    2. (b) 0

    3. (c) 1

    4. (d) +

    5. (e) +

    6. (f) 1

    7. (g) 0

    8. (h) +

    9. (i) 9

  24. 31. limwn=2

  25. 36. 17

Series

  1. 1.

    1. (a) Convergent

    2. (b) Divergent

    3. (c) Convergent

    4. (d) Divergent

    5. (e) Convergent

  2. 2.

    1. (a) Convergent; S=12

    2. (b) Convergent; S=1

    3. (c) Convergent; S=32

    4. (d) Convergent; S=1118

    5. (e) Convergent; S=516

    6. (f) Convergent; S=3

    7. (g) Convergent; S=98

    8. (h) Convergent; S=112

    9. (i) Divergent; liman=1

  3. 4. un=1nn+1

  4. 5.

    1. (a) an=2n. The series diverges since liman=+

    2. (b) an=213n1. The series converges and its sum is S=32

    3. (c) an=12n12n+1. The series converges and its sum is S=12

    4. (d) an=13n. The series converges and its sum is S=32

  5. 7.

    1. (a) Convergent

    2. (b) Convergent

    3. (c) Divergent

    4. (d) Convergent

    5. (e) Convergent

    6. (f) Divergent

  6. 9.

    1. (a) Convergent

    2. (b) Divergent

    3. (c) Convergent

    4. (d) Divergent

    5. (e) Divergent

    6. (f) Convergent

    7. (g) Convergent

    8. (h) Convergent

  7. 10.

    1. (a) Convergent

    2. (b) Convergent

    3. (c) Convergent

    4. (d) Divergent

    5. (e) Divergent

    6. (f) Convergent

    7. (g) Convergent

    8. (h) Convergent

    9. (i) Convergent

  8. 11.

    1. (a) Divergent

    2. (b) Divergent

    3. (c) Convergent

    4. (d) Convergent

    5. (e) Divergent

    6. (f) Convergent

    7. (g) Convergent

  9. 12.

    1. (a) Divergent

    2. (b) Divergent

    3. (c) Convergent

    4. (d) Divergent

    5. (e) Divergent

    6. (f) Convergent

    7. (g) Divergent

    8. (h) Convergent

    9. (i) Convergent

    10. 13. (a) Divergent

    11. (b) Divergent

    12. (c) Divergent

    13. (d) Divergent

    14. (e) Divergent

  10. 15.

    1. (a) Absolutely convergent

    2. (b) Conditionally convergent

    3. (c) Absolutely convergent

    4. (d) Absolutely convergent

    5. (e) Absolutely convergent

  11. 16.

    1. (a) Divergent

    2. (b) Divergent

    3. (c) Absolutely convergent

    4. (d) Absolutely convergent

    5. (e) Absolutely convergent

    6. (f) Absolutely convergent

    7. (g) Divergent

    8. (h) Absolutely convergent

  12. 17.

    1. (a) Pointwise convergent

    2. (b) Uniformly convergent

    3. (c) Uniformly convergent

    4. (d) Uniformly convergent

    5. (e) Uniformly convergent

  13. 26.

    1. (a) R=+ and I=],+[

    2. (b) R=15 and I=1515

    3. (c) R=1 and I=[1,1[

    4. (d) R=2 and I=[2,2[

    5. (e) R=1, and I=]1,1[

    6. (f) R=1 and I=11

  14. 30.

    1. (a) I=]1,5[

    2. (b) I=],+[

    3. (c) I=[1,1[

    4. (d) I=]0,4[

    5. (e) I=]1,1[

  15. 31.

    1. (a) D=]1,1[

    2. (b) D=]1,1]

    3. (c) D=R

    4. (d) D=]3,3[

    5. (e) D=]2,2]

  16. 32.

    1. (a) n=0+1nx2nn!

    2. (b) 5

  17. 33.

    1. (a) n=1+1n+122n1x42n12n1!

    2. (d) 1415

Advertisement

Acknowledgments

Luís Vieira was partially supported by CMUP (UID/MAT/00144/2013), which is funded by FCT (Portugal) with national (MEC) and European structural funds through the programs FEDER under the partnership agreement PT2020.

References

  1. 1. Apostol T. Calculus. 2nd ed. Vol. 1. New York: John Wiley & Sons; 1967
  2. 2. Courant R, John F. Introduction to Calculus and Analysis. Vol. 1. New York: Interscience Publishers; 1974
  3. 3. Marsden J. Elementary Classical Analysis. San Francisco: W. H. Freeman and Company; 1974
  4. 4. Nikolsky S. A Course of Mathematical Analysis. Moskow: MIR; 1977
  5. 5. Piskounov N. Calcul Différentiel et Intégral. Moskow: MIR; 1974
  6. 6. Shilov G. Analyse Mathématique. Moscow: MIR; 1973
  7. 7. Smirnov VI. A Course of Higher Mathematics. Vol. 1. Reading: Pergamon Press; 1964
  8. 8. Spivak M. Calculus. 4th ed. Cambridge: Publish or Perish, Inc.; 2008
  9. 9. Stewart J. Calculus. 4th ed. Boston: Brooks/Cole Publishing Company; 1999
  10. 10. Trench W. Introduction to Real Analysis. San Antonio: Pearson Education; 2003

Written By

Luís Vieira

Reviewed: 30 October 2023 Published: 22 May 2024