Conradie Notes
Conradie Notes
Real Analysis
(MAM2000W Module 2RA)
2010
Contents
1 Uncomfortable questions 1
4 Infinite series 86
4.1 Basic definitions and properties . . . . . . . . . . . . . . . . . . . . . 87
4.2 Tests for convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
4.3 Tests for non-negative series . . . . . . . . . . . . . . . . . . . . . . . 104
4.4 Regroupings and rearrangements of series . . . . . . . . . . . . . . . . 120
4.5 Power series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
ii
CONTENTS iii
7 Integration 179
7.1 What is integration? . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
7.2 Partitions, sums and integrals . . . . . . . . . . . . . . . . . . . . . . 181
7.3 When does the integral exist? . . . . . . . . . . . . . . . . . . . . . . 186
Uncomfortable questions
Real analysis presumably has something to do with real numbers, and analysis
suggests that whatever we are going to do, it will not be done superficially. In
fact, real analysis certainly has to do with a deeper understanding of real numbers
and what makes them tick, but it goes beyond that and also looks at real-valued
functions and their properties. You may feel that you have a pretty good idea of
what a real number is, and that you spent a substantial part of your first year
working with real-valued functions when you were doing calculus. And then there
was more of this in the Advanced Calculus module, only there you had to work with
real-valued functions of more than one variable. What more is there to say?
In the rest of this section we ask some questions which will hopefully make you
a little uncomfortable, or at least suggest that there are still some problems to be
solved.
Question 1. What is a real number?
This is easy, you may feel. A real number is a number that is either a rational
number or an irrational number.
Good. But what is a rational number?
Easy, you say. It is a number that can be expressed in the form p/q, where p is an
integer and q a non-zero integer.
Good (for the moment). But what is an irrational number?
Well, those are just all the numbers that are not rational, you say (a little irritated).
Ah, but what are these “numbers” you refer to?
The real numbers, you say, ... and then realise you have come full circle.
Perhaps saying that you know what a real number is was a little premature. But
you have worked with them a lot, and you do think that you have a pretty good feel
1
CHAPTER 1. UNCOMFORTABLE QUESTIONS 2
• • •1 •2 • •
-2 -1 0 3 3 1 2
1•
• • • • •
−2 −1 0 1 x 2
This is not at all such a daft question as it may seem to us nowadays. For a
very long time Greek mathematicians thought that every point on the number line
corresponded to a rational number. But Pythagoras’ theorem contained a threat to
CHAPTER 1. UNCOMFORTABLE QUESTIONS 3
this belief. If we construct a right-angled triangle with the two sides adjacent to the
right angle both equal to the unit length, then this theorem tells us that the length
x of the hypothenuse is such that x2 = 12 + 12 = 2. Hence there is a point on the
number line at distance x from 0.
Question 4. Does the point at a length x from 0 (where x > 0 and x2 = 2) represent
a rational number?
Well, let’s suppose it does. Then we can find two positive integers p and q such
that x = pq and p and q have no factors in common (if they did, we could cancel
the common factor). Since x2 = 2, we get p2 /q 2 = 2, or p2 = 2q 2 . It follows from
this that p2 is an even number. But then p is also an even number. (Prove this;
if you are stuck, look at Appendix A.) There is therefore a positive integer k such
that p = 2k. From this we get 2q 2 = p2 = 4k 2 , or q 2 = 2k 2 . But then q 2 is even,
and hence also q is even. We have shown both p and q are even; this means that
they have the factor 2 in common. But we have assumed that they have no factor
in common. This contradiction means that our assumption that there is a rational
number x such that x2 = 2 was wrong.
In terms of the number line this means that there is a point on the number
line that does not correspond to a rational number. If we think of real numbers as
numbers corresponding to points on the number line, then we now have (at least)
one real number that is not a rational number. Adapting the proof above enables
us to find many more.(You are asked to find one more in the exercises.)
The discovery that there are lengths that do not correspond to rational numbers
came as a profound shock to Greek mathematicians, and lead to a distrust of algebra.
They felt that the only safe way to do arithmetic was to work with lengths corre-
sponding to numbers, and to use geometric constructions to find sums, differences,
products and quotients of numbers. This cautious approach lead to stagnation in
Greek mathematics, and it was in India and the Arab world that the major advances
in algebra were made in the middle ages.
Using geometry to do arithmetic is not such a practical idea. But what use are
numbers to us if we cannot do arithmetic with them? We know how to do arithmetic
with rational numbers, but what about irrationals? Our problem is that our only
“definition” of a real number at the moment is “something that corresponds to a
length” (for positive real numbers), or, more generally, “a point on the number line”.
You have a flash of inspiration: decimal representations!
Question 5. Can’t we define real numbers using decimal representations?
The idea goes something like this: Rational numbers can be represented by ter-
CHAPTER 1. UNCOMFORTABLE QUESTIONS 4
5
= 0.833333 . . .
6
(This was to be expected, since 6 has factors 2 and 3.) But what does the right
hand side mean? If we see it as the same kind of abbreviation as before, it should
stand for
8 3 3 3
0+ + + + +···
10 100 1000 10000
But surely it is impossible to add infinitely many numbers! You can spend the rest
of your life trying to do this addition, and never finish. So how can we claim that 65
is equal to this infinite sum?
Question 7. Can we attach a meaning to an infinite sum?
If we look at the example above, there is a glimmer of hope. You may recognise
3 3 2
100
+ 1000 + 10000 + · · · as an infinite geometric series, and remember that there is
3
a formula for the sum of such a series. The series has first term 100 and common
CHAPTER 1. UNCOMFORTABLE QUESTIONS 5
1
ratio 10
, so the formula gives
3
8 3 3 3 4 100 5
+ + + +··· = + 1 = .
10 100 1000 10000 5 1 − 10 6
It works! The formula seems to have saved the day in this example. But will
it do the same for any rational number? If we use “long division” to find the
decimal representation for any rational number, will we always end up with an
infinite geometric series to which we can apply the formula to find its sum? You may
remember hearing that if we use this method to find the decimal representation of a
rational number, we always get a finite or recurring infinite decimal representation.
Do you know why? (Try to work this one out for yourself.) If we accept that this is
the case, then this recurring decimal representation, when written out in full, will
give us an infinite geometric series. This is nicely illustrated by the rational number
1
7
:
142857
1000000 142857 1
0.142857142857142857 . . . = 1 = = .
1 − 1000000 999999 7
But have you really solved the problem by quoting a formula? And where does
the formula come from? How does the formula get around the fundamental question
of how to add infinitely many numbers?
The problems don’t end there, unfortunately. We introduced decimal represen-
tations in the hope that this would enable us to do arithmetic with real numbers,
that this would give us a more practical way of doing this than geometrical construc-
tions with line segments. The following example illustrates a method often used to
convert recurring decimals to fractions, and suggests that it is business as usual:
Let x = 0.3333 . . .
Then 10x = 3.333 . . .
Hence 9x = 10x − x = 3.333 . . . − 0.333 . . . = 3.
1
and so x = .
3
It looks as if we can do calculations with infinite decimals in the same way as
with finite decimals. But how do we go about calculating 4 × 0.3333 . . .? We
can’t start multiplying on the right, because there is no “rightmost” digit. What
makes this more infuriating is that we know what the answer should be: surely
4 × 0.333 . . . = 4 × 31 = 43 = 1.3333 . . ..
So we do have a problem with doing arithmetic with infinite decimals, and this
problem crops up even in the case of decimals representing rational numbers. (In
CHAPTER 1. UNCOMFORTABLE QUESTIONS 6
this case we at least have an escape route: we can convert them to rationals, do
arithmetic with rationals, and convert back the answer to decimal form.)
In short:
Question 8. How does one do arithmetic with infinite decimals?
In case you may have forgotten, recall that we are still trying to say what a real
number is. The idea we are exploring is that we could think of real numbers as
being represented by finite or infinite decimal “expansions”. We have seen that
there is a problem with infinite decimal expansions: we need to be able to make
sense of infinite sums of (rational) numbers. In the case of recurring decimal ex-
pansions we could still try to hide behind a formula. Using this formula, we can
show that every rational number has a decimal expansion that is either finite, or
infinite but recurring. Conversely, we can show that every finite or recurring infinite
decimal expansion represents a rational number. The idea is that all the other dec-
imal expansions represent irrational numbers. But in the case of the other decimal
expansions, we cannot think of them as infinite geometric series any more (there is
no “common ratio”), so the crutch of the formula falls away. So we are left with:
Question 9. How can we make sense of non-recurring decimal expansions?
We have hit several stumbling blocks in trying to identify real numbers with dec-
imal expansions. The gist of the problem seems to be that we do not understand
“infinite sums” and how to work with them. This does not mean that the problem
is insurmountable, simply that there is work to be done. We’ll spend the first part
of this course trying to sort out these problems.
For the record we should also mention here that even our acceptance of a rational
number as “ a number that can be written in the form p/q, with p an integer and
q a non-zero integer” is problematical. What do we mean by “in the form p/q”?
Are 1/2 and 2/4 different rational numbers? Fortunately these problems can be
resolved, but not without some work. These issues are perhaps best dealt with in a
course on algebra, and so we’ll stick to our rather uncritical acceptance of rational
numbers in this course.
Real analysis is not only about understanding real numbers, but also about
understanding real-valued functions. It is perhaps not surprising that there are
troubling questions that we can ask about them as well. You may recall that there
were a number of rather important theorems in your calculus course that you did
not prove, such as the mean value theorem. We’ll address these omissions in this
module. But there are also new questions that arise when we try to add infinitely
many functions, a situation that arises surprisingly frequently in practice.
CHAPTER 1. UNCOMFORTABLE QUESTIONS 7
Figure 1.3: The function f (x) = ex and its first three Taylor polynomials.
Similar questions could of course be asked for any infinitely differentiable func-
tion, such as the sine and cosine functions for example.
CHAPTER 1. UNCOMFORTABLE QUESTIONS 8
To summarise:
Question 10. Where (if anywhere) does the Taylor series of an infinitely differen-
tiable function equal the function?
Suppose that the Taylor series of an infinitely differentiable function f about
x = 0 equals the function for every x in some interval (−a, a), then we can write
∞
X
f (x) = an xn
k=0
There is nothing that stops us from writing down a “Fourier polynomial of infinite
degree”, or Fourier series for this function. (To be honest, we have not explained
how these polynomials are found. For the moment we ask you to accept that it can
be done.) Once we have done this, we can start asking the same kind of questions
that we asked for Taylor series. We’ll keep it simple:
Question 12.Where (if anywhere) does the Fourier series of a periodic function
equal the function?
We have glossed over some technicalities in the last two questions. But too
much detail would probably have obscured the essence of the problems we raised.
Hopefully there was enough to convince you that there are interesting questions that
can be asked, and that there are no easy answers.
CHAPTER 1. UNCOMFORTABLE QUESTIONS 10
The next problem is easy to state. Suppose we are faced with the problem of
finding the solutions (if there are any) of the equation
x = cos x.
On the left in Figure 1.5 the graphs of the functions f (x) = cos x and g(x) = x
are shown; the solution of the equation above is given by the x-coordinate of the
point of intersection of the two graphs. From the figure it seems to be clear that the
equation has a solution. Unfortunately there is no formula for solving an equation
like this. One possibility is to resort to a “trial and error” method. As a first step,
To get a better “feel” for the method, we look at a second example. This time
we try to solve the equation x = e0.5x − 1. If we take x1 = 2 as our initial guess,
and put xn+1 = e0.5xn − 1, we see from Figure 1.6 that the sequence (xn ) obtained
seems to converge to the solution x = 0 This is not all that useful, since it is easy
to spot this solution immediately. Starting with x1 = 3 leads to a sequence with
terms just getting larger and larger. It is clear from the graph, however, that there
is a solution other than x = 0 to this equation. A little bit of experimentation and
a good look at the graph should be enough to convince you that we’ll end up with
a sequence converging to 0 or a sequence with terms getting larger and larger, no
matter which starting point (other than the solution itself) we choose. It is natural
to ask why the method works for the first equation, but not for the second.
The method can go wrong in other ways as well. If it is used to try to solve the
equation
2
x = 4x2 e−x ,
most starting points lead to a sequence with terms which eventually alternate be-
tween two different values and therefore does not get closer to any one of the solutions
of the equation.
On the left in Figure 1.7 on the next page we see what happens if we start with
x1 = 1, and on the right we see what happens if x1 = 2 is used as a starting point.
CHAPTER 1. UNCOMFORTABLE QUESTIONS 12
2
Figure 1.7: Attempting to solve the equation x = 4x2 e−x .
But even “worse” things can happen: Figure 1.8 shows an orbit with what seems
to be randomly distributed terms filling up an interval. This is known as chaotic
behaviour.
Sequences like these arise in the study of systems that change with time, provided
we only consider the time at discrete time intervals. We call such systems discrete
dynamical systems. All of this suggests our final question:
Question 14. Is there any way in which we can predict the behaviour of a discrete
dynamical system?
In this introductory chapter we have raised many questions, and answered none of
them. If you feel somewhat less comfortable with real numbers now than before you
have started reading, we have achieved our objective. Understanding real numbers
and their properties is far from a trivial exercise. But we hope that the some of the
examples have also shown that it may, in the long run, be a rewarding one.
CHAPTER 1. UNCOMFORTABLE QUESTIONS 14
Historical Notes
Brook Taylor (1685 – 1731)
Taylor was an English mathematician
best known for his discovery of the for-
mula for what is now known as the Tay-
lor expansion of a function. This first
appeared in his book Methodus incre-
mentorum directa et inversa of 1715, in
which he also develops the branch of
mathematics known as the ”calculus of
finite differences”. He is credited with
the invention of integration by parts,
and gave the first systematic account
of the basic principles of perspective.
In 1712 he was elected a Fellow of the
Royal Society and became a member
of the committee who had to settle the
priority dispute on the invention of the
calculus between Newton and Leibniz.
Joseph Fourier (1768 – 1830)
Fourier was born in Auxerre in France,
one of fifteen children. He showed great
interest in mathematics at an early age,
but initially decided to train for the
priesthood. However, he never took re-
ligious vows, but took a job as a teacher
in Auxerre, at the same time doing
mathematical research. It was the time
of the French Revolution and in 1793
Fourier became heavily involved in pol-
itics on the side of the revolutionaries.
His support for one faction within the
revolution led to his imprisonment on
two occasions. In 1794 he became one
of the first students at the newly estab-
lished Ecole Normal in Paris and later
taught mathematics at the College de
France and the Ecole Polytechnique.
CHAPTER 1. UNCOMFORTABLE QUESTIONS 15
Exercises
6. In this question you may assume that every real number is either a rational
or an irrational number, but never both ratinal and irrational.
Prove or disprove the following statements:
In the previous chapter we raised many problems about our understanding of real
numbers. Although we have a number of helpful ways of thinking about real num-
bers, none of them turned out to give us a useful and rigorous definition. This
chapter looks at ways to remedy this unsatisfactory situation.
Let’s start by putting our expectations on the table. As usual, we’ll use Q and
R to denote the sets of rational and the set of real numbers respectively. We would
like to define real numbers in such a way that:
• Q is a subset of R;
• we can do everything with the elements of R that we can do with those of
Q, in particular we can add subtract, multiply and divide them, and compare
them for size;
• R contains more elements than Q (and in particular contains an element x
such that x2 = 2);
• there is a one-to-one correspondence between the elements of R and the points
on the number line;
• there is a correspondence between the elements of R and decimals (recurring
and otherwise).
16
CHAPTER 2. NUMBERS AND SEQUENCES 17
We’ll opt for the axiomatic approach, but will try to soften the blow by taking some
time to motivate the axioms. This will be done in two stages. Since we want the
real numbers to be a set of numbers with all the properties of the rational numbers,
we’ll start by listing the essential properties of the rational numbers. Then we’ll
try to find a property that we feel the real numbers should have, but the rational
numbers does not have.
In the second stage we shall need to use sequences, and use them in a substantial
way. As you will soon see, this chapter about numbers will end up being mostly
about sequences. As luck will have it, a good understanding of sequences will in the
next chapter help us to come to grips with infinite sums as well.
M4 For each x ∈ Q such that x 6= 0, there exists a unique element x−1 ∈ Q with
the property that xx−1 ) = 1 (every non-zero rational number has a unique
inverse with respect to multiplication).
The set of rational numbers with its operations of addition and multiplication
is not the only example of this kind of structure in mathematics. If you are doing
the Introductory Algebra module, you will see others. Any set with operations of
addition and multiplication satisfying A1 – A4, M1 – M4 and D above is called a
field, as you may have seen in the Linear Algebra module already.
In a field we can define two further operations: For x, y ∈ Q,
From the properties listed above many other familiar algebraic properties of the
set of rational numbers can be derived. We shall not do this here, but will use them
freely. Since the properties above are the axioms defining a field, these deductions
can be done for any field. If you do the Introductory Algebra module, you will see
this and more.
We can also compare rational numbers: there is an order relation between ratio-
nal numbers, which we shall denote by the usual symbol < . This relation has the
following properties:
1 1
(c) If a > 0 and b > 0, then a ≤ b ⇔ ≥ .
a b
1 1
(d) If a < 0 and b < 0, then a ≤ b ⇔ ≥ .
a b
(e) If a ≥ 0 and b ≥ 0, then a ≤ b ⇔ a2 ≤ b2 .
Some of you may feel a bit uneasy about the fact that we seem to have taken
for granted the fact that Q with its usual addition and multiplication satisfies the
properties A1 – A4, M1 – M4, D and O1 – O4. If so, you may reassure yourself by
looking in Appendix B for more detail on how this may be proved.
In the rest of this chapter we’ll frequently use the absolute value of a number.
We give the definition for rational numbers below, but the definition makes sense in
any ordered field. Since we’ll be defining the set of real numbers in such a way that
it is also an ordered field, everything we prove here for absolute values of rational
numbers will remain true for absolute values of real numbers as well.
Recall that if x ∈ Q, then |x| = x if x ≥ 0 and |x| = −x if x < 0. We call |x|
the absolute value, or modulus, of x.
It follows almost immediately from this definition that
CHAPTER 2. NUMBERS AND SEQUENCES 20
• | − x| = |x|;
• if x, c ∈ Q and c > 0, then |x| ≤ c ⇔ −c ≤ x ≤ c;
• if x, a ∈ Q, then |x − a| = x − a if x ≥ a and |x − a| = a − x if a > x, and
therefore |x − a| is exactly the distance between x and a on the number line.
Proof: (a) and (b) are left as exercises. [Hint: consider cases.]
We derive (c) from (b):
(c) From (b) we have
|x| = |(x − y) + y| ≤ |x − y| + |y| and |y| = |(y − x) + x| ≤ |y − x| + |x| = |x − y| + |x|.
Combining these two inequalities we get
−|x − y| ≤ |x| − |y| ≤ |x − y|, and so | |x| − |y| | ≤ |x − y|.
The next step is to try to identify what it is that makes the set of real numbers
different from the set of rational numbers. We have to work for the moment with our
rather incomplete understanding of what a real number is. The route we’ll follow
will depend heavily on the use of sequences, and we therefore look at sequences of
rational numbers in the next section.
Summary:
In this section we have introduced the set of rational numbers Q axiomatically as a
set on which the two operation of addition and multiplication are defined, satisfying
certain properties with respect to these operations. Any set with two operations
satisfying these properties is called a field ; the set of rational numbers with its usual
operations is therefore an example of a field. We also listed some properties of the
order relation < on Q. Any field with a relation satisfying these properties is called
an ordered field, so that Q is an example of an ordered field. The section ended with
some important properties of inequalities and absolute values.
CHAPTER 2. NUMBERS AND SEQUENCES 21
Exercises
1. Use the definition of the absolute value to prove that for all x, y ∈ Q,
(a) | − x| = |x|
(b) if c > 0, then |x| ≤ c ⇔ −c ≤ x ≤ c
(c) if c > 0, then |x − y| ≤ c ⇔ y − c ≤ x ≤ y + c
(d) |xy| = |x| |y|
(e) |x + y| ≤ |x| + |y|.
2.2 Sequences
The idea of an infinite sequence of numbers is so natural that it is tempting to do
without a definition altogether. Informally, we think of such a sequence as a list
that can be written in the form
a1 , a2 , a3 , a4 , a5 , . . . ,
where the dots indicate that the numbers continue indefinitely. The subscripts on
the a’s are labels indicating the positions of the numbers in the list; so a3 , for
example is the third number in the list, and more generally ai is the i-th number.
The numbers ai could be real numbers, or rational numbers, or complex numbers, or
anything else; but in this section we will restrict our attention to rational numbers.
We will call a sequence of rational numbers a rational sequence for short. A rigorous
definition of an infinite rational sequence is not hard to give; here it is:
Two remarks are appropriate at this stage. Firstly, since we will be dealing
with rational sequences almost exclusively in this section, we will in the rest of this
section abbreviate “rational sequence” to “sequence”. Secondly, we will work only
with infinite sequences in what follows, so that we may as well drop the qualification
CHAPTER 2. NUMBERS AND SEQUENCES 22
“infinite” as well. This means that if we use the word “sequence” in the rest of this
section, it really means “infinite rational sequence”.
From the point of view of the definition we have given, a sequence should be
denoted by a single letter, like a, and its values by
a(1), a(2), a(3), a(4), . . . ,
but it has become customary to use subscript notation instead. In this notation, if
a : N+ → R is a sequence, we write
a1 , a2 , a3 , a4 , . . . instead of a(1), a(2), a(3), a(4), . . . .
We call an the n-th term of the sequence, and denote the entire sequence by
(an )n∈N+ , (an )∞
n=1 or, more simply, (an ). Note that there is a difference between an
(the n−th term of the sequence) and (an ) (the entire sequence).
Example 2.2.2 Sequences can be defined by giving an explicit formula for the n-th
term of the sequence. Here are some examples:
(d) an = sin nπ
2
, then (an ) = (1, 0, −1, 0, 1, 0, −1, 0, 1, 0, −1, 0, . . .)
Sequences can also be defined recursively. This means that a formula or description
is given for obtaining the n-th term in terms of one or more of the preceding terms.
In such cases we can only find a term once we have obtained the preceding terms.
We illustrate this in the following two examples.
CHAPTER 2. NUMBERS AND SEQUENCES 23
1 •
•
•
• • • • • • • • •
0 1 2 3 4 5 6 7 8 9 10 11 12
Example 2.2.5 Let u1 = 1 and u2 = 1 and un+2 = un + un+1 for all n ∈ N+ . Then
(un ) = (1, 1, 2, 3, 5, 8, 13, . . .).
This is the famous Fibonacci sequence, of which many of you may have heard before.
Note that in this case we need to know the preceding two terms to calculate the
next one.
Example 2.2.6 Let (an ) be any sequence and define the sequence (bn ) by putting,
for each n ∈ N+ , bn = an+3 . Then (bn ) = (a4 , a5 , a6 , . . .), so (bn ) is obtained by
omitting the first three terms of (an ).
For example, if (an ) = ( n1 ), then (an+3 ) = ( 14 , 51 , 61 , . . .).
More generally, for any fixed k ∈ N+ , the sequence (ak+n ) is obtained from the
sequence (an ) by omitting the first k terms of (an ).
There are certain types of sequences that we will frequently use in the rest of this
course. Although the names we use are generally quite good descriptions, we need
to give precise definitions. Such a definition will always be our starting point when
we prove something about a sequence of this type.
CHAPTER 2. NUMBERS AND SEQUENCES 24
Definition 2.2.7
(a) A constant sequence is a sequence with all its terms equal. More formally, a
sequence (an ) is constant if and only if there is a real number c such that an = c for
every n ∈ N+ . Writing this in terms of quantifiers and connectives, we get
(b) An eventually constant sequence is one in which from some term onward,
all its terms are equal. More formally, (an ) is eventually constant if and only if there
is a real number b and a k ∈ N+ such that, for any n ∈ N+ , if n ≥ k then an = b.
Using quantifiers and connectives, this condition reads:
Example 2.2.8
(a) If an = cos 2nπ for every n ∈ N+ , then (an ) is a constant sequence, since an = 1
for every n ∈ N+
41
(b) If bn is the n-th digit in the decimal expansion of 3
, then the sequence (bn ) is
eventually constant, since for n ≥ 3, bn = 6.
Some ideas that you are familiar with from calculus are also useful in the study of
sequences. Increasing or decreasing sequences are very similar to an increasing or
decreasing real-valued functions. Here are the definitions:
Definition 2.2.9 Let (an ) be a sequence of real numbers. We say that (an ) is
Warning: Note that the property of being increasing or decreasing is one that a
sequence either has, or does not have. A sequence other than a constant sequence
cannot be both increasing and decreasing, and no sequence can be both strictly
increasing and strictly decreasing. If the terms of a sequence increases up to a point
and then start decreasing (or the other way round), it is not monotone.
To prove that a sequence (an ) is increasing, you have to prove that an+1 ≥ an for
every n ∈ N+ . To prove that it is not increasing, it is enough to find one n ∈ N+
such that an+1 < an . Similar comments apply to decreasing sequences.
Example 2.2.10
1 1
(a) The sequence ( n1 ) is strictly decreasing, because n < n + 1 ⇒ > for all
n n+1
n ∈ N+ .
(b) the sequence (n!) is strictly increasing, because (n + 1)! = (n + 1)n! > n! for all
n ∈ N+ .
(d) The sequence ((−1)n+1 n) = (1, −2, 3, −4, 5, −6, . . .) is neither increasing nor
decreasing; so is not monotone.
Sequences that do not get out of hand, in the sense that their terms do not get
arbitrarily large positive or negative, are generally easier to work with and more
useful. We call such sequences bounded. The next definition makes this idea precise.
(a) The sequence (an ) is bounded above if and only if there is a rational number
c such that an ≤ c for all n ∈ N+ , or symbolically
(b) The sequence (an ) is bounded below if and only if there is a rational number
c such that an ≥ c for all n ∈ N+ , or symbolically,
(c) The sequence (an ) is bounded if and only if it is both bounded above and bounded
below.
(d) The sequence (an ) is called unbounded if and only if it is not bounded.
Remarks
•
•
• •
• • •
• •
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14
•
•
•
•
c
2. If a sequence has an upper bound c, then every real number larger than c is
also an upper bound for the sequence. This means that a sequence which is
bounded above has infinitely many upper bounds. Similarly, a sequence which
is bounded below has infinitely many lower bounds.
3. A sequence (an ) is not bounded above if it does not have an upper bound or,
equivalently, no real number is an upper bound. Symbolically this becomes
(an )is not bounded above ⇔ (∀c ∈ R)(∃n ∈ N+ )(an > c).
CHAPTER 2. NUMBERS AND SEQUENCES 28
Example 2.2.14
4+3n 4+3n 4
(a) The sequence n
is bounded. To see this, note that n
= n
+3 .
Since n > 0, n4 + 3 > 3, so certainly an = n4 + 3 ≥ 3 for all n ∈ N+ . This shows that
(an ) is bounded below, and that 3 is a lower bound.
Since n1 ≤ 1 and n4 ≤ 4, we have an = n4 + 3 ≤ 7 for all n ∈ N+ . It follows that (an )
is bounded above, and that 7 is an upper bound.
So 3 ≤ n4 + 3 ≤ 7 for all n ∈ N+ , and this sequence is bounded above and below; i.e.
it is a bounded sequence.
n
(b) The sequence (−1) n
= (−1, 12 , − 31 , 14 , − 15 , . . .) is bounded; −1 is a lower bound
1
and 2
is an upper bound.
(c) We show that the sequence (an ) = (n − 10) = (−9, −8, −7, −6, . . .) is bounded
below, but not bounded above. For any n ∈ N+ , n ≥ 1, so n − 10 ≥ −9. So −9 is
a lower bound for the sequence (n − 10). To show that (an ) is not bounded above,
we must show that (∀c ∈ Q)(∃n ∈ N+ ) (an > c). So let c ∈ Q be given. Choose
n ∈ N+ so that n > c + 10. Then an = n − 10 > (c + 10) − 10 = c, as required. It
follows that (an ) is not bounded.
(d) In much the same way it can be shown that the sequence (−n − 10) is bounded
above, but is not bounded below, and hence not bounded.
(e) The sequence (an ) = ((−1)n+1 n) = (1, −2, 3, −4, . . .) is not bounded above, nor
bounded below. We show that it is not bounded above. Let c ∈ Q be given. Let
n ∈ N+ be chosen so that n is odd and n > c. Then an = (−1)n+1 n = n > c, as
required.
If d ∈ Q and m ∈ N+ is chosen to be even and such that m > −d, then am =
(−1)m+1 m = −m < d, and so (an ) is also not bounded below.
(f) In the last example we show how induction can be used to prove that a recursively
defined sequence is bounded. Consider the sequence (cn ) defined by
1 2
c1 = 2, cn+1 = cn + for all n ∈ N+
2 cn
CHAPTER 2. NUMBERS AND SEQUENCES 29
The next result is useful when you need to show that a sequence is bounded,
because it allows you to check one condition, instead of two (bounded above and
bounded below).
Proof: ( ⇒ ) : Suppose (an ) is bounded. Then there exist c1 , c2 ∈ Q so that for all
n ∈ N+ , c1 ≤ an ≤ c2 . Let c = maximum {|c1 |, |c2 |}. We want to show that, for all
n ∈ N+ , |an | ≤ c, i.e. −c ≤ an ≤ c. Now, for any n ∈ N+ :
Summary:
In this section we gave a precise definition of an infinite rational sequence and
introduced the notation we’ll use for such sequences. We looked at different ways of
defining sequences. Special types of sequences were introduced:
CHAPTER 2. NUMBERS AND SEQUENCES 30
• monotone sequences;
• sequences that are bounded above, bounded below and upper and lower bounds
for sequences;
Exercises
(a) (an ) = (9 − n2 )
2n + 2
(b) (bn ) =
3n + 6
n
(c) (cn ) =
n+3
n
(d) (dn ) = 3
2
−n
2. For each of the sequences in Question 1(a) and (b) say whether it is (a) bounded
above (b) bounded below (c) bounded (d) unbounded. Give proofs for your
answers.
4. Define the sequence (an ) by a1 = α and an+1 = a2n − 2an + 2 for all n ∈ N+ .
Suppose that 1 < α < 2. Show that:
Is (an ) bounded?
(a) Show that, if 0 < β < 1, then (bn ) is increasing. Is (bn ) bounded in this
case? Give a proof for your answer.
(b) Show that, if β > 1, then (bn ) is decreasing. Is (bn ) bounded in this case?
Give a proof for your answer.
(c) What happens if β = 0? And if β = 1?
(a) If (an ) is bounded above, then the smallest of the upper bounds of (an ) is called
the supremum, or least upper bound (l.u.b.) of (an ) and denoted by supn an , or
sup{an : n ∈ N+ }.
(b) If (an ) is bounded below, then the largest of the lower bounds of (an ) is called
the infimum, or greatest lower bound (g.l.b.) of (an ) and denoted by inf n an , or
inf{an : n ∈ N+ }.
The definition does not say anything about the existence of the supremum or
the infimum of a sequence, and the example we started with suggests that this may
not be entirely unproblematic.
If you want to impress the snobs, note that supremum and infimum are Latin
words, and their plurals are respectively suprema and infima (rather than supre-
mums and infimums).
The following propositions gives useful characterisations of the supremum and
infimum.
CHAPTER 2. NUMBERS AND SEQUENCES 33
Proof: For (a), (ii) simply says precisely what it means for b to be the smallest upper
bound, so that (i) and (ii) are equivalent. To see that (ii) and (iii) are equivalent,
note that the second part of (iii) is simply the contra-positive of the second part of
(ii). The proofs for (b) are similar.
It is time for some examples. First we prove two simple but useful results about
the rational numbers.
Lemma 2.3.3
Example 2.3.4
(−1)n n
(c) Let an = + 1.
n+1
Then (an ) is not a monotone sequence. Since
n (−1)n n (−1)n n
0< < 1, −1 < < 1 and so 0 < an = +1<2
n+1 n+1 n+1
for all n ∈ N∗ . It follows that 0 is a lower bound and 2 an upper bound of (an ).
We show that 0 is the greatest lower bound. Suppose c is a lower bound and 0 ≤ c.
(−1)n n
Then 0 ≤ c ≤ + 1 for all n. In particular, whenever n = 2k − 1, we have
n+1
1
0≤c≤ , and this holds for every k ∈ N+ . it follows from Lemma 2.3.3 (b) that
2k
c = 0, so no lower bound can be larger than 0. It follows thatt inf an = 0; a similar
argument shows that supn an = 2. Note that neither the supremum nor the infimum
is a term of the sequence.
(a) If the sequence (an ) has a largest term (i.e. if there is an p ∈ N+ such that
an ≤ ap for every n ∈ N+ ), then we call ap the maximum of the sequence and
denote it by maxn an .
(b) If the sequence (an ) has a smallest term (i.e. if there is an p ∈ N+ such that
an ≥ ap for every n ∈ N+ ), then we call ap the minimum of the sequence and
denote it by minn an .
• If the range of the sequence (an ) is finite, that is if the set {an : n ∈ N+ } is
finite (this will be the case, for example, if the sequence is eventually constant),
then both maxn an and minn an exist.
A sequence can only have a least upper bound if it has at least one upper bound,
that is if it is bounded above. But does every sequence that is bounded above have
a supremum? The answer is: “It depends where you want it to be.”
To clarify this rather enigmatic answer, we need to return to the example at the
beginning of the section. There we constructed two sequences of rational numbers,
(an ) and (bn ) such that (an ) is increasing, (bn ) decreasing, bn − an = 10−(n−1) and
an < x < bn for every n ∈ N+ , where x2 = 2. It follows from this that x is an upper
bound for (an ) and a lower bound for (bn ). It also follows that for a number y to
be the least upper bound of (an ), it will have to satisfy an ≤ y ≤ bn (since for every
m ∈ N+ , bm is an upper bound of the sequence (an )).
Suppose now that there is a rational number y such that an ≤ y < x ≤ bn for
every n (that is, y is a rational upper bound for (an ), and smaller than x). Then
CHAPTER 2. NUMBERS AND SEQUENCES 36
• a sequence of rational numbers which is bounded above but does not have a
rational supremum;
• a sequence of rational numbers which is bounded below, but does not have a
rational infimum.
At the same time we have shown that if we allow a number which corresponds to a
point on the real line, the sequence (an ) has a supremum and the sequence (bn ) has
an infimum. We will define the real numbers in such a way that it is this property
(the existence of a supremum for every sequence which is bounded above) that will
distinguish the real numbers from the rational numbers.
Summary:
In this section we introduced the least upper bound, or supremum, of a sequence
that is bounded above, and the greatest lower bound, or infimum, of a sequence
that is bounded below. An example was given to show that a rational sequence
which is bounded above need not have a rational supremum. A sequence that is
bounded above need not have a maximum, but if it has, it equals the supremum of
the sequence. Similar comments applies to a minimum.
Exercises
1. For each of the following sequences (xn ), say whether supn xn , maxn xn , inf n xn
and minn xn exist, and where it does exist, find it.
(a) xn = (−1)n ;
(b) xn = 3−n ;
(c) xn = (−2)n ;
5n − 9
(d) xn = ;
3n
(e) xn = (2 + (−1)n )n + n1 .
CHAPTER 2. NUMBERS AND SEQUENCES 37
3. Let (an ) and (bn ) be bounded sequences such that for every n, m ∈ N+ ,
an ≤ bm . Show that supn an ≤ inf n bn .
such that (R, +, ×, <) is an ordered field, which also satisfies the axiom:
CHAPTER 2. NUMBERS AND SEQUENCES 38
You may wonder why the axiom LUB is phrased in terms of a supremum only.
What about infima? Here is the reason.
(a) If (an ) is bounded below, then the sequence (−an ) is bounded above and
inf n an = − supn (−an ).
(b) If (an ) is bounded above, then the sequence (−an ) is bounded below and
supn an = − inf n (−an ).
Proof: (a) Since (an ) is bounded below, there is a b ∈ R such that an ≥ b for all
n ∈ N+ . Hence −an ≤ −b for all n ∈ N+ , showing that (−an ) is bounded above.
By axiom LUB, c = supn (−an ) ∈ R, and so c ≥ −an , and hence an ≥ −c, for all
n ∈ N+ . This shows that −c is a lower bound for (an ). Suppose d is any lower
bound for (an ). Then an ≥ d, and hence also −an ≤ −d, for all n ∈ N+ . This means
that −d is an upper bound for (−an ), and since c is the least upper bound, we must
have c ≤ −d, or −c ≥ d. It follows that −c is the greatest lower bound of (an ), and
so inf n an = −c = − supn (−an ).
(b) The proof is similar.
There are a number of important and uncomfortable questions we can ask about
the definition of the real numbers above.
2. If the answer to the previous question is “yes”, is there only one such set (i.e.,
is R unique)?
Fortunately, the answer to each one of the questions is in the affirmative. The proofs
are lengthy and not easy, and if we were to do them properly it would take up a
large part of this course. Since there are many other interesting things to discover
about real numbers, we’ll only make some comments about the proofs and leave it
at that.
That there is a set satisfying the properties in the definition can be shown by
constructing such a set from the rational numbers, using Q and its properties. There
are several ways to do this, all in some sense using the idea that a real number should
be thought of as a number that can be “approximated” by rational numbers. The
problem is to do it in such a way that a sequence of approximations can be thought
of as a number in its own right, and to define addition, multiplication and the
order relation for such “numbers”. If you are not satisfied with starting with the
rational numbers, you could go all the way back to the natural numbers, define them
axiomatically and then construct first the integers and then the rational numbers.
The proof that, once we have constructed an example of a set with operations
satisfying the properties of the definition, all other examples must essentially be the
same (in much the same way that two isomorphic vector spaces are essentially the
same, as vector spaces) is beyond the scope of this course.
To see that we can think of Q as a subset of R, we think of the additive identity
of R as the natural number 0, and the multiplicative identity as the number 1. Since
R is closed under addition, it must contain 1 + 1, 1 + 1 + 1, . . ., and these we can
identify with the numbers 2, 3, . . .. The additive inverse of each of these numbers
must also be in R, and we can identify them with the numbers −1, −2, −3, . . .. This
shows that Z can be regarded as a subset of R. Since each q ∈ N+ can now be
regarded as an element of R, its multiplicative inverse q −1 must also be in R. We
can now identify the rational number p/q (with p ∈ Z, q ∈ N+ ) with the element
pq −1 of R. This allows us to think of Q as a subset or R. This is not quite the end
of the story, since we also need to check that the usual addition, multiplication and
the order relation in Q is compatible with this identification of Q with a subset of
R. This can be done, but we won’t do it here. If you want a bit of a conceptual
challenge, you can try to work out what it is that has to be proved.
From now on we’ll assume the existence and uniqueness of the real numbers in
the sense explained above, and also that we may regard every real number as a
rational number. In the rest of this section we start exploring the consequences of
the definition.
A bit of a reassurance: we have seen that we can identify the point on the real
line at a distance x from 0, where x2 = 2, with the supremum of a sequence of
rational numbers, and by our definition of R, it must therefore be a real number.
CHAPTER 2. NUMBERS AND SEQUENCES 40
√
From now on we’ll denote it by 2, as usual.
The two properties of Q listed in Lemma 2.3.3 also hold in R, but we need the
axiom LUB in the proof.
Theorem 2.4.5 Let (an ) and (bn ) be two sequences of real numbers such that
(a) (an ) is increasing and (bn ) is decreasing;
(b) an < bn for every n ∈ N+ ;
(c) for every c > 0 in R, there is an n ∈ N+ such that bn − an < c.
Then there is a unique x ∈ R such that an ≤ x ≤ bn for every n ∈ N+ .
Proof: The sequence (an ) is bounded above (by b1 ) and therefore x = supn an exists
and is a real number, and an ≤ x for every n. As we have shown before, every term
of the sequence (bn ) is an upper bound for (an ), and therefore and x ≤ bn for every
n ∈ N+ . This shows that an ≤ x ≤ bn for every n. Suppose that there is a y ∈ R
such that an ≤ y ≤ bn for every n; without loss of generality we may assume that
y ≥ x. Then 0 ≤ y − x ≤ bn − an for every n. For every m ∈ N+ , there is, by the
assumption (c), an n ∈ N+ such that bn − an < m1 . From this we have that for every
m ∈ N+ , 0 ≤ y − x < m1 , and so by Proposition 2.4.4 (b) y = x.
The Nested Interval Property gives us a way of seeing how we can identify points
on the real line with real numbers as we have defined them here. Given a point on
CHAPTER 2. NUMBERS AND SEQUENCES 41
the real line, we first find an interval of length 1 with integer endpoints containing
the point. Next we divide this interval into two intervals with equal length and
choose the one which contains the point. Continuing with this method of bisecting
intervals we can find a nested sequence of intervals all of which contain the point
we started with and such that the length of the n-th interval is 2−(n−1) . It follows
from Proposition 2.4.4(a) that for every c > 0 we can find an n ∈ N+ such that
2−n < c (check this).The intervals therefore satisfy the conditions of the NIP, which
then guarantees that there is a unique real number common to all the intervals, and
by construction this number corresponds to the point we started with.
If, conversely, we start with an element x ∈ R, then we know that for any integer
n we must have x = n, x < n or x > n. We can use this property to find an integer
m such that m ≤ x ≤ m + 1. The number x therefore corresponds to a point in
the interval [m, m + 1] on the real line. Using the same bisection method as above
we can find a nested sequence of intervals, starting with [m, m + 1], all containing
x, and the point on the real line common to all these intervals then represents the
number x.
Summary:
In this section we defined the real numbers R as an ordered field which satisfies the
least upper bound axiom, and showed that this implies that R also satisfies a greatest
lower bound axiom and has the Archimedean and Nested Interval properties. Using
this we could set up a correspondence between the real numbers and points on the
number line.
Exercises
1
1. Prove that for every x ∈ R,with x > 0, there is an n ∈ N such that < x.
n
Deduce that if y is a real number such that 0 ≤ y < x for every positive real
number x, then y = 0.
2. Prove that if the real sequence (an ) is bounded above, the the sequence (−an )
is bounded below and inf(−an ) = − supn an .
3. In this question we give definitions for sets of real numbers that are very similar
to the ones we have already given for sequences.
Let A be a non-empty subset of R.
We’ll see later (Theorem 3.5.12) that it can be shown that every non-empty
set of real numbers which is bounded above has a supremum, and every non-
empty set of real numbers which is bounded below has an infimum. You may
assume this for the moment.
For each of the following subsets of R, say whether it is bounded above, and
if it is, find its supremum. Also say whether it is bounded below, and if so,
find its infimum.
Two of the most fundamental concepts in analysis are those of convergence and
limits. In your calculus course you have certainly worked with limits a great deal;
in fact, the two basic notions in calculus, those of differentiation and integration,
depend for their definitions on the idea of a limit. In addition, the precise definition
of the continuity of a function at a point also makes use of a limit. In this section
we are going to look at limits in a slightly different context: that of sequences.
Looking at the examples and questions raised in the first chapter, you will realise
that satisfactory answers to the questions asked there will have to rely on being able
to say what we mean by the convergence of a sequence, and being able to determine
whether a sequence converges, and if so, to what. We make a start here to finding
the tools to deal with such problems.
43
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 44
would therefore qualify. But one could also argue that a sequence like
1 1 1 1
1, , 1, , 1, , 1, , 1, . . .
2 4 8 16
satisfies this criterium (we can get as close as we like to 0 using the odd-numbered
terms), but we would not feel comfortable with saying that the sequence converges
to 0 (because the even-numbered terms stay far from 0). To avoid this, we could
make our requirement a bit more stringent: ‘(an ) converges to 0 if we can make the
distance between an and 0, and the distance between all subsequent terms and 0, as
small as we like by taking n large enough.”
A definition in words like this is not very useful when we want to prove things
about convergent sequences. We need a very precise definition. Such a precise
definition is given below. At first sight it is not pretty. It contains three quantifiers,
and a Greek letter (). You don’t have to use a Greek letter (any other letter will
do), but the quantifiers have to be there, and they have to be in the correct order.
If all the symbols are a bit overwhelming, you may find the following translation
into words less of a shock to the system: A non-negative sequence (an ) of real
numbers converges to 0 if and only for every positive real distance, we can find a
term in the sequence such that this term and all subsequent terms are within this
distance from 0.
It is time for some examples.
Example 3.1.2 There would be something seriously wrong with our definition if
the sequence (an ) = ( n1 ) did not converge to 0. So let us check that it does indeed
converge to 0 according to the definition. Informally, what we need to do is to find,
for every given a distance ∈ R+ , a term aN in the sequence such aN and all terms
after it is at a distance less than from 0. Since the sequence ( n1 ) is decreasing, it
will be enough to find an N ∈ N+ such that N1 < , or equivalently, N > 1 . The
Archimedean property guarantees that we can find such an N.
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 45
This gives a good idea of how to give a formal proof. Let ∈ R+ . Then 1 ∈ R+ ,
and so it follows from Proposition 2.4.4 that there is an N ∈ N+ such that N > 1 ,
and hence N1 < . Then for every n ∈ N+ , if n ≥ N, then 0 ≤ n1 ≤ N1 < . This
shows that the sequence ( n1 ) converges to 0.
1
Example 3.1.3 You claim that the sequence (n− 2 ) converges to 0.
1
Your opponent, the doubter, comes up with = 10 and challenges you to produce a
corresponding N.
You claim that N = 101 will do.
√ √ √ 1 1
Your opponent checks: If n ≥ 101, then n ≥ 101 > 100 = 10, and so √ < ,
n 10
so she has to admit that you won this round.
1
Your opponent’s next challenge is = 1000 .
You respond with N = 1000001. √ √
Your opponent checks: If N = 1000001, then if n ≥ 1000001, then n ≥ 1000001 >
√ 1 1
1000000 = 1000, and so √ < , so she has to admit that you won the second
n 1000
round as well.
By now your opponent has probably worked out that you have a winning strategy.
1
If she comes up with an , you choose N to be any integer larger than 2 . She
1
realises that you will always win, and concedes that (n− 2 ) converges to 0.
1
we proposed (choosing N > ) is legitimate: there is such an N. The same is true
for the last example.
Can we also use the definition to prove that a sequence does not converge to 0?
Example 3.1.4 At the start of this section we argued that the sequence defined by
does not qualify as a sequence converging to 0. Can we use the definition to show
this does not converge to 0? The odd-numbered terms are all equal to 1, and they
certainly never get to within a distance 21 of 0. If we take = 21 , then a2n−1 = 1 > .
It follows that for every N ∈ N+ , there will be an n ≥ N such that an = 1 > .
Hence for = 12 there will be no N that will satisfy the definition, and so the
sequence does not converge to 0.
It would be tedious in the extreme if we had to come up with a new winning strategy
for every sequence we wanted to prove converges to 0. To avoid this, we are going
to prove some results that will help us to treat large classes of examples at the same
time. You will probably find that none of these results are surprising. So you may
well ask why we bother to prove things that seem obvious. There are a number of
reasons:
• To give credibility to the formal definition by showing that results you think
should be true (using your informal intuition) can indeed be proved (using
only the formal definition);
• To build up a useful stock of results that will make it easier to calculate limits
in future.
Proposition 3.1.5 If an ) and (bn ) are real sequences such that 0 ≤ an ≤ bn for
every n ∈ N+ and bn → 0, then an → 0 as well.
Example 3.1.7 We use the sandwich theorem to show that the sequence (2−n )
converges to 0. Since we have 2n > n for every n ∈ N+ (if you do not believe this,
you can prove it by induction!), we can deduce that 0 ≤ 2−n ≤ n1 for all n ∈ N+ .
But we have seen above that n1 → 0, so Proposition 3.1.5 tells us that 2−n → 0 as
well.
The same proof will in fact show that for every k ∈ N+ with k ≥ 2, the sequence
(k −n ) converges to 0.
We can do even better. Let 0 < r < 12 . We show that the sequence (r n ) converges
to 0. Since 0 < r < 12 , 0 < r n < 2−n for all n ∈ N. Since (2−n ) → 0, we also have
r n → 0.
1
The case 21 < r < 1 needs some more care. We can write r = , for some
1+b
0 < b < 1. Then we have, for all n ∈ N+ ,
n(n − 1) 2 1 11
(1 + b)n = 1 + nb + b + · · · + bn > nb, and so r n = < .
2 (1 + b)n bn
11
Since → 0, also r n → 0.
bn
Proof: (a) We first look informally at what we have to prove. To show that can → 0,
we need to show that for every > 0, we can find an N ∈ N+ such that if n ≥ N,
0 ≤ can < , or 0 ≤ an < c−1 . Since an → 0 and c−1 > 0, we can use c−1 in the
definition of convergence to find a corresponding N. The formal proof consists of
showing that this N works to show that can → 0.
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 48
So let > 0. Since an → 0 and c−1 > 0, we can find an N ∈ N+ such that for
every n ∈ N+ , if n ≥ N, 0 ≤ an < c−1 . But then it follows that for every n ∈ N+ ,
if n ≥ N, 0 ≤ can < . This proves that can → 0.
(b) Informally, given > 0, we want 0 ≤ an + bn < for large enough n. To do
this, it will be enough to ensure that an < 21 , and bn < 12 , for then we’ll have
0 ≤ an + bn < . Will this be possible? Since an → 0, and 21 > 0, we can choose
an N1 such that 0 ≤ an < 21 for n ≥ N1 . In the same way, since bn → 0, we can
choose an N2 such that 0 ≤ an < 21 for n ≥ N2 . If n is larger than or equal to both
N1 and N2 , both inequalities will be satisfied.
For the formal proof, let > 0. Since an → 0 and bn → 0, we can find an
N1 ∈ N+ and an N2 ∈ N+ such that for every n ∈ N+ , if n ≥ N1 , then 0 ≤ an < 21
and if n ≥ N2 , then 0 ≤ bn < 21 . Let N = max{N1 .N2 }. Then for every n ∈ N+ , if
n ≥ N, we have n ≥ N1 and n ≥ N2 . Hence 0 ≤ an + bn < 12 + 21 = . This shows
that an + bn → 0.
(c) We are given that (bn ) is bounded, so there is a c > 0 such that bn ≤ c for all
n ∈ N+ , and so 0 ≤ an bn ≤ an c. The result now follows from (a) and the sandwich
theorem.
We finish this section by returning to the Nested Interval Property (NIP) that
we introduced in the previous chapter. We saw there that the axiom LUB implies
the NIP, and promised to show that the converse is also true. We are now in a
position to do this. To start note that now that we have defined what it means for
a non-negative sequence to converge to 0, we can restate the NIP, as is done in the
theorem below. Check that you can see why this statement of the NIP is equivalent
to the one given in Theorem 2.4.5.
Theorem 3.1.9 Suppose that the NIP holds: If (an ) and (bn ) are two sequences of
real numbers such that
(a) (an ) is increasing and (bn ) is decreasing;
(b) an < bn for every n ∈ N+ ;
(c) bn − an → 0,
then there is a unique x ∈ R such that an ≤ x ≤ bn for every n ∈ N+ .
Then every non-empty subset of R which is bounded above has a supremeum.
Summary:
We started our investigation of convergent sequences in this section by defining what
it means for a non-negative sequence to converge to 0. The definition was used to
find some useful examples of such sequences, and we also saw how it can be used
to show that a sequence does not converge to 0. A few important properties of
sequences converging to 0 were proved. Finally we proved that the Nested Interval
Property is equivalent to the Least Upper Bound axiom.
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 50
Exercises
4. Prove or disprove:
Definition 3.2.1 Let (an ) be a real sequence and a ∈ R. Then we say that the
sequence (an ) converges to a if and only if the non-negative sequence (|an − a|)
converges to 0.
More formally, using quantifiers and connectives, this reads:
2n + 1
Example 3.2.2 Let an = .
n
Looking at the first few terms, it seems fairly obvious that (an ) converges to 2.
According to the definition we need to check that the sequence 2n+1
n
− 2 converges
to 0. Since
2n + 1 2n + 1 − 2n 1
−2 = = ,
n n n
1
and we already know that ( n ) converges to 0, this is indeed the case.
There are many ways of writing the fact that (an ) converges to a. The following all
mean the same:
• (an ) converges to a
• (an ) tends to a
• an → a
• an → a as n → ∞
• lim an = a
n→∞
• lim an = a
n
• lim an = a
It might make more sense to write (an ) → a rather than an → a, but most books
write an → a, so we will stick to this tradition.
We start with some easy deductions from the definition.
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 52
Proof: (a) and (b) follows at once from the definition and Exercises 1 and 2 of
section 3.1. Make sure that you know why.
(c) Informally, we want, given > 0, to have n large enough so that |an+k − a| < .
Since an → a, we can find an N ∈ N+ such that |an −a| < if n ≥ N, or |an+k −a| <
if n + k > N. This gives us enough information to write down a formal proof.
For the formal proof, let > 0. Since an → a, there is an N ∈ N+ such that
for all n ∈ N+ , if n ≥ N, |an − a| < . Hence for all n ∈ N+ , if n ≥ N − k, then
n + k ≥ N and so |an+k − a| < . This shows that an+k → a.
The result in (c) above shows that omitting a finite number of terms from a
convergent sequence will not change the fact that it converges, or its limit.
In the case of a constant or eventually constant sequence, the terminology “an
tends to a” is not really appropriate, since in these cases an is always, or eventually,
equal to a. The terminology is so firmly established that we will stick to it even in
these two cases, even though it is not strictly speaking correct.
A convergent sequence can have only one limit:
0 ≤ |b − a| ≤ |b − an + an − a| ≤ |b − an | + |an − a|.
(a) an + bn → a + b;
(c) an − bn → a − b;
Now let C = max{|a1 |, |a2|, . . . , |aN −1 |, 1 + |a|}. Then it follows that for all n ∈ N+ ,
|an | ≤ C, and so (an ) is bounded.
|an bn − ab| = |an bn − abn + abn − ab| ≤ |an − a||bn | + |a||bn − b|.
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 54
Proposition 3.2.8 Let (bn ) be a sequence with all its terms non-zero, and suppose
bn → b, with b 6= 0.
1
(a) The sequence is bounded.
bn
1 1
(b) → .
bn b
an a
(c) If also an → a, then → .
bn b
Proof: (a) It follows as in the proof of Proposition 3.2.6 that taking = 21 |b|
we can find an N ∈ N+ such that if n ≥ N, then | |bn | − |b| | < 21 |b|, and so
− 12 |b| < |bn | − |b| < 12 |b|. From this we can deduce that for n ≥ N, |bn | ≥ 12 |b|,
1 2
or ≤ . If we let C = max{|b1 |−1 , |b2 |−1 , . . . , |bN −1 |−1 , 2|b|−1 }, then it follows
|bn | |b|
1
that ≤ C for all n ∈ N+ . (b) Use the fact that ( b1n ) is bounded, bn → b and
|bn |
that
1 1 |b − bn |
− =
bn b |b||bn |
1 1
to deduce that → , by Proposition 3.1.8 (c).
bn b
(c) This now follows immediately from (b) and Proposition 3.2.7.
n2 + 1
Example 3.2.9 We show that the sequence (an ), with an = , converges
3n2 − 2
and find its limit.
n2 + 1 1 + 1/n2
For any n ∈ N+ , an = =
3n2 − 2 3 − 2/n2
(divide numerator and denominator by n2 ).
1 1 1 2 2 1
Now → 0, so 2 → 0, so 1 + 2 → 1. Also − 2 → 0, so 3 − 2 → 3. So bn → .
n n n n n 3
Make sure that you can give a reason or a reference for each of these steps.
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 55
+
Proof: Since an ≥ 0 for √
every n ∈ N√ , it follows from
√ Theorem
√ 3.2.10 that a ≥ 0.
√ √ √
Since |an − a| = | an − a| | an + a| ≥ | an − a| a, we have that for every
n ∈ N+ ,
√ √ |an − a|
| an − a| ≤ √ .
a
|an − a| √ √
But √ → 0, so it follows from Proposition 3.1.5 that an → a.
a
+
A similar argument can be used to show that if p ∈ N√ , p ≥ 2 and (an ) is a
√
non-negative sequence such that an → a, then also an → a. We leave this as an
p p
exercise.
We can now state and prove a general version of the sandwich theorem.
Theorem 3.2.13 (Sandwich theorem) Let (an ), (bn ) and (cn ) be real sequences such
that an ≤ bn ≤ cn for every n ∈ N+ , and an → s, cn → s. Then (bn ) is convergent,
and bn → s.
√
Proposition 3.2.14 (a) Let a ∈ R+ . Then lim n
a = 1.
n→∞
√
(b) lim n
n = 1.
n→∞
a = (1 + kn )n
n(n − 1) 2
= 1 + nkn + kn + . . . + knn using the Binomial Theorem
2
≥ 1 + nkn (Why can we omit the remaining terms?)
> nkn .
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 57
a a
It follows that, for any n ∈ N+ , 0 < kn < . But → 0, so by the Sandwich
√ n n
Theorem, kn → 0. So 1 + kn → 1; hence n a → 1.
1 √ q 1
(iii) Suppose a < 1. Write b = . Then a = n 1b = √
n
. But if a < 1, then
a n
b
√ 1
b > 1. So n b → 1 (by (ii)), so √ n
→ 1.
b
√
(b) For any n ∈ N+ √, if n ≥ 2, then n n > 1. (Why?) In this case, we can again
write, for every n, n n = 1 + kn for some kn > 0. Then
n = (1 + kn )n
n(n − 1) 2
= 1 + nkn + kn + . . . + knn
2
n(n − 1) 2
> kn
2
r
2 2
It follows that kn2 < , so kn < .
r n−1 rn − 1
2 2
Now → 0, so since 0 < kn < , we have kn → 0.
n − 1√ n−1
It follows that n n = 1 + kn → 1.
Given a sequence, it is quite often fairly obvious that it is likely to converge,
and not too difficult to guess its limit. It is then a question of confirming your
suspicion by using the definition of convergence, or more preferably one or more of
the limit rules. But sometimes, as in the two cases above, it is not quite so obvious,
and certainly not so easy to prove. It would be very useful to have a property or
properties of a sequence that would ensure its convergence. We have seen that a
convergent sequence is always bounded. But boundedness alone is not enough to
guarantee convergence; we need something more. The additional property that will
do the trick is monotonicity.
Theorem 3.2.15 Every real sequence that is increasing and bounded above con-
verges. Its limit is its supremum.
Proof: Let (an ) be an increasing real sequence that is bounded above, and let
a = supn an . We show that an → a. This means that given ∈ R+ , we must find
an N ∈ N+ so that for any n ∈ N+ , if n ≥ N, then |an − a| < . Since a ≥ an for
every n ∈ N+ , we can write this last inequality as a − an < .
Let ∈ R+ be given. Then a − < a. Since a is the smallest upper bound of (an ),
a − is not an upper bound of (an ). So there exists k ∈ N+ such that a − < ak .
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 58
Corollary 3.2.16 Every real sequence that is decreasing and bounded below con-
verges. Its limit is its infimum.
Proof: If (an ) is bounded below and decreasing, (−an ) is bounded above and
increasing. By the theorem −an → supn (−an ) = − inf n an . Hence an → inf n an .
2
Example 3.2.17 Define (an ) by a1 = 4 and an+1 = 3 − for all n ∈ N+ . In
an
Question 3 of the Exercises for Section 2.2 you were asked to show that
2 ≤ an ≤ 4 for all n ∈ N+ , and
(an ) is decreasing.
So (an ) is a decreasing sequence that is bounded below.
It follows from the corollary above that (an ) converges. Suppose limn→∞ an = a. By
Theorem 3.2.10 an → a and an ≥ 2 for every n ∈ N+ implies that a ≥ 2. It follows
from Propositions 3.2.3, 3.2.5 and 3.2.8 that
2 2
a = lim an+1 = lim 3 − =3− .
an a
Then
0 = a2 − 3a + 2 = (a − 2)(a − 1), and so a = 2 or a = 1.
1 2
Example 3.2.18 Consider the sequence (cn ) defined by c1 = 2, cn+1 = cn +
2 cn
for all n ∈ N+ .
We show by induction that c2n ≥ 2 for all n ∈ N+ .
Clearly c21 ≥ 2.
Suppose that c2k ≥ 2 for some k ∈ N+ . We show that c2k+1 ≥ 2.
2
1 4 1 4 1 2
c2k+1 −2 = c2k +4+ 2 −2 = 2
ck − 4 + 2 = ck − ≥ 0.
4 ck 4 ck 4 ck
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 59
Note that it follows from Theorem 3.2.10 that if cn is positive and c2n ≥ 2 for
every n ∈ N+ and cn → c then c is a positive number with c2 ≥ 2. So, in particular,
we can divide by c in the above argument.
Theorem 3.2.19 If a and b are real numbers and a < b, then there is a rational
number x such that a < x < b.
Proof: Let A = {n ∈ N : n > max{(b − a)−1 , b−1 }}. Then A is non-empty. Choose
any q ∈ A. Now let B = {n ∈ N : n < bq}. Then B is a non-empty finite set. Let
p = max B. Then p ∈ B and p + 1 ∈ / B.
p
We show that if we put x = , then a < x < b.
q
p p+1
It is clear that < b. Since b ≤ , it also follows that
q q
p+1 1 p
a = b − (b − a) < − = .
q q q
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 60
Corollary 3.2.20 Every real number is the limit of a sequence of rational numbers.
Summary:
In this section we defined the notion of convergence of a real sequence to a real
number, by saying that a sequence (an ) converges to a iff the distance between an
and a converges to zero. From this definition we were able to prove a large number of
“limit rules”, which enable us to find limits of sequences without having to resort to
the definition. We also proved that bounded monotone sequences always converge:
an increasing sequence to its supremum, and a decreasing sequence to its infimum.
We finished the section by showing that every real number is the limit of a sequence
of rational numbers.
Exercises
1. For each of the following sequences, show that it converges and find its limit.
You may use any of the results proved in this section.
3
2n + n
(a)
n3 + 4n2
4
n + n2 − 1
(b)
3n5 + 2
n
2 + 3n
(c)
1 + 5n
√ !
5n4 + 3n
(d)
7n − 5n2 + 2
n
8 + n3
(e)
n!
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 61
cos(n)
(f) .
2 + n2
2. Suppose that (xn ) converges, and a, b ∈ R.
(a) Show that if xn ≥ a for all n ∈ N+ , then lim xn ≥ a. [Hint: This should
be very quick; use a theorem in the notes.)
(b) Show that if xn ≤ b for all n ∈ N+ , then lim xn ≤ b.
(c) Conclude that if xn ∈ [a, b] for all n ∈ N+ , then lim xn ∈ [a, b].
3. Let α ∈ R be such that α < 7. Consider the sequence (dn ) defined by:
dn + 7
d1 = α and dn+1 = for all n ∈ N+ .
2
(a) Show that dn ≤ 7 for all n ∈ N+ .
(b) Show that (dn ) is increasing.
(c) Is (dn ) convergent? If so, find its limit.
Show that
The limits limn→∞ zn and limn→∞ yn are known as the lim inf and lim sup
of the sequence (xn ) respectively, and denoted by lim inf n xn and lim supn xn .
n n
6. Let xn = (−1) + . Find supn xn , inf n xn , lim inf n xn and lim supn xn .
n+1
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 62
To get some practice in proving that a sequence diverges, we illustrate the formal
definition of divergence with a sequence we have always thought of as not convergent.
Example 3.3.1 Let (an ) = ((−1)n+1 ) = (1, −1, 1, −1, 1, −1, . . .).
We show that (an ) diverges. We do this by showing that (an ) cannot converge to a
non-negative number a, because every second term of the sequence equals −1 and
is therefore at a distance more that 12 from a. Likewise, (an ) cannot converge to
a negative number a, because every second term of the sequence equals 1 and is
therefore at a distance more that 12 from a.
Here is the formal proof: Let a ∈ R be given.
Choose = 12 .
Let N ∈ N+ be given.
We must find an n ∈ N+ such that n ≥ N and |an − a| ≥ 21 . We do this as follows:
If a ≥ 0, choose any n ≥ N for which an = −1.
Then n ≥ N and |an − a| = | − 1 − a| ≥ 1 > 12 (because | − 1 − a| is the distance
between a and −1, and if a ≥ 0 this distance is greater than or equal to 1).
On the other hand, if a < 0, choose any n ≥ N for which an = 1. Then n ≥ N and
|an − a| = |1 − a| ≥ 1 > 12 , as required.
How easy is it to spot divergent sequences? There is one large class of sequence
that we can immediately classify as divergent:
Proof: This follows immediately from the fact that every convergent sequence is
bounded (Proposition 3.2.6).
The converse of this result is not true, and we have an immediate example: We
have just proved that ((−1)n ) is divergent, but it clearly is a bounded sequence.
Two special kinds of divergent sequences are of interest. We single them out in
the following definition.
Remarks:
(a) Strictly speaking, before we can classify the types of sequences we have just
defined as divergent, we have to prove that if a sequence diverges to ∞ or to −∞,
then it cannot converge to any real number a. This looks so obvious that one can be
forgiven for not even thinking that it is necessary. But it has to be done, and doing
it is in fact a surprisingly useful exercise which helps to get one to think carefully
about the definitions. We ask you to do this in the exercises.
(b) Not all divergent sequences are of the two types we have just defined. The
sequence ((−1)n+1 ) diverges, but does not diverge to ∞ or −∞.
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 65
Proof: This follows almost immediately from the Archimedean property. Write out
the proof for yourself.
√
Example 3.3.6 We show that − n → −∞ as n → ∞.
+
+ +
√ Given K ∈ R we must
Let’s first look informally at what we need to do. √ find
N ∈ N so √ that for any n ∈ N , n ≥ N ⇒ − n < −K. However, − n <
−K ⇐⇒ n > K ⇐⇒ n > K 2 . If we choose N > K 2 , then the required
condition will be satisfied.
Here is the formal proof: Let K ∈ R+ be given. Choose N ∈ N+ such that
N > K 2 (the Archimedean property guarantees that we can do this). Then, for any
n ∈ N+ , √ √
n ≥ N ⇒ n > K 2 ⇒ n > K ⇒ − n < −K,
as required.
The next proposition gives us a useful way of avoiding the definition in many
cases by making use of sequences we know diverge to ∞ or −∞.
Proof: In both cases this follows almost immediately from the definition. Write
out the one-line proofs yourself. newpage
Example 3.3.8 We show that the sequence (an ), with an = (n(2+(−1)n )), diverges
to ∞.
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 66
We finish this section by establishing some fairly obvious links between divergent
and convergent sequences.
1
(a) If an 6= 0 for all n ∈ N+ and an → ∞, then → 0.
an
1
(b) If an > 0 for all n ∈ N+ and an → 0, then → ∞.
an
an
(c) If bn 6= 0 for all n ∈ N+ , bn → ∞ and (an ) is a bounded sequence, then → 0.
bn
Proof: (a) Let > 0, then also K = −1 > 0. Since an → ∞, there is an N ∈ N+
1 1
such that for all n ∈ N+ , if n ≥ N, then an > K = , or 0 < < . This shows
an
1
that → 0.
an
(b) The proof is similar, and left as an exercise.
(c) This follows from (a) and Proposition 3.1.8.
Example 3.3.10 For each of the following sequences, decide whether it converges
or diverges:
1 − 3n
2n + 2
(a) (cn ) = (b) (dn ) = √ .
n3 + n n2 + n − n
1 − 3n 1/n − 3 1
(a) For any n ∈ N+ , cn = 3 = 2 . Now − 3 → −3 and so the sequence
n +n n +1 n
1
− 3 is bounded. Since n2 + 1 ≥ n for every n ∈ N+ , n2 + 1 → ∞, so by
n
Proposition 3.3.9(c), cn → 0.
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 67
2n + 2 2 + 2/n
dn = √ =p ;
2
n +n−n 1 + 1/n − 1
2 p
(divide numerator and denominator by n). Now 2 + → 2 and 1 + 1/n − 1 → 0.
n
Then p
1 1 + 1/n − 1 0
= → = 0.
dn 2 + 2/n 2
p 2 1
Also > 0 for all n ∈ N+ , so
1 + 1/n − 1 > 0 and 2 + > 0 for all n ∈ N+ .
n dn
1
This means we have everything we need to apply Proposition 3.3.9(b) with an = .
dn
1
We conclude that → ∞, i.e. dn → ∞.
an
Summary:
Sequences that do not converge are called divergent. In this section we saw how the
definition of convergence can be used to prove that a sequence diverges. We saw
that the class of divergent sequences contain every unbounded sequence, but many
others as well. Two special kind of divergent sequences were defined: those that
diverge to ∞, and those that diverge to −∞. A link between sequences converging
to 0 and sequences diverging to ∞ was established.
Exercises
6. (a) Suppose that (bn ) is an increasing sequence. Show that if (bn ) is not
bounded above, then bn → ∞.
(b) Give an example to show that the statement in (a) is false if the assumption
that (bn ) is increasing is omitted.
3.4 Subsequences
We have on occasion found it useful when working with a sequence to look at
a “part” of the sequence only. As an example, when calculating the limit of a
convergent recursively defined sequence (an ), we used the fact that the sequence
(an+1 ) obtained by omitting the first term has the same limit as the sequence itself. A
sequence such as ((−1)n +( 21 )n ) can be better understood by looking at the sequence
of even-numbered terms and the sequence of odd-numbered terms separately. It is
natural to call such a sequence obtained by picking some of the terms of the original
sequence (or, equivalently, by leaving out some of the terms of the original sequence)
a subsequence. It is this idea we are going to explore in more detail in this section.
A subsequence of a sequence is obtained by selecting some of the terms of the
original sequence. It is important that we be very clear about how this selection
should be done.
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 69
We can say how we select the terms by listing the numbers (i.e. the subscripts) of
the terms of the original sequence we select. This list will be a list of infinitely many
positive integers, in other words a sequence of positive integers. To ensure that
the terms in the subsequence appear in the same order as in the original sequence,
the sequence of integers has to be increasing. To ensure that we do not choose a
term more than once, this sequence of integers will have to be strictly increasing.
It is traditional to write n1 for the number of the first term we select, n2 for the
number of the second term, and so on. The list will then be the sequence of integers
(nk ). (We could make this even clearer by writing (nk )∞k=1 , to indicate that k is the
running index.). The subsequence then is the sequence
Definition 3.4.1 Let (an ) be a real sequence and let (nk ) = (n1 , n2 , n3 , . . .) be a
strictly increasing sequence of positive natural numbers, i.e. n1 < n2 < n3 < . . ..
Then the sequence (ank ) = (an1 , an2 , an3 , . . .) is called a subsequence of (an ).
For example, given a sequence (an ) = (a1 , a2 , a3 , a4 , . . .) the following are some
subsequences of (an ):
The condition that (nk ) be strictly increasing is important. It means, for exam-
ple, that a sequence like
(a2 , a1 , a4 , a3 , a6 , a5 , a8 , a7 , . . .)
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 70
Proposition 3.4.3 Suppose that (an ) is a convergent sequence. Then every subse-
quence of (an ) also converges, and has the same limit as (an ).
Proof: Let’s first have an informal look at how to go about the proof. Suppose
an → a and (ank ) is a subsequence of (an ). We must show that ank → a as k → ∞.
(Note that since k is the index that changes, it is k that tends to infinity.) So, given
∈ R+ , we need N ∈ N+ so that for all k ∈ N+ , k ≥ N ⇒ |ank − a| < . But since
an → a, we can find M ∈ N+ so that for all k ∈ N+ , k ≥ M ⇒ |ak − a| < . Since
nk > k, it looks as if taking N = M will do the trick.
Here is the formal proof: Suppose that an → a and that (ank ) is a subsequence
of (an ). Let ∈ R+ be given. Since ak → a as k → ∞, there exists N ∈ N+ so
that for all k ∈ N+ , k ≥ N ⇒ |ak − a| < . Then since nk ≥ k for all k ∈ N+ , by
Lemma 3.4.2,
k ≥ N ⇒ nk ≥ k ≥ N ⇒ |ank − a| < .
Corollary 3.4.4 If the sequence (an ) has two subsequences that converge to different
limits, then (an ) diverges.
This corollary often provides an easy way to show that a sequence diverges. For
example, we now have a painless way to show that (an ) = ((−1)n ) diverges. The
subsequence (a2n ) converges to 1 (it is a constant sequence!), and the subsequence
(a2n−1 ) converges to −1. Therefore (an ) diverges.
The next proposition is probably the most surprising result we will show about
subsequences. It indicates that subsequences can often be better behaved than the
original sequence. The proof is quite cunning; it is easy to get confused trying to
prove this, if you do not hit on the right idea.
Proof: We identify certain terms of a sequence for which we use the quite descriptive
name peak point.
Given a sequence (an ) we say that aM is a peak point of the sequence (an ) if and
only if for all n ∈ N+ , n ≥ M ⇒ aM ≥ an .
Informally, if you’re standing on a peak point and you look right (in the direction of
the positive n-axis) all the terms of the sequence are lower down than you, or on the
same level. This is shown in the graph below. Usually the graph of a sequence will
consist of dots only, but here we have joined them with straight lines to illustrate
where the name peak point comes from.
a2
a4
a8
a3
a6 a10
a9
a1 a7 a12
a5 a11
0 1 2 3 4 5 6 7 8 9 10 11 12
For the sequence in this graph, a2 , a4 , a8 , a9 , a10 are peak points; a1 , a3 , a5 , a6 , a7 , a11
are not.
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 72
Proposition 3.4.5 gives some indication why monotone sequences are important.
While it is certainly not true that every sequence is monotone, every sequence does
have a monotone subsequence. The next result is even better.
Proof: Suppose (an ) is a bounded sequence. By Proposition 3.4.5 (an ) has either an
increasing or a decreasing subsequence, (ank ). Now (ank ) is also bounded. (One can
use the upper and lower bounds of (an ).) By Theorem 3.2.15 and Corollary 3.2.16,
(ank ) converges.
Summary:
In this section we made precise the important notion of a subsequence of a sequence.
Subsequences inherit good properties from their parent sequences: a subsequence of
a bounded sequence is itself bounded and a subsequence of a convergent sequence is
also convergent, to the same limit. But a subsequence may have even better proper-
ties than its parent sequence: Every bounded sequence, not necessarily convergent
itself, has at least one convergent subsequence (the Bolzano-Weierstrass theorem).
Historical Notes
Bernard Placidus Johann Nepomuk Bolzano (1781 – 1848)
Bolzano was born in Prague (now in the Czech Republic) and went to the University
of Prague in 1796 to study philosophy and mathematics. From the start he was
interested in the philosophical and foundational aspects of mathematics; as he put
it “those parts of mathematics which was at the same time philosophy”. In 1800 he
started three years of theological study and at the same time began on a doctoral
thesis on geometry. In it he investigated the nature of proof in mathematics. In
1804 he was awarded the doctorate, and ordained as a priest in the Roman Catholic
church. He was also appointed as a professor in philosophy and religion at the
University of Prague. He was a pacifist and had strong views on economic justice
which did not make him popular with the authorities. This lead to him losing his
position in 1819, being placed under house arrest and forbidden to publish.
Bolzano embarked on a series of books
on the foundations of mathematics.
The first volume was published in 1810,
the second was written but never pub-
lished. Instead he began work on an
attempt to free calculus from the con-
cept of an infinitesimal. He wanted to
develop rigorous definitions of the no-
tions of limit, convergence and deriva-
tive that did not depend on geometry.
He used his new approach to give a
proof of the intermediate value theo-
rem, and also gave a definition of a
what is now called a Cauchy sequence
before Cauchy did so himself!
After 1817, Bolzano did not publish
any mathematical work for many years,
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 74
but worked on a theory of science and knowledge. He did return to the foundations
of mathematics, and worked on an attempt to put the whole of mathematics on a
logical foundation. In the process he discovered a number of paradoxes concerned
with infinite sets. He was the first to use the term “set”, and to give an example
of an infinite set whose elements is in one-to-one correspondence with the elements
of a proper subset of itself. He anticipated Cantor’s theory of infinite sets in many
ways, but much of his work was only published after his death in 1848.
Exercises
informal description is not good enough to use when you need to prove things, so
we will have to give a rigorous definition of this property. A sequence with this
property is called a Cauchy sequence, named after the mathematician we may hold
responsible for its introduction.
Definition 3.5.1 A real sequence (an ) is a Cauchy sequence if and only if, for
any ∈ R+ there exists N ∈ N+ so that for any n, m ∈ N+ , if n ≥ m ≥ N then
|an − am | < .
(an ) is Cauchy ⇐⇒
(∀ ∈ R+ )(∃N ∈ N+ )(∀n ∈ N+ )(∀m ∈ N+ )[n ≥ m ≥ N ⇒ |an − am | < ].
Very roughly, the terms of a Cauchy sequence get closer and closer to one another.
There is a danger, however, in relying too heavily on such a rough idea. It is possible,
for example, to give an example of a sequence with the property that the distance
between successive terms tend to 0, but the sequence is not a Cauchy sequence (see
the exercises).
To start, we need to know that convergent sequences are Cauchy sequences. We
are, of course, hoping that the converse will be true as well, since if it is, we’ll be
able to identify convergent sequences without knowing what their limits are.
The proposition above expresses precisely the idea that, if the terms of a sequence
are getting closer and closer to some limit, then they are also getting closer to one
another.
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 78
Corollary 3.5.3 If a real sequence is not a Cauchy sequence, then it is not conver-
gent.
Our main aim in this section will be to prove the converse of Proposition 3.5,
i.e. to show that every Cauchy sequence of real numbers converges. First we look
at some examples. Since every convergent sequence is a Cauchy sequence, there is
no lack of examples. What we are looking for are examples of sequences that are
not obviously convergent, but which we can show are Cauchy sequences.
1
Given ∈ R+ , we could choose N ∈ N+ so large that |b − a| < . (Find
+
2N −2
such an N.) Then for any n, m ∈ N ,
1
n ≥ m ≥ N ⇒ |cn − cm | ≤ |b − a| < .
2N −2
This shows that (cn ) is a Cauchy sequence.
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 79
This shows that (sn ) is not a Cauchy sequence, and it therefore follows from Corol-
lary 3.5.3 that (sn ) is not convergent.
Proof: The proof is very similar to the proof that every convergent sequence is
bounded.
Suppose (an ) is a Cauchy sequence. Taking = 1 in the definition of a Cauchy
sequence, we can find an N ∈ N+ such that for every n, m ∈ N+ , if n ≥ m ≥ N,
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 80
Proposition 3.5.8 Suppose that (an ) is a Cauchy sequence. If (an ) has the con-
vergent subsequence (ank ), then (an ) converges and its limit is the same as that of
the subsequence (ank ).
Proof: Suppose that (an ) is a Cauchy sequence and (ank ) is a subsequence that
converges to a. We show that an → a as well. Let ∈ R+ be given.
Since ank → a as k → ∞, there exists N1 ∈ N+ so that for any k ∈ N+ ,
k ≥ N1 ⇒ |ank − a| < 21 .
Since (an ) is Cauchy, there exists N2 ∈ N+ so that for any n, m ∈ N+ ,
n ≥ m ≥ N2 ⇒ |an − am | < 21 .
Let N = maximum {N1 , N2 }.
So, for any n, k ∈ N+ , if n ≥ N and k ≥ N then:
|ank − a| < 21 , because k ≥ N1
|an − ank | < 12 , because n ≥ N2 and nk ≥ k ≥ N2 . Hence
|an − a| = |an − ank + ank − a|
≤ |an − ank | + |ank − a|
< +
2 2
= .
Proof: Put together Proposition 3.10, Corollary 3.13 and Proposition 3.14.
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 81
Example 3.5.10 We can now return to Example 3.5.4. There we showed that
1
the sequence defined inductively by c1 = a, c2 = b and cn+2 = (cn+1 + cn ) for
2
all n ∈ N+ is a Cauchy sequence. We can now say that it must therefore be
convergent. But what is its limit? Let’s put limn→∞ cn = c. Trying a method we
used before, we can then say limn→∞ cn+2 = limn→∞ cn+1 = limn→∞ cn = c and so
c = limn→∞ cn+2 = 12 (limn→∞ cn+1 + limn→∞ cn ) = 21 (c + c). This is reassuring, but
hardly useful; we can still not say what c is. We try something different: From the
definition of cn+2 , we easily obtain cn+2 + 21 cn+1 = cn+1 + 21 cn . Applying this result
repeatedly, we get
1 1 1 1 1
cn+2 + cn+1 = cn+1 + cn = cn + cn−1 = . . . = c2 + c1 = a + b.
2 2 2 2 2
Taking limits in this equation gives
3 1 1 1 2 1
c = c + c = lim cn+2 cn + cn+1 = a + b, and so c = a+ b .
2 2 n→∞ 2 2 3 2
If you look back at what we have done in this chapter, you will realise the
important role played in the results we proved by the least upper bound axiom
(LUB): Every sequence in R that is bounded a bove has a supremum. Using this
we were able to prove that every bounded monotone sequence converges, and this
in its turn lead to a proof that every Cauchy sequence converges.
It is interesting to ask whether we could have taken another axiom in the place
of the least upper bound axiom in the definition of the real numbers. Here is the
answer:
(a) Every sequence of real numbers that is bounded above has a a least upper bound
in R.
(b) Every increasing sequence of real numbers that is bounded above converges to a
real number.
Proof: Since we have already shown that (a) implies (b) (Theorem 3.2.15) and that
(b) implies (c) (Theorem 3.5.9), we only need to prove that (c) implies (a). We leave
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 82
this as a challenging exercises for the prospective mathematicians amongst you. You
will find some hints in the exercises.
It follows that we could equally well have used (b) or (c) in the palace of (a)
in the definition of the real numbers. What is more, we have already seen that (a)
in the theorem above (the axiom we denoted by LUB) is equivalent to the Nested
Interval property (NIP). This means that we could also have replaced the axiom
LUB in the definition of the real numbers by the NIP.
In the exercises for Section 2.4 we defined the supremum of an arbitrary non-
empty subset of the real numbers. At the time we did not ask whether the supremum
of a non-empty subset of R which is bounded above always has a supremum, but
simply asked you to find the suprema of some specific sets. The axiom LUB guaran-
tees that if we can write the elements of the set as a sequence, then the supremum
will exist. Fortunately, more is true:
(a) Every sequence of real numbers that is bounded above has a least upper bound in
R.
(b) Every non-empty set of real numbers that is bounded above has a least upper
bound in R.
Proof: It is clear that (b) implies (a). We leave the proof that (a) implies (b) as
another challenging exercise. You will find some hints in the exercises.
Summary:
In this section we found a property that ensures that a real sequence converges,
but does not require us to guess the limit first. This is the property of being a
Cauchy sequence. We introduced Cauchy sequences and showed that every conver-
gent sequence is a Cauchy sequence. Since a Cauchy sequence is bounded, it has a
convergent subsequence. Although a sequence with a convergent subsequence need
not converge itself, we showed that if the sequence is also a Cauchy sequence, then
it will converge. From all of this we can deduce that a real sequence converges to a
real number if and only if it is a Cauchy sequence.
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 83
Historical notes
Augustin Louis Cauchy (1789–1857)
Cauchy was born in Paris at the time of the French Revolution. When he was four,
conditions in Paris became so dangerous that the family moved to Arcuiel. There
conditions were so hard that they soon returned to Paris. Cauchy’s father took an
active interest in his son’s education, and sought advice from the mathematicians
Laplace and Lagrange. Cauchy’s formal education started with two years devoted to
classical languages, after which he studied mathematics at the Ecole Polytechnique
and engineering at the Ecole des Ponts et Chaussèes. After graduation he took
on a job at Cherbourg, where he worked on harbour facilities for Napoleon’s fleet.
Although he worked extremely hard, he made time for mathematical research and
published his first paper.
In September 1812 Cauchy became ill and returned to Paris. While recuperating
he continued his research and published a further paper. He did not want to return
to Cherbourg, and managed to obtain a posting to the Ourcq Canal project, where
he worked as a student earlier. Cauchy wanted an academic position, but failed in
a number of applications. He did man-
age to obtain further sick leave, and a
stoppage on the canal project caused
by political events gave him two years
in which to do research. One of the im-
portant achievements of this time was
a memoir on integration which became
the basis of his theory of complex func-
tions.
In 1815 he did manage to secure a po-
sition at the Ecole Polytechnique, and
the next year was awarded a prize by
the French Academy of Sciences for a
paper on waves. This was followed by a
solution to a problem posed by Fermat,
and election to the French Academy of
Sciences.
In 1817 he obtained a post at the Collège de France. His textbook Cours
d’analyse of 1821 attempted to establish a rigorous foundation for the study of
the calculus. This included a rigorous definition of the notion of an integral, and a
rigorous treatment of convergence of infinite series.
CHAPTER 3. CONVERGENT AND DIVERGENT SEQUENCES 84
Cauchy’s relations with his fellow scientists were not particularly good. He was
a staunch Catholic and tended to bring religion into his work. He supported the
Jesuits in their attack on the Academy of Sciences. His manner was often abrupt
and very critical, and the young mathematicians Abel and Galois amongst others
were treated badly by him.
After the revolution of 1830, Cauchy decided to leave Paris and he spent some
time in Switzerland. When he returned to Paris, he was asked to swear an oath
of allegiance to the new government. When he refused to do this, he lost all his
positions. He went to Turin, and was appointed professor of theoretical physics
there in 1832. The next year he went from there to Prague to tutor the grandson of
Charles X, but Cauchy’s quick temper and the prince’s lack of interest meant that
this was not a particularly successful venture. Cauchy returned to Paris in 1838,
but could not teach because he still refused to take the oath of allegiance. This
even meant that he could not take up a new position to which he was appointed.
His religuous and political views also resulted in him not being appointed to a
professorship at the Collège de France, even though he was the best candidate.
He continued to do research in mathematical physics, astronomy and differential
equations.
Political changes in France in 1848 led to Cauchy regaining his old appointments.
When he applied for the chair at the Collège de France in 1850, he narrowly lost
out to Liouville, and this soured relations between them. Another dispute (in which
Cauchy was proved to be wrong) led to a great deal of bitterness in the last years of
Cauchy’s life. He died in 1857. His name lives on in many terms in mathematics, like
Cauchy sequences, the Cauchy-Schwarz inequality, the Cauchy-Riemann equations
and the Cauchy integral formula. He contributed to all the then-known areas of
mathematics, and his contributions show an amazing creativity and insight.
Exercises
√
1. Let the sequence (an ) be defined by: a1 = 0 and an+1 = 1 + an for all
n ∈ N+ .
(e) Deduce that (an ) is a Cauchy sequence. (Use the definition of a Cauchy
sequence.)
(f) Does (an ) converge? If so, to what?
2. Let (an ) be a sequence such that |an+1 − an | < Ck n for all n ∈ N+ , where k,
C ∈ R and 0 < k < 1 . Prove that (an ) is a Cauchy sequence.
Infinite series
In the previous chapter we studied infinite sequences and we were particularly inter-
ested in what happens to the terms of a sequence “in the long run”; more precisely,
we wanted to be able to say whether a sequence converges or not, and if it does
converge, what its limit is. These are questions that only make sense for infinite
sequences; for a finite sequence, there is no “long run”. In this chapter we ask what
we can say about the sum of terms of a sequence. If the sequence is finite, there
is not much to say: we can always add finitely many real numbers, and get a real
number as answer. But we saw in Chapter 1 that there are situations that arise
quite naturally where we may want to add infinitely many real numbers.
Taken at face value, the idea of adding infinitely many real numbers seems ridicu-
lous. This is a process that can never stop; there will always be another number to
add. There is clearly no hope of getting a real number as answer. But on second
thoughts there do seem to be situations where we are in fact claiming that we can
add infinitely many numbers and get a real number as answer. The most familiar
case is that of decimal representations for real numbers. An infinite decimal is an
abbreviation for a sum of infinitely many fractions, and we are usually quite com-
fortable with the claim that such a sum equals a real number. We do not think
twice about writing something like
3 3 3 1
0.333 . . . = + + +··· = .
10 100 1000 3
How do we make sense of all this? The answer is that before we can work
rigorously with infinite sums, we have to attach a meaning to such a sum; it is clear
that we cannot take it at face value. “Attach a meaning to” for a mathematician
means “giving a definition of”. We would like to give a natural and useful definition,
one that makes you want to say: “Now this is a sensible way of going about it, coming
86
CHAPTER 4. INFINITE SERIES 87
to think about it, we couldn’t really have done it any differently.” . If the definition
is such that it can accommodate our idea of decimal representations, and results
in infinite sums having many of the properties of finite sums as well, we could feel
well satisfied. This chapter starts with such a definition, and goes on to explore its
consequences.
It is only fair to issue two warnings right at the beginning. The first is that not
all infinite sums are going to cooperate and yield a real number as answer. You’ll
have to accept that there is probably something seriously amiss in a world where
the sum
1+1+1+1+···
yields a real number as answer. So it will be prudent to admit defeat in some cases.
Not every infinite sum has an “answer”.
The second warning is never to take anything infinite lightly. Weird and wonder-
ful things happen when working with the infinite. Infinite sums are no exception.
If you assume that infinite sums behave exactly like finite ones, you do so at your
own peril. You have been warned!
a1 + a2 + a3 + a4 + · · · .
We can add finitely many of the terms of the sequence. In particular, we can find
the sums
s1 = a1
s2 = a1 + a2
s3 = a1 + a2 + a3
s4 = a1 + a2 + a3 + a4
sn = a1 + a2 + a3 + · · · + an .
We call these sums partial sums of the sequence (an ); specifically, sn is called the n-th
partial sum. The fact that we can form such a partial sum for every n ∈ N+ means
that starting with an infinite sequence (an ), we can form a new infinite sequence
(sn ). This in turn allows us to ask what will happen to sn in the long run. More
CHAPTER 4. INFINITE SERIES 88
precisely, the question is whether the sequence (sn ) converges or not. If it does
converge, the limit would be a very natural candidate for the “sum” of all the terms
of the sequence (an ). We formalise this intuitive idea in a definition:
Definition 4.1.1 An infinite real sequence (an ) is called summable if the sequence
(sn ) of its partial sums converges. If this is the case, we
P∞ call the limit of the sequence
(sn ) the sum of the sequence (an ), and denote it by k=1 ak . In short:
∞
X n
X
a1 + a2 + a3 + · · · = lim sn , or ak = lim ak .
n→∞ n→∞
k=1 k=1
∞
X
P P∞
We sometimes write ak or 1 ak instead of the more precise ak .
k=1
Our first example (really a whole class of examples) is a familiar one. It illustrates
that sometimes we have to admit defeat: not all sequences are summable.
Whether the sequence is summable or not will depend on the value of r. We consider
four cases:
(a) If |r| < 1, we have seen before that r n → 0 . It follows from this that if |r| < 1,
1
then (sn ) is convergent and lim sn = . Hence if |r| < 1, the sequence (r n ) is
n→∞ 1−r
1
summable and has sum .
1−r
(b) If r = 1, sn = 1 + 1 + · · · + 1 = n (we are simply adding n 1’s), and hence (sn )
is not convergent. Therefore the constant sequence with all terms equal to 1 is not
summable.
1 − (−1)n 1
(c) If r = −1, sn = = (1−(−1)n ). The sequence (sn ) does not converge,
1 − (−1) 2
and so the sequence ((−1)n ) is not summable.
CHAPTER 4. INFINITE SERIES 89
(d) If |r| > 1, the sequence (r n ) does not converge, and therefore (sn ) also does not
converge. Therefore in this case as well the sequence (r n ) is not summable.
We can now settle and old question: What does 0.999 . . . equal? This is an
1
infinite sum of the type we discussed above, in disguise, with r = 10 . We have
9 9 9 1 1 9 1
0.999 . . . = + +··· = 1+ + +··· = 1 = 1.
10 100 10 10 100 10 1 − 10
1 1 1
We use the fact that for every k ∈ N+ , = − .
k(k + 1) k k+1
The n-th partial sum of this series is
1 1 1
sn = + + ...+
1·2 2·3 n(n + 1)
1 1 1 1 1
= 1− + − + ...+ −
2 2 3 n n+1
1
= 1−
n+1
CHAPTER 4. INFINITE SERIES 90
∞
X 1
So lim sn = 1 and hence = 1.
k=1
k(k + 1)
∞
X 1 1 1
Example 4.1.4 The series = 1+
+ + . . . arises so often that it has a
k 2 3
k=1
1 1
special name. It is called the harmonic series. In this case sn = 1 + + . . . + .
2 n
If you look at Example 3.5.5, you will see that we showed there that (sn ) is not a
∞
X 1
Cauchy sequence, so it does not converge. So diverges.
k
k=1
To create new convergent series from existing ones, we need to know how we
can combine convergent series to form convergent series. We start with sums and
constant multiples. The proofs rely on similar results for sequences.
P P
Proposition 4.1.5 Suppose ak and bk are convergent series. If c ∈ R, then
P P P P
(a) (ak + bk ) converges and (ak + bk ) = ak + bk ;
P P P
(b) (cak ) converges and (cak ) = c ak .
P
Proof: (a) To begin, we calculate the n-th partial sum of the series (ak + bk ).
n
X
(ak + bk ) = (a1 + b1 ) + (a2 + b2 ) + . . . + (an + bn )
k=1
= (a1 + · · · + an ) + (b1 + . . . + bn )
Xn X n
= ak + bk .
k=1 k=1
Pn
ak ) and ( nk=1 bk ) converge, by assumption, the sequence
P
Since the sequences ( k=1
n
! n
! n
!
X X X
(ak + bk ) = ak + bk
k=1 k=1 k=1
CHAPTER 4. INFINITE SERIES 91
by Proposition 3.2.5(a).
(b) The proof is similar and left as an exercise.
So far infinite series seem to behave in much the same way as finite sums. The
proofs of the two properties above relies on the corresponding properties of finite
sums. But as we warned you in the introduction to this chapter, it is dangerous to
assume that it is always business as usual.
Example 4.1.6 (a) We have seen in Example 4.1.2 that the series
∞
X
(−1)k = −1 + 1 − 1 + 1 − 1 + 1 . . .
k=1
diverges .
(b) Now let’s see what happens if we change this series into a new one by adding
brackets. Consider the series
X
ak = (−1 + 1) + (−1 + 1) + (−1 + 1) + . . . .
What we see in this example is that, although the original series in (a) diverges,
the two new series in (b) and (c) converge. This shows that we have created two
new series in (b) and (c) by introducing brackets, and that even the position of the
brackets changes matters. This is in sharp contrast with the situation for finite
series, where the associative law holds and the way we group terms has no effect on
the sum of the terms.
The next result says that the convergence or divergence of a series is not changed
if we omit or insert finitely many terms at the start of the series. The sum will
change, of course (but not the fact that the series converges).
CHAPTER 4. INFINITE SERIES 92
t1 = am
t2 = am + am+1
..
.
tn = am + am+1 + · · · + am+n−1
sn = a1 + . . . + an
= (a1 + . . . + am−1 ) + (am + . . . + an )
= (a1 + . . . + am−1 ) + tn−m+1
= sm−1 + tn−m+1 .
Since (tn ) is convergent, it follows from Proposition 3.4.3 that the subsequence
(tn−m+1 ) also converges. Since sm−1 is a constant, it follows from the above that
(sn ) also converges.
The proof of the converse is similar, and is left as an exercise.
Summary:
In this section we made precise the idea of P an infinite sequence (an ) being summable
or not, or equivalently of the infinite series ∞k=1 ak converging or diverging. A series
converges (respectively diverges) if and only if its sequence of partial sums converge
(respectively diverges). As illustration we looked at the important examples of the
geometric and harmonic series. We showed that sums and constant multiples of
converging infinite series are again convergent, but that the associative law does not
hold for infinite series.
Exercises
2. Say whether the following series are convergent or divergent. Give a reason
for your answer.
P −k
(a) e
P 1
(b)
5k
P 1
(c) ( k + ( 23 )k ).
P
(d) (a + (k − 1)d), where a and d are real constants.
P
Proposition 4.2.1 If ak converges, then an → 0.
P
Proof: SupposeP ak = s and sn = a1 + . . . + an is the n-th partial sum of the
series. Since ak converges, sn → s, and hence also sn−1 → s. For all n ≥ 2,
an = sn − sn−1 , so
P
Corollary 4.2.2 (n-term test for divergence) If an 6→ 0, then ak diverges.
CHAPTER 4. INFINITE SERIES 94
Example
P 1 4.2.3 We have already seen (in Example 4.1.4) that the harmonic se-
1
ries n
diverges. However, n
→ 0. This example shows that the converse of
Proposition 4.2.1 is false.
Proposition 4.2.1 and its corollary are far more useful than it looks at first sight.
Unfortunately, though, these results are often misunderstood and misused. We
summarise exactly what they say (and do not say) as follows:
P
• an 6→ 0 ⇒ an diverges.
P
• an → 0: Does not tell us anything about the convergence of the series an .
P
• an converges ⇒ an → 0.
P
• an diverges : Does not tell us anything about the convergence or divergence
of the sequence (an ).
Example 4.2.4
X k
(a) Does converge or diverge?
4k + 1
n 1 1 P
Here the n-th term is an = = → . Since an 6→ 0, an diverges.
4n + 1 4 + 1/n 4
X
(b) Does (ln(k + 1) − ln k) converge or diverge?
n+1 1
Here the n-th term is an = ln(n + 1) − ln n = ln( ) = ln(1 + ) → 0 (since
n n
ln 1 = 0). This does not tell us anything about the convergence or divergence of the
series. So we calculate the n-th partial sum of the series instead:
sn = (ln 2 − ln 1) + (ln 3 − ln 2) + . . . + (ln(n + 1) − ln n)
= ln(n + 1) − ln 1
= ln(n + 1)
P
As n → ∞, ln(n + 1) → ∞ so (sn ) diverges. So (ln(k + 1) − ln k) diverges.
As we have seen, Proposition 4.2.1 can be useful if we want to show that a series
diverges. It cannot, however, be used for showing that a series converges.
In the previous chapter we proved that a sequence converges if and only it is
a Cauchy sequence. Since an infinite series converges if and only if its sequence of
partial sums converges, it follows that a series converges if and only if its sequence of
partial sums is a Cauchy sequence. This leads almost immediately to the following
criterion for convergence:
CHAPTER 4. INFINITE SERIES 95
+
Theorem 4.2.5 An infinite series ∞
P
k=1 ak converges if and only if for every ∈ R
+ + +
there is an N ∈ N such that such that if n ∈ N , m ∈ N and n > m ≥ N, then
n
X
ak < .
k=m+1
Pn
Proof: Let sn = k=1 be the n-th partial sum of the series. Then for n > m,
n
X m
X n
X
|sn − sm | = ak − ak = ak .
k=1 k=1 k=m+1
with pk ≥ 0 for all k ∈ N+ , where the first form gives a series starting with a
non-negative term, and the second one starting with a non-positive term. The test
is named after the mathematician who introduced the notation that we still use in
calculus today.
Theorem 4.2.6 (The Leibniz Test) Let (pn ) be a sequence such that
(c) pn → 0.
also converges.
CHAPTER 4. INFINITE SERIES 96
Corollary 4.2.7 Let (pn ) satisfy the conditions of the Leibniz test. Then, for any
m ∈ N+ ,
X∞ m
X
(−1)k+1 pk − (−1)k+1 pk ≤ pm+1 .
1 1
k+1
In words: The absolute value of the error made by approximating ∞
P
1 (−1) pk
P m k+1
by 1 (−1) pk is at most pm+1 , or, more loosely: The value of the error made
by approximating the sum of the series by a partial sum is at most equal to the
absolute value of the first term not used in the partial sum.
k+1
Proof: For brevity, write s = ∞
P
k=1 (−1) pk . As usual, let sn be the n-th partial
sum of the series. Then s = lim sn . For any m ∈ N+ ,
∞
X m
X
k+1
(−1) pk − (−1)k+1 pk = |s − sm |
1 1
= | lim sn − sm |
n→∞
= | lim (sn − sm )|
n→∞
= lim |sn − sm |
n→∞
≤ pm+1
This last inequality comes from the proof of Theorem 4.2.6. Make sure that you
know where each of the equalities above come from.
This is of more interest than it may appear, at first sight. You may recall – if
you don’t, don’t worry – that the Taylor series for ln(1 + x) about x = 0 is given by:
x2 x3 x4
ln(1 + x) = x − + − + ...
2 3 4
and so
1 1 1
ln 2 = 1 − + − + ....
2 3 4
This means that what we are really finding here are estimates for ln 2.
X (−1)k+1 3k
Example 4.2.9 Does converge or diverge?
k2 + 1
3n 3/n
Let pn = 2 . Then pn = → 0. Also pn ≥ 0 for all n ∈ N+ and (pn ) is
n +1 1 + 1/n2
decreasing. (This requires a bit of calculation - check it yourself.) So Leibniz’s test
X (−1)k+1 3k
applies, and converges.
k2 + 1
X (−1)k+1 3k 2
Example 4.2.10 Does converge or diverge?
k2 + 1
3n2 3
Let pn = 2 . Then pn = → 3. The Leibniz test does not apply.
n +1 1 + n−2
However, Corollary 4.2.2, the n-th term test for divergence, does apply. If we write
3n2
an = (−1)n+1 2 for the n-th term of the series, then the sequence (an ) diverges.
n +1
(This follows from the fact that a2n → −3 and a2n+1 → 3.)
X X (−1)k+1 3k 2
So, in particular, an 6→ 0 and so ak = diverges.
k2 + 1
Our main aim in this chapter will be to develop tests that allow us to decide
whether a given series converges or diverges. Sometimes we will be able to find the
precise sum, sometimes not. We’ll find that, generally speaking, it is easier to work
with series for which all the terms are non-negative. If a series has both positive
and negative terms, we can create a new series by replacing the terms of the original
series by their absolute values; the new series will then have only non-negative terms.
Our next step will be to explore the relationship between these two series. We first
introduce some helpful terminology:
P
Definition 4.2.11
P A series ak is called absolutely convergent if and only
if the series |ak | is convergent. (More formally, the sequence (an ) is absolutely
summable if and only if the sequence (|an |) is summable.)
CHAPTER 4. INFINITE SERIES 99
At first sight, there is no reason to think that this notion could be at all helpful.
We begin with a series that has positive and negative terms in it, make all the terms
positive and see whether the resulting series converges. Why should this tell us
anything about the convergence of the original series? The next theorem tells us
why.
Theorem 4.2.12
(b) A series is absolutely convergent if and only if the series formed from its positive
terms and the series formed from its negative terms both converge.
P P
Proof: (a) Suppose we are given
P a series ak such that |ak | converges. We use
Theorem 4.2.5 to show that ak also converges.
Let ∈PR+ be given.
Since |ak | converges, it follows from Theorem 4.2.5 that there is an N ∈ N+ so
that for all n, m ∈ N+ , if n > m ≥ N, then |am+1 | + . . . + |an | < . Hence for
n > m ≥ N,
Note that
a+
P P
(i) k is the series of positive terms of ak
(−a−
P P
(ii) k ) is the series of negative terms of ak
(iii) |ak | = a+ −
k + ak for any k ∈ N
+
(iv) ak = a+ − +
k − ak for any k ∈ N .
CHAPTER 4. INFINITE SERIES 100
P + P
First suppose Pthat aP
k and (−a−k ) both converge. Then, by (iii) and Proposi-
a+ −
P P P
tion 4.1.5(a), |ak | = k + ak . This shows that |a |
k converges, i.e. ak
converges absolutely.
P P
Next suppose that ak is absolutely convergent, i.e. that |ak | converges. If we
add the two equations (iii) and (iv) above, we get
1
|ak | + ak = 2a+ +
k , so ak = (|ak | + ak ).
2
P P
P + |a
Now k | converges
P by hypothesis, so ak also converges
P (by part (a)). So
ak = 21 P |ak | + 21 P
P
ak , i.e. the series of positive terms of ak converges. P
− 1 1
P
Similarly, (−ak ) = 2 ak − 2 |ak | and so the series of negative terms of ak
also converges.
From the theorem above it follows that any convergent series with only positive
terms can be used to obtain infinitely many other convergent series, simply by
putting in minus signs at random.
However, not every convergent series can be obtained in this way. There are series
which are convergent, but not absolutely convergent. They get a special name:
P P (−1)k+1 1 1 1
Example 4.2.14 Consider the series ak = = 1 − + − + . . ..
P k 2 3 4
In Example 4.2.8, we saw that ak converges.
P 1 1
In Example 4.1.4, we saw that |ak | = 1 + + + . . . diverges.
2 3
P (−1)k+1
So is a conditionally convergent series.
k
Corollary 4.2.15 If a series is conditionally convergent then the series formed from
its positive terms and the series formed from its negative terms both diverge.
CHAPTER 4. INFINITE SERIES 101
P
Proof: Suppose ak is a conditionally Pconvergent series. We use the same notation
+
as in thePproof of Theorem
P 4.2.12, so ak is the series obtained from the positive
P
terms of ak , and (−ak ) Pis the series obtained
−
from the negative terms of ak .
a+ (−a−
P P
By Theorem 4.2.12, either k diverges or k ) diverges (since
P + ak is not
absolutely
P convergent).
P Now
P +use aP proof by contradiction. Suppose ak converges.
Then (−ak ) = ak − ak , so (−a−
−
k ) converges – a contradiction. One obtains
a similar contradiction if one assumes that (−a−
P
k ) converges.
Summary:
In this section we started the process of finding tests that will enable us to decide
whether a given series converges or diverges, without having to find a sum for the
series. The terms of a convergent series tend to 0, and therefore a series with terms
that do not converge to 0 must be divergent. Leibniz’s test gives sufficient conditions
for an alternating series to converge, and if it converges, also gives an estimate of
the error made when approximating the sum by a partial sum. A series is absolutely
convergent if the series obtained by taking the absolute value of each of its terms
converges. An absolutely convergent series is convergent, but not conversely: a series
which is convergent but not absolutely convergent is called conditionally convergent.
A series is absolutely convergent ifand only if the series formed from its positive
terms and the series formed from its negative terms both converge. If a series is
conditionally convergent, both these series diverge.
CHAPTER 4. INFINITE SERIES 102
Historical Notes
Gottfried Wilhelm von Leibniz (1646 – 1716)
a committee to decide the matter. The clearly biased finding of the committee was
in Newton’s favour, and the Leibniz spent a large part of the last years of his life
trying to refute the report of the committee.
Another major contribution of Leibniz to mathematics was the development of
the binary system of arithmetic. He also made substantial contributions to dynam-
ics. He left Paris in 1676 to take up a position in Hanover, working for the Duke of
Hanover, and stayed there until his death in 1716.
Leibniz travelled extensively during his lifetime, met many of the leading scien-
tists and philosophers of his time, corresponded with more than 600 of them and
put much effort into setting up and promoting the aims of scientific societies. He
had extremely wide-ranging interests, and believed in working in many different
disciplines, ignoring the traditional boundaries between them.
Exercises
Proof: It follows from the fact that all the terms of the series are non-negative that
the sequence of partial sums is increasing. Since an increasing sequence is convergent
if and only if it is bounded, the sequence of partial sums will be convergent if and
only if it is bounded. The result follows from the fact that an infinite series converges
if and only if its sequence of partial sums converges.
We can now use this result to find a test that allows us to determine whether
a series converges or diverges by comparing it to one that is known to converge or
diverge.
Theorem 4.3.2 (Comparison test) Let (an ) and (bn ) be real sequences.
We summarise the uses and limitations of the comparison test in the table below.
The comparison test can only be used if ak ≥ 0 and bk ≥ 0 for every ninN+ .
P P
Know bk converges Know bk diverges
P
ak ≤ bk Deduce ak converges Cannot say anything
P
ak ≥ bk Cannot say anything Deduce ak diverges
As our first application of the comparison test we show that all infinite decimal
expansions (not only repeating ones) represent real numbers.
CHAPTER 4. INFINITE SERIES 106
Proposition 4.3.4 For each k ∈ N+ , let ak ∈ {0, 1, 2, . . . , 9}. Then the series
∞
X ak
converges to a real number x, and 0 ≤ x ≤ 1.
k=1
10k
ak 1
Proof: For every k ∈ N+ , k
≤ 9 k.
P −k 10 10
Since 10 is a convergent geometric series, it follows from the comparison test
∞ ∞
X ak X 1
that k
converges to a real number x, and that 0 ≤ x ≤ 9 k
= 1.
k=1
10 k=1
10
∞
X 1 1 1
Example 4.3.5 Consider the series 2
= 1 + + + . . ..
k=1
k 4 9
1 1
For all k ≥ 2, 2
≤ .
k k(k − 1)
∞
X 1
We saw in Example 4.1.3 that ) = 1. Hence
k=1
k(k + 1
∞ ∞
X 1 X 1
= = 1.
(k − 1)k k(k + 1
k=2 k=1
∞ ∞
X 1 X 1
It follows from the comparison test that converges and ≤ 1.
k=2
k2 k=2
k 2
∞ ∞
X 1 X 1
Hence, by Proposition 4.1.7, converges, and ≤ 2.
k=1
k2 k=1
k2
X 1 1 1 1
Next we look at p
= 1 + p + p + p + . . . for p ≥ 2. Since k 2 ≤ k p for
k 2 3 4
+ 1 1 +
all k ∈ N , we have 2 ≥ p for all k ∈ N , so we can use the comparison test to
k k
P 1
conclude that converges.
kp
Example 4.3.6 Often the comparison test can be used to analyse very complicated
looking series in which most of the complication is irrelevant (for convergence, at
least). For example, consider the series
X 2 + sin4 (k + 1)
.
2k + k 2
CHAPTER 4. INFINITE SERIES 107
2 + sin4 (k + 1) 3
For all k ∈ N+ , 0 ≤ k 2
≤ k (Why?).
2 +k 2
∞
X 3
Also is a convergent geometric series. Since the terms of the original series
k=1
2k
are all positive, the comparison test applies, and the series converges.
∞
X k+7
Example 4.3.7 Consider the series .
k=1
3k 2 − 1
We first try to get a feeling for the kind of series we are dealing with. For large k,
k+7 k 1 X1
the expression is approximately equal to = . Since diverges,
3k 2 − 1 3k 2 3k k
we expect the given series to diverge as well.
k+7 k 1
Now ≥ = ≥ 0 for all k ∈ N+ .
3k 2 − 1 3k 2 3k
X1 X 1
Since diverges, diverges as well. The comparison test applies,
k 3k
X k+7
and diverges.
3k 2 − 1
Sometimes a little more ingenuity is needed to find an appropriate inequality. If we
∞
X k−1
are given the series , we could argue as above that we would expect it
k=1
3k 2 + 1
X1
to behave like , and therefore to diverge.
k
We have, for all k ≥ 3, that k − 1 ≥ 21 (k + 1) and 3k 2 + 1 ≤ 3(k + 1)2 . Therefore
1
k−1 2
(k + 1) 1 1
2
≤ 2
= for all k ≥ 3.
3k + 1 3(k + 1) 6k+1
∞ ∞
1X 1 1X1
Since the series = diverges, it follows from the comparison test
6 k=3 k + 1 6 k=4 k
∞ ∞
X k−1 X k−1
that 2
diverges as well. But then the series 2+1
will also diverge.
k=3
3k + 1 k=1
3k
The next test is arguably the easiest to apply, and has the further merit that it
only requires that the terms of the series be non-zero. There are series, however, for
which this test is inconclusive.
CHAPTER 4. INFINITE SERIES 108
Theorem 4.3.8 (Limit Ratio Test) Let an 6= 0 for every ninN+ and suppose
an+1
there is an ` ∈ R such that lim = `.
n→∞ an
P
(a) If ` < 1, then an converges absolutely (and hence converges).
P
(b) If ` > 1, then an diverges.
(c) If ` = 1, the test is inconclusive (i.e., it cannot say whether the series converges
or diverges).
Proof: (a) We begin with the case ` < 1. (We know ` ≥ 0. Why?)
`+1
Choose b ∈ R such that ` < b < 1. (For example, b = would do.) Then
2
b − ` > 0, so there exists N ∈ N+ so that for all n ∈ N+ ,
an+1
n≥N ⇒ −` < b−`
an
an+1
⇒ −(b − `) < −`< b−`
an
an+1
⇒ <b (ignore the other inequality)
an
⇒ |an+1 | < b|an |.
So for all n > N,
|an | < b|an−1 | < b2 |an−2 | < . . . < bn−N |aN | = bn (b−N |aN |).
Now bn is a convergent geometric series (since 0 < b < 1), andP
therefore b−N |aN | bn
P P
∞
converges as well.
P∞It now followsP from the comparison test that k=N |ak | converges,
hence so does k=1 |ak |, i.e. ak is absolutely convergent.
(b) Now consider the case ` > 1.
`+1
Choose b ∈ R such that 1 < b < `. (Again b = would do.) Then ` − b > 0 so
2
there exits N ∈ N+ so that for all n ∈ N+ ,
an+1
n≥N ⇒ −` <`−b
an
an+1
⇒ −(` − b) < −` <`−b
an
an+1
⇒ b< (ignore the other inequality)
an
⇒ |an+1 | > b|an |.
CHAPTER 4. INFINITE SERIES 109
Since b > 1, bn−N |aN | > |aN | > 0 for all n > N So for any n ∈ N+ , if n > N then
|an | > b|an−1 | > b2 |an−2 | > . . . > bn−N |aN | > |aN |.
P
Now |aN | > 0, so this shows that an 6→ 0 as n → ∞. So, by Corollary 4.2.2, ak
diverges.
X1 X 1
(c) To see that the case ` = 1 is inconclusive, consider the series and .
k k2
P X1 an+1 1/(n + 1) n
If ak = then = = → 1.
k an 1/n n+1
P X 1 an+1 1/(n + 1)2 n2
If ak = then = = → 1.
k2 an 1/n2 n2 + 1
an+1 X 1 X1
So in both cases we have that lim = 1, but converges, whereas
n→∞ an k2 k
diverges.
is absolutely convergent.
If r = 0, the result is clear.
If r 6= 0, apply the Ratio Test:
r n+1
an+1 (n + 1)! r n+1 n! |r|
= n = · n = → 0.
an r (n + 1)! r n+1
n!
an+1 X rk
Since limn→∞ = 0 < 1, converges absolutely (and so, of course,
an k!
converges).
used k = 1.)
1
In this case an = , so
(ln n)n
1
n
(ln n)n
an+1 (ln(n + 1))n+1 1 ln n
= = = · .
an 1 (ln(n + 1))n+1 ln(n + 1) ln(n + 1)
(ln n)n
an+1 1
0≤ ≤ .
an ln(n + 1)
1 an+1
As n → ∞, → 0 so, by the Sandwich Theorem, → 0.
ln(n + 1) an
∞
X 1
According to the ratio test converges.
k=2
(ln k)k
en+1
2
en+1 n2
an+1 (n + 1)2 1
= = n · =e → e > 1.
an en e (n + 1)2 1 + 1/n
n2
P
By the ratio test, ak diverges.
Remark: A careful look at the proof of the ratio test shows that it is not essential
an+1
for lim to exist. We can in fact use much the same argument to show that
n→∞ an
the following more general version of the ratio test holds.
Theorem 4.3.13 (Ratio test) Let (an ) be a sequence of non-zero real numbers.
|an+1 |
(a) If there is an N ∈ N+ and an 0 < ` < 1 such that ≤ ` for n ≥ N, then
P |an |
an converges absolutely.
|an+1 |
(b) If there is an N ∈ N+ such that
P
≥ 1 for n ≥ N, then an diverges.
|an |
Proof: Exercise.
The proof of the next test is quite similar to that of the ratio test, and we
therefore leave it as an exercise. It also has a limit version, which is easier to apply.
+
p
(a)
P f there is an N ∈ N and an 0 < ` < 1 such that n
|an | ≤ ` for n ≥ N, then
an converges absolutely.
p
(b) If there is an N ∈ N+ and an ` > 1 such that n |an | ≥ ` for n ≥ N, then
P
an
diverges.
p P
(c) In particular, if limn→∞ n |an | = ` exists, then an converges absolutely if
` < 1 and diverges if ` ≥ 1. If ` = 1, the test is inconclusive.
CHAPTER 4. INFINITE SERIES 112
Proof: Exercise.
Example 4.3.15 In Example 4.3.12 we applied the limit ratio test to the series
k
P∞
k=1 kr . We now try to apply the limit root test to the same series. We have
n
an = nr , so p √
lim n |an | = lim n n|r| = |r|.
n→∞ n→∞
It follows from the root test that the series converges for |r| < 1 and diverges for
|r| > 1.
The last test we consider in this section can be used only for non-negative series,
and uses improper integrals.
exists.
Proof: The following graph illustrates the inequalities that we will need:
Let us write
n
X
sn = ak+1
k=1
n Z
X k+1 Z 2 Z n+1 Z n+1
tn = f (x) dx = f (x) dx + . . . + f (x) dx = f (x) dx
k=1 k 1 n 1
n
X
un = ak .
k=1
The sequences (sn ), (tn ), (un ) are all increasing, because f is positive, and from (*)
we have
sn ≤ tn ≤ un for all n ∈ N+ (∗∗)
P
If ak converges, then (un ) is bounded above. But (**) then shows that (tn ) is
bounded above, and since it is an increasing sequence, it converges.
Z n+1
Hence lim f (x) dx = lim tn exists.
n→∞ 1 n→∞
Z n
If lim f (x)dx exists, then (tn ) converges and so is bounded above. But (**)
n→∞ 1 P∞
then shows
P∞ that (s n ) is bounded above, and so converges, i.e. k=2 ak converges.
Then k=1 ak also converges also, by Proposition 4.1.7
1
Let f (x) = p , for p ≥ 0. Then f is positive and decreasing on [1, ∞).
x
Now if p 6= 1
Z n Z n −p+1 n
1 x 1 1
f (x) dx = p
dx = = −1 .
1 1 x −p + 1 1 1 − p np−1
If p = 1,
n n
1
Z Z
f (x) dx = dx = [ln x]n1 = ln n.
1 1 x
We now have to consider three cases:
n
1 −1
Z
• If p > 1, p − 1 > 0, and so p−1 → 0, so lim f (x) dx = , hence
n n→∞ 1 1−p
X 1
converges.
kp
Z n X1
• If p = 1, lim f (x) dx = lim ln n = ∞, so diverges.
n→∞ 1 n→∞ k
(This is just the harmonic series, so this result is not new.)
Z n
1 1−p 1
• If 0 ≤ p ≤ 1, 1 − p ≥ 0, so p−1 = n → ∞, so lim p
dx = ∞, hence
n n→∞ 1 x
X 1
diverges.
kp
1
If p < 0, the function f (x) = p is increasing on [0, ∞), and therefore the integral
x
test cannot be used. We leave it as an exercise for you to show that for p < 0, the
X 1
series diverges.
xp
We summarise our results for the different cases of the p-series:
X 1
• converges for p > 1
kp
X 1
• diverges for p ≤ 1
kp
Then f is decreasing and f (x) ≥ 0 for x ∈ [2, ∞), so the Integral Test applies. In
this case we begin with n = 2, since ln 1 = 0 so f (1) would be undefined.
Z n Z n
1
f (x) dx = dx = [ln(ln x)]n2 = ln(ln n) − ln(ln 2) → ∞.
2 2 x ln x
n ∞
1
Z X
Since lim f (x) dx does not exist, diverges.
n→∞ 2 k ln k
k=2
Summary:
In the last two sections we have looked at a number of tests which enables us to
determine whether a given infinite series converges or diverges. It is often not so
easy to decide which test would be the most appropriate. In the place of the usual
short summary, we give here a strategy for tackling this kind of problem.
P∞
Suppose you are given an infinite series n=1 an and have to decide whether it
converges or diverges.
1. The first, easy test that you can apply to any series
P is the n-th term test
for divergence: If (an ) does not tend to 0, then an diverges.
WARNING: This is only a test for divergence. If we find that an → 0, this
does not help us to decide whether the series converges or diverges. It could
do either.
Pan → 0, you will have to look for another test to decide whether the series
If
an converges or diverges. Here are the possibilities:
2. If the series is an alternating series (i.e. the terms are alternately positive and
negative) we may be able to apply the Leibniz Test: If an = (−1)n+1 pn ,
CHAPTER 4. INFINITE SERIES 116
Exercises
1. Classify the following series as convergent or divergent. Give reasons for your
answers.
X
(a) e−k
X 1 + 2 + ...+ k
(b)
k2
X k
(c)
k3 + 1
X 1
(d)
3k + 1
X ek
(e)
k!
X ln k
(f)
2k 3 − 1
X (k!)2
(g)
(2k)!
X√
(h) ( 1 + k 2 − k).
X 3k
(i)
k5k
∞
X 1
(j) p
for p > 0
k=2
k(ln k)
∞
X 1
(k) for p > 0
k=2
k(ln k)(ln(ln k))p
X 5(−1)k+1
(c)
k 2 − 21
X 1
1
(d) −
k! k
3. Let a ∈ R such that a > 1. Show that the series
converges. [Hint: For the second series, first show that there is a telescoping
series similar to it that converges.]
P1 P 1 P 1
4. The series , and all diverge. Investigate the convergence
k 2k 2k − 1
or divergence of
X 1 1
X 1 1
(a) − (b) − .
k 2k 2k 2k − 1
7. This question is for those of you interested in decimal expansions and the
structure of the real numbers. We have already seen, in Example 4.3.4, that
every infinite decimal expansion represents a real number. We now show that
every real number can be represented by a decimal expansion. We do this for
real numbers between 0 and 1, but the method applies to all real numbers.
(a) Let r be any real number. If 0 ≤ r < 10, show that there is an a ∈ N+
such that
0 ≤ a ≤ r < a + 1 ≤ 10 (1)
(b) Let x ∈ [0, 1). Show that there is an a1 ∈ {0, 1, 2, . . . , 9} such that
a1 a1 1
≤x< + .
10 10 10
[Hint: Consider 10x and apply (1).]
(c) Show that if
k k
X ai X ai 1
i
≤x< i
+ k (2)
i=1
10 i=1
10 10
[Hint: Multiply (2) by 10k and juggle things so that you can apply (1).]
(d) By induction, for all i ∈ N+ there exist ai ∈ {0, 1, 2, . . . , 9} such that, for
all n ∈ N+ ,
n n
X ai X ai 1
i
≤x< i
+ n.
i=1
10 i=1
10 10
n
X ai
(e) Let sn = and show that sn → x. Then you have proved that
i=1
10i
∞
X ai
= x, or more informally, x = 0.a1 a2 a3 a4 a5 . . ..
i=1
10i
diverges, but that different groupings of its terms give series that converge (possibly
to different sums).
On the other hand, we have also seen an example where regrouping P does not
change the sum. It follows from the proof of Proposition 4.1.7 , if ak converges,
then ∞ ∞
X X
ak = (a1 + . . . + am ) + ak
k=1 k=m+1
CHAPTER 4. INFINITE SERIES 121
for any m ∈ N+ . Here grouping the first m terms together does not change the sum.
There is one obvious difference between these two examples – the one series
diverges, the other converges.
P
Theorem 4.4.1 Given a convergent series ak , we can regroup consecutive terms
of the series without changing the sum of the series.
P
Proof: Suppose we are given that P ak = ` for some ` ∈ R. Consider another
series b1 + b2 + b3 + . . . formed from ak byPregrouping the terms so that b1 is the
sum of one or more of the initial terms in ak , b2 is obtained by adding one or
more of the next terms to come along, and so on. We are not allowed to change the
order of the terms, only to regroup them. For convenience, we write:
b1 = a1 + a2 + . . . + ak1
b2 = ak1 +1 + ak1 +2 + . . . + ak2
b3 = ak2 +1 + ak2 +2 + . . . + ak3
..
.
Note that km ≥ m for all m ∈ N+ . Denote the partial sums of the two series by
sn = a1 + . . . + an and tm = b1 + . . . + bm . Then
tm = b1 + b2 + . . . + bm = a1 + a2 + . . . + akm = skm .
P
We are given thatP ak = `, i.e. that sk → ` as k → ∞.
We must show that bn = `, i.e. that tm → ` as m → ∞.
We know (sk ) converges to `. But for each m ∈ N+ , tm = skm , and (s Pkm ) is a
subsequence of (sk ), and therefore must converge to ` as well. But then bn = `
as needed.
We look at rearrangements next. Here the story is considerably more compli-
cated. We begin with an example.
What is happening here? The example above shows that it is possible to take a
convergent series, rearrange the terms and get a series that converges to a different
sum. However, this does not always happen. If you experiment with different
rearrangements of
X 1 1 1 1
= 1 + + + + ...
2k−1 2 4 8
you will see that the sum is always 2. What is the difference between these two
X 1
series? Well, one difference is that contains only positive terms. That is
2k−1
important (as we shall see), but it is not the whole story, as the next example shows.
1 1
In Example 4.4.2, the series of positive terms 1 + + + . . . diverges,
3 5
1 1 1
and the series of negative terms − − − − . . . also diverges.
2 4 6
X (−1)k+1
The difference between the two examples is that is absolutely con-
2k
X (−1)k+1
vergent, whereas is conditionally convergent.
k
For absolutely convergent series all is well:
P
Theorem 4.4.4 Given an absolutely convergent series ak , we can rearrange the
terms of the series without changing the sum of the series.
Proof: We omit the full proof. The basic idea is not difficult. You know that in
a conditionally convergent series, the series formed from its positive terms diverges,
as does the series formed from its negative terms. The terms of the series converges
to 0.
Choose enough positive terms so that your sum is bigger than c. Then add enough
negative ones so that the sum is less than c again. Then add positive ones till the
sum is bigger than c. In this way, the sum gets closer and closer to c.
CHAPTER 4. INFINITE SERIES 124
Summary:
We looked at grouping and rearrangements of infinite series in this section. Grouping
of the terms of a divergent infinite series can change it into a convergent series, and
different groupings can even produce different sums. But groupings of the terms of
a convergent series will not change the sum of the series. Rearrangements of the
terms of an absolutely convergent series will not change the fact that it converges,
or the sum of the series. But the terms of a conditionally convergent series can be
rearranged to converge to any real number, or even to diverge.
Exercises
(b) Why can we choose N2 so that the inequality in line (4) holds?
(c) Why can M be chosen, as claimed in line (5)?
(d) Explain how the inequality in line (6) follows.
(e) Explain how the expression in line (7) is obtained.
(f) How does the theorem follow in line (8)?
Definition 4.5.1 A power series in x about the point a is a series of the form
∞
X
ak (x − a)k = a0 + a1 (x − a) + a2 (x − a)2 + . . . + an (x − a)n + . . .
k=0
In the rest of this section we will take a = 0. Such a power series has the form
∞
X
ak xk = a0 + a1 x + a2 x2 + . . . .
k=0
The results we prove can easily by generalised to the case where a 6= 0, by a simple
substitution.
The familiar Taylor series, for instance
x2 x3 x4
ln(1 + x) = x − + − + ...
2 3 4
CHAPTER 4. INFINITE SERIES 126
x2 x3
ex = 1 + x + + + ...
2! 3!
x3 x5
sin x = x − + − ...
3! 5!
are examples of power series.
Note that it is usual in a power series to begin with k = 0, to allow for the term
0
a0 x = a0 , which is independent of x.
P∞ k
Theorem 4.5.2 Suppose w ∈ R and k=0 ak w converges.
k
P∞
Then k=0 ak x converges absolutely for all x such that |x| < |w|.
k n n
Proof: Since ∞
P
k=0 ak w converges, an w → 0 as n → ∞. So (an w ) is a bounded
sequence and so we can find C ∈ R+ so that, for all n ∈ N+ , the inequality |an w n | ≤
C holds.
Now, for any x ∈ R such that |x| < |w|, and any n ∈ N,
xn x n x n
|an xn | = an w n · n
= |an w n | · ≤C .
w w w
∞
x x X x k
Since does not depend on n and < 1, the series converges (why?),
w w k=0
w
∞
X x k
also converges and, by the Comparison Test, ∞ k
P
so C k=0 |ak x | converges
k=0
w
too. This shows that ∞ k
P
k=0 ak x is absolutely convergent.
ak w k diverges.
P
Corollary 4.5.3 Suppose
ak xk diverges for all x with |x| > |w|.
P
Then
ak xk converged. By
P
Proof: Suppose that, for some x ∈ R, |x| > |w| and
k
P
Theorem 4.5.2, ak w would also converge – a contradiction.
In the proof of the next theorem we will make use of the idea of the supremum
of a set (rather than a sequence) of real numbers. If A is a set of real numbers, an
upper bound of A is a real number C such that x ≤ C for all x ∈ A. If A has an
upper bound, we say it is bounded above, and then it also has a least upper bound,
or supremum, denoted by sup A. It is an upper bound with the property that if
c < sup A, there is an x ∈ A such that x > c (see Question 3 in the exercises for
Section 2.4, and Theorem 3.5.12).
CHAPTER 4. INFINITE SERIES 127
(c) there exists a positive real number R such that the series converges absolutely if
|x| < R and diverges if |x| > R.
(Note this theorem gives no information about the case |x| = R.)
Proof: Suppose (a) and (b) do not hold; we show that (c) must then hold. Define
∞
X
R = sup{|x| : ak xk converges }.
k=0
xn
We apply the Ratio Test, with an = . Then
n
an+1 xn+1 n n
= · n = |x| · → |x|.
an n+1 x n+1
So the series converges absolutely for |x| < 1 and diverges for |x| > 1. So it has
radius of convergence R = 1. To find its interval of convergence, we must still
consider the cases x = 1 and x = −1.
1
If x = 1, the series becomes ∞
P
k=0 , the harmonic series, which diverges.
k+1
(−1)k+1
If x = −1, the series becomes ∞
P
k=0 , which converges (see Example 4.2.8).
k+1
So this power series has interval of convergence [−1, 1).
xn
We can use the ratio test again, with an = :
n!
an+1 |x|n+1 n! |x|
= n = →0
an |x| (n + 1)! n+1
for all x ∈ R. From this it follows that this series converges for all x ∈ R, so its
interval of convergence is R.
x
This is a geometric series. It converges absolutely for < 1, i.e. for |x| < e. It
e
x x
diverges for ≥ 1, i.e. for |x| ≥ e. (Note that the case = 1 is taken care
e e
of automatically by the geometric series.) The interval of convergence is therefore
(−e, e).
Summary:
In this section we considered power series, that is series for which each terms is a
constant multiple of a power of a real variable x. We showed that there is an interval
such that for every x in the interval the series converges, and for every x outside the
interval the series diverges. This interval is called the interval of convergence of the
power series, and its radius is the radius of convergence.
Exercises
1. Find the radius and interval of convergence for each of the following power
series
∞ ∞
X (3x)n X (2x)n
(a) √ (b)
n=0
n+1 n=0
(n + 1)3
∞ ∞
X xn X (nx)n
(c) (d) .
n=0
nn n=0
n!
∞ ∞
X xn X 2n n
(e) (f) x .
n=1
n2 2n n=0
n!
an xn is given by
P
2. Show that the radius of convergence of the power series
|an+1 |
lim whenever this limit exists.
n→∞ |an |
CHAPTER 4. INFINITE SERIES 130
R1 ≥ R3 .
(14)
So finally R1 = R2 = R3 . (15)
(a) How do you know that there is an r with the properties claimed in line
(3)?
(b) Why is the series in line (4) absolutely convergent?
(c) Why is the sequence (|an r n |) bounded, as claimed in line (5)?
(d) Show how the inequality in line (7) is obtained.
(e) Why is the series in line (8) convergent?
(f) How is line (9) obtained from the previous working?
(g) Why does line (10) follow?
(h) Show how the inequality in line (12) is obtained.
(i) Why does line (13) follow?
(j) Explain how the inequality in line (14) is obtained.
4. Show that the three series of Question 3 need not have the same interval of
1
convergence, by considering the case where an = .
n
5. (a) Show that if any one of the series in Question 3 converges for all x ∈ R,
then so do the other two.
(b) Show that if any one of the series in Question 3 has interval of convergence
equal to {0} then so do the other two.
Chapter 5
In this chapter we shift our focus from sequences and series to real-valued functions.
We will look at some familiar notions from the calculus with the same rigour that
we applied to sequences and series. We will not be abandoning sequences altogether,
though. As you will soon see, it is possible to define some key concepts, such as limits
and continuity, in terms of convergence of sequences. This has the huge advantage
that we can use all our knowledge of sequences to prove results about limits and
continuous functions.
Series also will have their day, but that will have to wait till the next chapter,
when we consider series of functions.
132
CHAPTER 5. CONTINUOUS AND DIFFERENTIABLE FUNCTIONS 133
1◦ 1◦
0 •0
◦ −1 ◦ −1
Now g is defined at x = 0, but we have not managed to fill the gap. The problem
is that if we approach x = 0 from the left, we stay on the line y = −1, while if we
CHAPTER 5. CONTINUOUS AND DIFFERENTIABLE FUNCTIONS 134
approach from the right, we stay on the line y = 1. You may recall from your first-
year calculus that we say in this case that the left-hand limit limx→0− g(x) = −1,
and the right-hand limit limx→0+ g(x) = 1. Since the two are different, there is no
hope of filling the gap in the graph. We say in such cases that the limit limx→0 g(x)
does not exist. If this is the case, the function cannot be continuous.
(c) Let’s try to change g in such a way that the limit at x = 0 exists. If we put
h(x) = |g(x)|, we get
1 if x 6= 0
h(x) =
0 if x = 0
Now limx→0− h(x) = 1 = limx→0+ h(x), so limx→0 h(x) does exist. But there is still
a gap in the graph of h at x = 0. The problem is that limx→0 h(x) = 1 6= 0 = h(0).
To fill the gap and ensure continuity at x = 0 we need to have h(0) = limx→0 h(x).
(d) To change h into a continuous function, we put k(x) = 1 for all x ∈ R. Then we
have k(0) = limx→0 k(x), the gap at x = 0 is filled and k is a continuous function.
1◦ 1
0 0
(a) f is defined at a;
There are a number of legitimate concerns that can be raised about this defini-
tion.
• What is limx→a f (x)? When does it exist? In many cases it may be intuitively
clear whether this limit exists and if so what it equals, but this is certainly
not always the case.
• Does it make sense to talk about continuity, and limits, unless the function is
defined not only at a, but also “near” a?
Before we can make much progress, we will have to say clearly what we mean
by the limit of a function at a point, and get clarity on when this limit exists, and
when not.
It is now clear from the examples above and the definition that the functions f, g
and h in Example 5.1.1 are not continuous at 0, but that the function k, defined by
k(x) = 1 for all x ∈ R, is continuous. These are rather trivial examples, but they
do illustrate the essential ideas. There are far more complicated ones to come!
We now turn to differentiability. Geometrically speaking, a real-valued function
f is differentiable at a point a ∈ R if a (unique) tangent can be drawn to the graph
of f at the point (a, f (a)). This means that the graph of f is “smooth” at this
point – it does not change direction abruptly. This idea was used in your first-year
calculus course to motivate the following definition:
Proof: Suppose that f is differentiable at a, i.e. that f 0 (a) exists. We must show
that limx→a f (x) = f (a). Now
lim f (x) = lim f (a + h)
x→a h→0
f (a + h) − f (a)
= lim · h + f (a)
h→0 h
f (a + h) − f (a)
= lim · lim h + lim f (a)
h→0 h h→0 h→0
0
= f (a) · 0 + f (a)
= f (a), as required.
Warning: The converse of this theorem is false. The standard example is the
function f defined by f (x) = |x|, which is continuous at x = 0, but not differentiable
there.
Note that we are using rules like “the limit of a sum is the sum of the limits”
and “the limit of a product is the product of the limits” in the proof above. Before
we can prove results like these, we need a precise definition of the limit of a function
at a point.
We rest our case for the need of a definition. In the next section we do limits
properly!
Summary:
In this section we reminded you of the usual defininitions of continuity and differ-
entiability of a real-valued function at a point, and we saw that in order to make
them rigorous, a precise definition of the limit of a function at a point is needed.
The phrase “near a” is not very precise. What we will require is that there be some
open interval I (i.e., I must be an interval of the form (c, d) = {x ∈ R : c < x < d})
which contains the point a and such that f is defined everywhere on I except perhaps
at a itself. This means that f must be defined on I \ {a} = {x ∈ R : x ∈ I, x 6= a}.
(We sometimes call I \ {a} a deleted interval.)
Our intuitive idea of the limit of a function f at a point a is that it is the number
that the function values f (x) tend to as x tends to a. In the definition that follows,
we use sequences converging to a to approach the point a.
Proposition 5.2.2 Suppose f is defined near a. If there exist sequences (xn ) and
(yn ) such that xn → a and yn → a but f (xn ) → `1 and f (yn ) → `2 with `1 6= `2 ,
then limx→a f (x) does not exist.
Example 5.2.3 (a) Let c ∈ R and f be the constant function defined by f (x) = c
for all x ∈ R. Let a be any real number. We show that limx→a f (x) = c.
So suppose (xn ) is a sequence such that xn → a. Then f (xn ) = c for all n ∈ N+ , so
(f (xn )) is the constant sequence with each term equal to c, which certainly converges
to c.
(b) Let g(x) = x for all x ∈ R. Let a be any real number; we show that
limx→a g(x) = a.
So suppose (xn ) is a sequence such that xn → a. Then g(xn ) = xn so (g(xn )) = (xn ),
so g(xn ) → a as required.
Now for a rather remarkable example: a function for which the limit does not exist
at any point!
We are now in a position to use all the “limit rules” we have for convergent
sequence to prove similar rules for limits of functions.
Theorem 5.2.5 Let a ∈ R, and let f and g be functions such that limx→a f (x) and
limx→a g(x) both exist. Then in each case below the limit on the left exists, and
Proof: Let (xn ) be a sequence such that xn → a. Then f (xn ) → ` and h(xn ) → `.
Also f (xn ) ≤ g(xn ) ≤ h(xn ) for all n ∈ N+ , so by the Sandwich Theorem for
sequences (Theorem 3.2.13) g(xn ) → `.
Summary:
In this section we gave the definition of the limit of a function at a point in terms
of convergence of sequences. This allowed us to prove results for limits of functions
that look very similar to the results for sequences.
Historical Notes
University of Bresslau in 1827. Standards at the university were low, and Dirichlet
took up a position at the Berlin Military College a year later. Soon afterwards he
was appointed at the University of Berlin, where he stayed until 1855. He had a high
teaching load, partly because he was required to keep on teaching at the Military
College even after he was appointed at the University. Improvements in his salary
and an appointment to the Berlin Academy of Sciences in 1831 enabled Dirichlet to
marry Rebecca Mendelssohn, the sister of the composer Felix Mendelssohn.
When Gauss died in 1855, Dirichlet was offered his position at the University of
Göttingen. He initially tried to use the offer to negotiate improved conditions in
Berlin, but when he was unsuccessful, he accepted the chair in Göttingen. Here
conditions were far more favourable, and he had more time for research and excellent
research students. Unfortunately he was not destined to enjoy this for long. While
at a conference in Switzerland in 1858 he suffered a heart attack, and never recovered
fully. He died in Göttingen the next year, shortly after his wife.
Dirichlet made important contributions to many fields of Mathematics. His initial
work was mainly on various aspects of number theory, and is considered to be the
father of analytic number theory. He was also the first to propose the modern
definition of a function. He did outstanding work in mechanics and is also regarded
as the founder of the theory of Fourier series.
Exercises
n
X
1. Prove that if f is a polynomial function (i.e. f is of the form f (x) = ak xk ),
k=0
then for every a ∈ R, limx→a f (x) = f (a).
p(x)
2. Prove that if f (x) = , where p and q are polynomial functions, then
q(x)
limx→a f (x) = f (a) for every a ∈ R such that q(a) 6= 0.
3. Prove or disprove:
(a) If lim f (x) and lim (f (x) + g(x)) exist, then lim g(x) exists.
x→a x→a x→a
(b) If lim f (x) exists and lim g(x) does not exist, then lim (f (x).g(x)) does
x→a x→a x→a
not exist.
(c) If lim f (x2 ) exists, then lim f (x) also exists.
x→0 x→0
CHAPTER 5. CONTINUOUS AND DIFFERENTIABLE FUNCTIONS 142
Proof: Combine Definition 5.1.2 and Definition 5.2.1. Note that for the lim f (x) to
x→a
be defined we need f to be defined on a deleted open interval containing a, and for
f to be continuous at a, we need f to be defines at a as well, hence the requirement
that f be defined on an open interval containing a.
To avoid statements about continuity at a point a becoming very long and te-
dious, we’ll in future omit the fact that the function f is assumed to be defined on
some open interval containing a; whenever we say something like “f is continuous at
a”, we’ll assume that that f is defined on such an interval. When working with two
functions both continuous at a, each one will be defined on such an open interval,
and these intervals may be different. But the intersection of the two open intervals
will then still be an open interval containing a.
In future, when we want to prove that a function f is continuous at a, we’ll start
of by saying “Let (xn ) be a sequence such that xn → a”; the understanding is that
when we say this, there is an open interval I containing a and (xn ) is a sequence in
I, i.e. xn ∈ I for every n ∈ N+ .
Our first two examples are so obvious it hardly seems worth mentioning, but
these give us the building blocks for many other continuous functions.
The next proposition says that if a function is continuous and positive at a point,
then there will be some open interval containing the point where it will still be
positive. This simple result is surprisingly useful, as we’ll soon see.
f
(b) If g(a) 6= 0 then is continuous at a.
g
Proof: (a) Let (xn ) be a sequence such that xn → a. Since f and g are continuous
at a, f (xn ) → f (a) and g(xn ) → g(a). So (by Proposition 3.2.5(a)) f (xn ) + g(xn ) →
f (a) + g(a), so f + g is continuous at a. The other proofs are similar.
(b) Since g(a) 6= 0, there exists δ ∈ R+ so that g(x) 6= 0 for all x ∈ (a − δ, a + δ),
f
by Proposition 5.3.4. Then is defined on the open interval (a − δ, a + δ). Let
g
(xn ) be a sequence contained in the interval (a − δ, a + δ), such that xn → a. Then
f (xn ) f (a) f
f (xn ) → f (a) and g(xn ) → g(a), so → (by Proposition3.2.8). So is
g(xn ) g(a) g
continuous at a.
Corollary 5.3.6 Polynomial functions are continuous at every point of the real line
and rational functions are continuous at every point where the denominator is not
zero.
Proof: Let (xn ) be a sequence such that xn → a. Then f (xn ) → f (a) (why?) and
g(f (xn )) → g(f (a)) (why?). So (g ◦ f )(xn ) → g(f (a)), as required.
(If you are observant, you may notice that we have to be more careful here about
the open intervals on which f and g are defined. See if you can work out what the
problem is, and if you are stuck, look ahead at Theorem 5.3.23.)
So far we have been concentrating on the idea of a function being continuous at
a point. Often, however, we will be more interested in whether or not a function is
continuous on an interval, or on the whole real line. (In what follows, when we say
“interval”, we’ll include the real line as well.) We state precisely what we mean by
this in the next definition.
Definition 5.3.8 Let I be any interval on the real line and f : I → R be a function.
We say that f is continuous on the interval I if and only if for every sequence
(xn ) in I such that xn → x and x ∈ I also, we have f (xn ) → f (x).
If I is an open interval (or the whole real line), this amounts to saying that f is
continuous at every point of I.
This is not so in general, however. Suppose I = [a, b]. If f is continuous on [a, b],
then for every sequence (xn ) such that a ≤ xn ≤ b for every n ∈ N+ and xn → a,
f (xn ) → f (a).
For f to be continuous at the point a, however, we must have f (xn ) → f (a) for
every sequence (xn ) in an open interval containing a, including sequences sequences
(xn ) for which xn ≤ a for some or all n.
Is f continuous on the interval [−4, 0]? Use the sequence (xn ) = (− n1 ) again to
show that f is not continuous on the interval [−4, 0].
We leave it to you to show that f is continuous on the interval (−4, 0).
As for sequences, we can prove that a set S is bounded if and only if there is a C ∈ R+
such that |x| ≤ C for every x ∈ S. We now extend the notion of boundedness to
functions as well.
Continuous functions defined on closed, bounded intervals have some very useful
properties. We look at some of the most important ones here. The first one says
that such a function is bounded (see Figure 5.4 on page 148).
Example 5.3.13 (a) The conclusion of the theorem need not hold if the interval on
1
which f is defined is not closed and bounded. For example f (x) = is continuous
x
on (0, 1) (check!) but is not bounded on this interval.
CHAPTER 5. CONTINUOUS AND DIFFERENTIABLE FUNCTIONS 147
(b) The conclusion of the theorem also need not hold if f is not continuous. For
example,
1
if x 6= 0
x
f (x) =
0 if x = 0
is defined on the closed interval [0, 1] but is not bounded on [0, 1]. It is not continuous
on [0, 1].
The Least Upper Bound axiom says that every sequence which is bounded above
has a least upper bound, or supremum. From this we could also deduce that every
sequence which is bounded below has an infimum, or greatest lower bound. We saw
in Section that it follows from this that something similar is true for sets of real
numbers. We state this result again in the form that we will need it:
As is the case with sequences, if a set S has a minimum (denoted by min S), it is
also the infimum of the set. A non-empty set in R which is bounded below always
has an infimum, but need not have a minimum. Similarly, if a set S has a maximum
(denoted by max S), it is also the supremum of the set. A non-empty set in R which
is bounded above always has a supremum, but need not have a maximum.
Suppose a function f is continuous on an interval [a, b]. Let
S = {f (x) : x ∈ [a, b]}.
From Theorem 5.3.12 we know that S is a bounded subset of R, and so sup S and
inf S exist. In fact, we can prove more: max S and min S exist, as the following
theorem shows.
Figure 5.4: A continuous function is bounded and has a maximum and a minimum.
Proof: Let S = {f (x) : x ∈ [a, b]}, and write M = sup S. We shall find d ∈ [a, b]
such that f (d) = M; this will show that M = max S.
For each n ∈ N+ , M − n1 is not an upper bound of S, so there exists xn ∈ [a, b] such
that M − n1 < f (xn ) ≤ M. Since n1 → 0, f (xn ) → M by the Sandwich Theorem.
On the other hand, (xn ) is contained in [a, b], so is bounded, so has a convergent
subsequence, (xnk ) say. Suppose xnk → d. Then d ∈ [a, b]. By continuity of f ,
f (xnk ) → f (d). However, f (xnk ) → M (since (f (xnk )) is a subsequence of (f (xn ))).
So, finally, f (d) = M (since limits of convergent sequences are unique).
A similar argument shows that if m = inf S, there exists c ∈ [a, b] such that f (c) =
m; i.e. m = min S. Fill in the details yourself.
1 3 3
Example 5.3.16 Consider the function f (x) = x − x2 + 2x on the interval
3 2
[0, 2]. Since it is a polynomial function, f is continuous on [0, 2] (in fact, on R).
Let S = {f (x) : x ∈ [0, 2]}. You can use a bit of first-year calculus to show that
max S = 65 = f (1) and min S = 0 = f (0).
Now consider the same function f as above, but on the interval (0, 2). Then max S =
5
6
= f (1) still holds, but min S does not exist. However, inf S does exist and
inf S = 0.
Now h is not continuous on [0, 2] (why not?). Let S = {h(x) : x ∈ [0, 2]}. Then
min S = 0 = h(0) or h(2), so in this case f attains its minimum at two points in
[0, 2]. However, max S does not exist, but sup S exists and sup S = 2.
The next theorem is one that makes good intuitive sense: If a continuous function
is negative at some point and positive at another, its graph must cross the x-axis
at some point in between the two. You probably used this result frequently when
trying to find approximations to the zeros of a continuous function; now we can
prove it.
Theorem 5.3.18 If the function f is continuous on [a, b], f (a) < 0 and f (b) > 0
then there exists c ∈ (a, b) such that f (c) = 0.
Proof: Before starting the proof we note that there may be more than one such
number c; our strategy will be to find the smallest one. Let
Since b ∈ S, S is not empty; since S ⊆ [a, b], S is bounded. Hence c = inf S exists.
We show that c ∈ (a, b) and f (c) = 0.
Now c ∈ [a, b] (why?); we check that c 6= a and c 6= b. There exists δ ∈ R+
so that f (x) < 0 for all x ∈ [a, a + δ) (this can be proved in the same way as
Proposition 5.3.4). So f (x) ≥ 0 implies that x ≥ a + δ; so c ≥ a + δ. This gives
c 6= a; a similar argument shows c 6= b.
CHAPTER 5. CONTINUOUS AND DIFFERENTIABLE FUNCTIONS 150
Next we show that f (c) = 0. Since c = inf S, for each n ∈ N+ , c + n1 is not a lower
bound of S, so there exists xn ∈ S such that c ≤ xn < c + n1 . So xn → c, and so
f (xn ) → f (c) by continuity of f . Since f (xn ) ≥ 0 for all n ∈ N+ , f (c) ≥ 0.
Now suppose f (c) > 0. Then there exists δ ∈ R+ so that f (x) > 0 for all x ∈
(c − δ, c + δ) ⊆ [a, b]. So (c − δ, c + δ) ⊆ S, which gives inf(c − δ, c + δ) ≥ inf S,
i.e. c − δ ≥ c. This is a contradiction, so f (c) ≤ 0. Putting this all together we get
f (c) = 0.
(In other words, if a continuous function takes on two values, it also takes on any
value in between those two values.)
Proof: Suppose that f (a) ≤ w ≤ f (b) . If w = f (a), we obviously take c = a; if
w = f (b) we can take c = b. So now consider the case f (a) < w < f (b). Define
g(x) = f (x) − w.
Then g is continuous on [a, b] (why?). Also g(a) = f (a) − w < 0 and g(b) =
f (b) − w > 0. So we can apply Theorem 5.3.18 to the function g to obtain c ∈ (a, b)
such that g(c) = 0, i.e. f (c) = w.
The Intermediate Value Theorem is useful in establishing the existence of real
numbers with various properties, as the following two results show.
Theorem 5.3.20 Every non-negative real number has a non-negative square root.
Proposition 5.3.22 Every continuous function f : [0, 1] → [0, 1] has a fixed point.
CHAPTER 5. CONTINUOUS AND DIFFERENTIABLE FUNCTIONS 151
Proof: Assume that f (0) > 0 and f (1) < 1. (Why can we do this?) Define
g(x) = f (x) − x.
Then g(0) = f (0) − 0 = f (0) > 0 and g(1) = f (1) − 1 < 0.
By Theorem 5.25, there exists c ∈ (0, 1) such that g(c) = 0, i.e. such that f (c) = c.
Note that although Proposition 5.3.22 gives the existence of a fixed point of a
function, it does not provide any method for finding it. For example, if f : [0, π2 ] →
[0, 1] is defined by f (x) = cos x, then Proposition 5.3.22 states that there exists
c ∈ [0, π2 ] such that cos c = c. To find an approximation to this solution, we can use
the method described in Chapter 1.
The theory of fixed points is very important, particularly in obtaining the exis-
tence of solutions of differential equations.
Before concluding this section on continuity, we give a characterization of conti-
nuity that you may well encounter as the definition of continuity in other textbooks.
It is the so-called “ − δ definition”.
f : I → R is continuous at a
⇐⇒ (∀ ∈ R+ )(∃δ ∈ R+ )(∀x ∈ I)[|x − a| < δ ⇒ |f (x) − f (a)| < ]. (∗)
Proof: Suppose first that the condition (∗) holds. Let (xn ) be a sequence contained
in I such that xn → a. We must prove that f (xn ) → f (a). Let ∈ R+ be given.
From (∗), there exists δ ∈ R+ so that, for any x ∈ I, |x − a| < δ ⇒ |f (x) − f (a)| < .
Since xn → a there exists N ∈ N+ so that, for any n ∈ N+ , n ≥ N ⇒ |xn − a| < δ.
Putting this together gives, for any n ∈ N+ ,
such an xn for each n ∈ N+ . This gives a sequence (xn ) in I such that xn → a but
f (xn ) 6→ f (a), so f is not continuous at a.
The theorem gives a necessary and sufficient condition for a function f to be
continuous at a point a. Essentially the same proof gives a necessary and sufficient
condition for f to be continuous on an interval :
f : I → R is continuous on I
⇐⇒ (∀ ∈ R+ )(∀a ∈ I)(∃δ ∈ R+ )(∀x ∈ I)[|x − a| < δ ⇒ |f (x) − f (a)| < ].
The condition in the corollary will be satisfied if for a given > 0, for each a ∈ I,
there is a δ > 0 for this a such that if |x − a| < δ, then |f (x) − f (a)| < , i.e. the δ
could be different for different a ∈ I. If, for a given > 0 we can get a δ > 0 that
will do for all a ∈ I, we say that the function is uniformly continuous on I:
f : I → R is uniformly continuous on I
⇐⇒ (∀ ∈ R+ )(∃δ ∈ R+ )(∀x, y ∈ I)[|x − y| < δ ⇒ |f (x) − f (y)| < ]. (∗∗)
Summary:
In this section we used the precise definition of the limit of a function at a point
to give a rigorous definition continuity of a function at a point. Continuity of a
function on an interval was then defined, and two important results about functions
continuous on closed bounded intervals were proved: they are bounded and attain
there maximum and minimum value on the interval at points in the interval. A third
important theorem says that if a continuous function on a closed bounded interval
changes signs between two points, it must be zero somewhere between these two
points. From this the intermediate value theorem can be derived, and this theorem
has several useful consequences. An equivalent definition of continuity, not using
sequences, the − δ definition) was also derived. We also introduced the notion
of uniform continuity, and showed that a function which is continuous on a closed
bounded interval is uniformly continuous there.
Exercises
1. Let h(x) = x − [x] where [x] denotes the largest integer less than or equal to
x.
2. Let I be an interval and suppose the function f : I → R has the property that
there is a constant C ∈ R+ such that |f (x) − f (y)| ≤ C|x − y| for all x, y ∈ I.
Prove that f is continuous on I.
[Functions with this property are called Lipschitz-continuous.]
3. Suppose f is continuous on [a, b].
(a) Show that if f (a) < 0 then there exists δ ∈ R+ so that f (x) < 0 for all
x ∈ [a, a + δ) ⊆ [a, b].
(b) Show that if f (b) > 0 then there exists δ ∈ R+ so that f (x) > 0 for all
x ∈ (b − δ, b] ⊆ [a, b].
4. Prove or disprove:
√
(a) The function f (x) = x is continuous on [0, ∞).
(b) The function g(x) = 17x7 − 19x5 − 1 has a zero in the interval (−1, 0).
5. Show that if the functions f and g are continuous at a ∈ R, then the function
h defined by h(x) = max{f (x), g(x)} is continuous at a.
[Hint: First show that 2 max{f (x), g(x)} = |f (x) − g(x)| + f (x) + g(x).]
6. If f : R → R is continuous and f (r) = 1 for all r ∈ Q, then f (x) = 1 for all
x ∈ R. [Use the fact that between any two real numbers there is a rational
number.]
7. Prove that if n is an odd positive integer and a ∈ R, then there is an x ∈ R
such that xn = a.
[Hint: apply the Intermediate Value Theorem.]
1
8. Show that the function f : (0, 1) → R defined by f (x) = x
is continuous on
(0, 1), but not uniformly continuous there.
You may recall that the differentiability (and hence continuity) of the trigono-
metric and inverse trigonometric functions can be deduced from those of the function
sin x. The differentiability of sin x at any non-zero x can be derived from its differ-
entiability at 0. To show that sin x is differentiable at x = 0 you probably used a
geometric argument, which in its turn relies on a geometric definition of sin x. There
are ways around this. It is possible to define sin x, and many other functions by
means of infinite series. This idea will come up in the next chapter.
Differentiability of a function f at a point a, and the derivative f 0 (a) at a, was
defined in Section 5.1. We now look at differentiability on an open interval.
Theorem 5.4.4 (Rolle’s Theorem) If the function f is continuous on [a, b], dif-
ferentiable on (a, b) and f (a) = f (b), then there exists c ∈ (a, b) such that f 0 (c) = 0.
Then f has a minimum point, c, on (a, b), and by Proposition 5.4.3 again, f 0 (c) = 0.
Example 5.4.5 The following two examples show that both conditions in the the-
orem are essential.
(a) Let
x2
if x ∈ [0, 2)
h(x) =
−x + 4 if x ∈ [2, 4]
Then h(0) = 0 = h(4) but there is no c ∈ (0, 4) such that h0 (c) = 0. The function is
of course not continuous on [0, 4] (it is not continuous at x = 0).
(b) Let g(x) = |x| for x ∈ [−2, 2]. Then g(−2) = 2 = g(2) but there is no c ∈
(−2, 2) such that g 0(c) = 0. In this case, g is not differentiable on (−2, 2) (it is not
differentiable at x = 0).
Proof: Define
f (b) − f (a)
g(x) = f (x) − (x − a).
b−a
Then g is continuous on [a, b] and g(a) = g(b) = f (a). Apply Rolle’s Theorem to g.
f (b) − f (a)
This gives c ∈ (a, b) such that g 0 (c) = 0, so f 0 (c) − = 0.
b−a
f (b) − f (a)
So f 0 (c) = , as required.
b−a
Note that Rolle’s Theorem is a special case of the Mean Value Theorem. We did
Rolle’s Theorem first because the proof of the Mean Value Theorem is easy using it.
f (b) − f (a)
Note also that is the slope of the line joining the points (a, f (a))
b−a
and (b, f (b)). So the Mean Value Theorem states that there is a c ∈ (a, b) at which
the tangent line to f is parallel to this line.
The Mean Value Theorem is a key ingredient in the proofs of many of the stan-
dard results about differentiability. We conclude with a few of these.
CHAPTER 5. CONTINUOUS AND DIFFERENTIABLE FUNCTIONS 157
Corollary 5.4.7 If the function f is differentiable on (a, b) and f 0 (x) = 0 for all
x ∈ (a, b) then f is constant on (a, b).
Proof: Let x1 , x2 ∈ (a, b). Then f (x2 ) − f (x1 ) = (x2 − x1 )f 0 (c) for some c ∈ (a, b).
(Why?) But f 0 (c) = 0 then, so f (x2 ) = f (x1 ).
Corollary 5.4.8 If the function f and g are both differentiable on (a, b) and f 0 (x) =
g 0 (x) for all x ∈ (a, b) then there is a constant c ∈ R such that f (x) = g(x) + c.
Summary:
In this section we looked at some familiar results from calculus involving derivatives,
such as Rolle’s theorem and the mean Value Theorem, and showed that we are now
able to prove them rigorously using the results about limits and continuity proved
in the previous sections.
Historical notes
Michel Rolle (1652 – 1719)
Michel Rolle was born in Ambert, France, the son of a shopkeeper. He received very
little formal education beyond primary school. Before going to Paris in 1675, he
worked as an assistant to several attorneys in his home town. In Paris he worked
as a scribe and an expert in arithmetic, and soon married. He taught himself
mathematics and first came to the attention of the mathematical world when he
solved a problem posed by the mathematician Ozanam. This lead to him being
awarded a state pension and an appointment in the civil service and as a tutor to
CHAPTER 5. CONTINUOUS AND DIFFERENTIABLE FUNCTIONS 158
the sons of the French Secretary of War. He was elected to the Royal Academy of
Sciences in 1685, and subsequently did impressive work in mathematics. He suffered
a stroke in 1708, and this put an end to his mathematical work. He died in 1719
after a second stroke.
Rolle’s main contributions to mathematics were in the field of number theory,
algebra an geometry. He published a book entitled Traité de Algèbre, on the theory
of equations, in 1690, in which he uses ideas that foreshadow the techniques of the
calculus. Despite this, he was a fierce critic of the calculus then being developed,
and this lead to vigorous and sometimes acrimonious debates in the Academy. To
his credit, he later admitted that he had been wrong.
The theorem we now know as Rolle’s theorem appeared in work published by
Rolle in 1691 to
√ justify the methods he used to solve equations. Rolle also introduced
the notation x, and popularised the use of the equality sign (=).
n
Exercises
(a) Show that f has a fixed point in [0, 1] (i.e. that there is an x ∈ [0, 1] such
that f (x) = x).
(b) Let x1 = 1, xn+1 = f (xn ) for n ≥ 1. Use Question 6 to show that there
is a constant k such that 0 < k < 1 and |xn+1 − xn | ≤ k n−1 for all
n ∈ N+ , n ≥ 2.
(c) Use the inequality in (b) to show that (xn ) is a Cauchy sequence.
(d) Show that (xn ) converges to a fixed point of f in [0, 1].
8. Prove the following generalisation of the Mean Value Theorem, known as the
Cauchy Mean Value Theorem:
If the functions f and g are continuous on the interval [a, b] and differentiable
on (a, b), then there is a c ∈ (a, b) such that
In this chapter we combine the notions of sequences, series and functions and look
at sequences and series of functions. This is not such an unfamiliar idea as it may
appear at first. We have in fact already studied one special example of a series of
functions: power series. The terms of a power series are functions (rather simple
ones, namely functions of the form fn (x) = an xn , or power functions as they are
sometimes called). Taylor series are in turn special kinds of power series, and you
have probably realised by now that they are useful. Another very familiar example
of a series of functions is a Fourier series; in this case the functions are sine or cosine
functions of the form an cos nx or bn sin nx.
In the first section of the chapter we will be looking at sequences of functions
and define, in a very natural way, a limit function for such a sequence. We will be
primarily interested in whether the limit function inherits properties like continuity,
differentiability and integrability from the functions in the sequence. In the second
section we ask similar questions about series of functions, and specialise in the end
to power series.
160
CHAPTER 6. SEQUENCES AND SERIES OF FUNCTIONS 161
This function f is called the pointwise limit of the sequence (fn ) of functions. We
write
fn → f pointwise on B.
x
Example 6.1.2 For each n ∈ N+ , let fn (x) = .
n
Fix an x ∈ R and look at the sequence
x x x x
(f1 (x), f2 (x), f3 (x), . . .) = x, , , , . . . , , . . . .
2 3 4 n
This sequence is convergent, and tends to 0. This is the case for every x ∈ R+ . This
means that
B = {x ∈ R : lim fn (x) exists } = R and f (x) = lim fn (x) = 0 for all x ∈ R.
n→∞ n→∞
So fn → f pointwise on R.
Example 6.1.5
(a) In Example 6.1.3 each gn is continuous on the interval [0, 1], but g, their pointwise
limit, is not.
x2n
(b) For each n ∈ N+ , define fn (x) = for each x ∈ R.
1 + x2n
Then each fn is differentiable on the whole real line. Consider various values of x:
If |x| < 1, x2n → 0 so fn (x) → 0.
1
If |x| = 1, x2n → 1 so fn (x) → .
2
If |x| > 1, x2n → ∞ and fn (x) → 1.
Define
0 if |x| < 1
1
f (x) = if |x| = 1
2
1 if |x| > 1
CHAPTER 6. SEQUENCES AND SERIES OF FUNCTIONS 163
4 g4
3 g3
2
g2
1 g1
1 1 1
0 4 3 2 1
Figure 6.1: Graphs of g1 , g2 , g3 and g4 .
On the interval [0, n1 ] the graph of gn forms an isosceles triangle of height n, with
area 21 .
If x = 0, gn (x) = 0 for all n ∈ N+ so gn (x) → 0. If x 6= 0, there exists N ∈ N+ so
that n1 < x for all n ≥ N. So gn (x) = 0 for all n ≥ N, so gn (x) → 0. What this
shows is that gn → g pointwise, where g is the constant zero function.
R1 R1
Now 0 gn (x) dx = 21 for all n ∈ N+ , but 0 g(x) dx = 0, so
R1 R1
0
g(x) dx 6
= lim 0 n
g (x) dx.
n→∞
Our idea of pointwise convergence, while perhaps easy to work with, gives
counter-intuitive (and not very useful) results. What is wrong? Even though fn → f
pointwise there is some sense in which fn is not necessarily “close” to f , even for
large n. (Look at Example 6.1.5(c) again.) We will remedy this state of affairs by
introducing a stronger form of convergence than pointwise convergence – it is called
uniform convergence.
CHAPTER 6. SEQUENCES AND SERIES OF FUNCTIONS 164
nx2
Example 6.1.8 For each n ∈ N+ , let gn (x) = and g(x) = x on the interval
1 + nx
[0, 1]. Then gn → g uniformly on [0, 1], since
nx2
dn = sup −x
x∈[0,1] 1 + nx
x
= sup
x∈[0,1] 1 + nx
1
= sup
x∈[0,1] n + 1/x
1
= ,
1+n
1
and so dn = → 0 as required.
1+n
CHAPTER 6. SEQUENCES AND SERIES OF FUNCTIONS 165
n ≥ N ⇒ |dn | <
⇒ sup |fn (x) − f (x)| <
x∈I
⇒ |fn (x) − f (x)| < for every x ∈ I.
• fn → f pointwise on I ⇐⇒
(∀x ∈ I)(∀ ∈ R+ )(∃N ∈ N+ )(∀n ∈ N+ )[n ≥ N ⇒ |fn (x) − f (x)| < ].
• fn → f uniformly on I ⇐⇒
(∀ ∈ R+ )(∃N ∈ N+ )(∀n ∈ N+ )(∀x ∈ I)[n ≥ N ⇒ |fn (x) − f (x)| < ].
x
Example 6.1.10 We look at the sequence (fn ) on R defined by fn (x) = again.
n
Recall that the pointwise limit function f in this case is the constant zero function.
We have already seen that fn → f pointwise but not uniformly.
When checking pointwise convergence, you would begin with:
Let x ∈ I and ∈ R+ be given.
x |x|
We need N ∈ N+ so that n ≥ N ⇒ | | < . That’s easy: take N > .
n
CHAPTER 6. SEQUENCES AND SERIES OF FUNCTIONS 166
Look at Examples 6.1.5(b) and (c) again – in both cases the convergence cannot
be uniform.
Rb Rb
So a
fn (x) dx → a f (x) dx as n → ∞.
Rb
We define a f (x) dx formally in the next chapter show there that if f is contin-
Rb
uous on [a, b] then a f (x) dx exists. The properties of the integral required in the
above proof are simple, and should be familiar from your first-year course. In the
proof of the next result we will also need the Fundamental Theorem of Calculus.
Proof: Exercise.
In comparison with our nice results about continuity and integration (Theo-
rem 6.1.12 and Theorem 6.1.14), the result about differentiability (Theorem 6.1.15)
is a bit disappointing. It is possible for (fn ) to converge uniformly to f on I , for
differentiable functions fn , but for the conclusion f 0 (x) = lim fn0 (x) to be false.
n→∞
sin nx
Example 6.1.17 Consider fn (x) = on [0, 2π]. If f is the constant zero
n
function, fn → f uniformly on [0, 2π]. To see this, note that
sin nx 1
dn = sup |fn (x) − f (x)| = sup ≤ ,
x∈[0,2π] x∈[0,2π] n n
so dn → 0.
Now fn0 (x) = cos nx for all x ∈ [0, 2π], and f 0 is the constant zero function. However
fn0 does not even converge pointwise to f 0 ; for example fn0 (π) = (−1)n and this does
not converge.
Summary:
In this section we introduced the pointwise limit of a sequence of functions, and
showed that in general the limit of a sequence of continuous functions need not be
continuous, and that differentiation and integration are not interchangeable with the
limiting process. We then introduced the stronger notion of uniform convergence
of a sequence of functions and showed that for a uniformly convergent sequence of
functions we obtain far more satisfactory results.
CHAPTER 6. SEQUENCES AND SERIES OF FUNCTIONS 169
Exercises
1. Find the pointwise limit f of the sequence (fn ), where this limit exists, if
fn : R → R and fn (x) is
1 n2 x xn
(a) e−nx (b) xe−nx (c) (d) (e) .
1 + nx 1 + n3 x2 1 + xn
2. Find the pointwise limit f of the sequence (fn ), where this limit exists, if
fn : R+ → R and fn (x) is
x nx2
(a) (b)
1 + nx 1 + nx
3. For each of the sequences (fn ) in Questions 1 and 2 that have a pointwise limit
f , determine whether (fn ) converges uniformly to f or not.
6. Answer the questions (a) - (c) of the previous question for the sequence of
functions gn : R → R defined by
x2
gn (x) = .
n + x2
CHAPTER 6. SEQUENCES AND SERIES OF FUNCTIONS 170
Definition 6.2.1 Let I be an interval and for each n ∈ N+ , let fn , and f be real-
valued functions defined on I.
For each x ∈ I and each n ∈ N+ , let
n
X
sn (x) = f1 (x) + f2 (x) + . . . + fn (x) = fk (x).
k=1
∞
X
We say that the series of functions fk converges pointwise to f if and only if
k=1
the sequence of functions (sn ) converges pointwise to f .
X∞
Similarly fk converges uniformly to f if and only if the sequence (sn ) con-
k=1
verges uniformly to f .
X (−1)k+1 x
Example 6.2.2 Consider the series on the interval [0, 1].
1 + kx
x
First let’s check that this series converges pointwise. Fix x ∈ [0, 1]. Now →0
1 + nx
x
as n → ∞ and (for fixed x, remember) is a decreasing sequence. So the
1 + nx
X (−1)k+1 x
Leibniz test applies and converges to some real number, which we
1 + kx
denote by f (x).
Next let’s check that this series converges uniformly:
By the error estimate associated with the Leibniz test we have, for any n ∈ N+ ,
n
X (−1)k+1 x x
− f (x) ≤ .
k=1
1 + kx 1 + (n + 1)x
CHAPTER 6. SEQUENCES AND SERIES OF FUNCTIONS 171
Then
n
X (−1)k+1 x
dn = sup − f (x)
1 + kx
x∈[0,1] k=1
x
≤ sup
x∈[0,1] 1 + (n + 1)x
1
= .
1 + (n + 1)
x
The last equality follows because is an increasing function of x (for
1 + (n + 1)x
1
fixed n, now). Since → 0, dn → 0 also.
n+2
∞
X
Theorem 6.2.3 Suppose fk converges uniformly to f on [a, b].
k=1
∞
X
(c) If each fn is differentiable and fk0 converges uniformly on [a, b] to a continuous
k=1
∞
X
function, then f is differentiable on (a, b) and f (x) = 0
fn0 (x) for all x ∈ (a, b).
n=1
The next theorem provides a useful criterion for showing that a series of functions
converges uniformly.
CHAPTER 6. SEQUENCES AND SERIES OF FUNCTIONS 172
∞
X xn
Example 6.2.5 We show that 2 + x2
converges uniformly on [0, 1].
n=1
n
xn
+ 1 1
For any x ∈ I and n ∈ N , ≤ . So we can use (M n ) = .
n2 + x2 n2 n2
∞
X 1 X xn
Since converges, the Weierstrass M-test shows that converges
n2 n=1
n2 + x2
For the rest of this section, we turn our attention to power series again. The
next few theorems make it clear why power series are so very useful in calculus.
∞
X
Theorem 6.2.6 Let ak xk be a power series.
k=0
∞
X
For each x in its interval of convergence I put f (x) = ak xk .
X k=0
If [−r, r] ⊂ I then ak xk converges absolutely and uniformly to f on [−r, r].
∞
X
Theorem 6.2.7 Let an xn be a power series with radius of convergence R. For
n=0
∞
X
each x in its interval of convergence, put f (x) = an xn . Then f is differentiable
n=0
∞
X
on (−R, R), and for all x ∈ (−R, R), f 0 (x) = nan xn−1 .
n=1
Proof: Let x ∈ (−R, PR) and choose real numbers r and s such that |x| < r < s < R.
n
By Theorem 6.2.6, an x converges uniformly and absolutely on [−s, s]. We show
n−1
P
below that nan x converges uniformly on [−r,Pr]. This requires a bit of calcula-
tion. But once we have done that, we know that nan xn−1 converges P
to a contin-
uous function, by Theorem 6.1.12, and P so, by Theorem 6.1.15, f (x) =P nan xn−1 .
0
Now, we show uniform convergence of nan xn−1 . Note that, since an sn is ab-
n
solutely convergent, the sequence (|an |s ) converges to 0, and so is bounded. Let B
be an upper bound for it. Then, for any x ∈ [−r, r],
≤ n|an |r n−1
n|an |r n sn
=
rsn
n r n
= |an |sn
r s
B r n
≤ n .
r s
X r
Use the Ratio Test to check that n( )n converges. You will need the fact that
P s n−1
r < s. By The Weierstrass M-test, nan x converges uniformly on [−r, r], as
required.
We follow up this result with a similar one for integrals. As before the working
for integrals is somewhat easier than that for derivatives.
Proof:
Z d Z ∞
dX
f (x)dx = an xn dx
c c n=0
Z d n
X
= lim ( ak xk )dx
c n→∞
k=0
Z d n
X
= lim ( ak xk )dx
n→∞ c k=0
n
XZ d
= lim ak xk dx
n→∞ c
k=0
∞
X an
= (dn+1 − cn+1 ).
n=0
n+1
∞ ∞ ∞
X
n
X X an n+1n−1
Proposition 6.2.9 The three series an x , nan x and x
n=0 n=1 n=0
n+1
have the same radius of convergence. However, they need not have the same interval
of convergence.
CHAPTER 6. SEQUENCES AND SERIES OF FUNCTIONS 175
∞
X x2k+1
Example 6.2.10 Consider f (x) = (−1)k .
k=0
(2k + 1)!
To find the radius and interval of convergence, apply the Ratio Test. For any x ∈ R,
Summary:
In this section we introduced the notion of pointwise and uniform convergence of
series of functions. The results of the previous section can be applied to the sequence
of partial sums of such series. The Weierstrass M-test is a very useful test for uniform
convergence of series of functions. When we apply the general results to power
series, we find that such series are very well behaved, and can be differentiated and
integrated term by term on any interval inside its interval of convergence.
CHAPTER 6. SEQUENCES AND SERIES OF FUNCTIONS 176
Exercises
∞
X x2k+1 x3 x5 x2n+1
f (x) = (−1)k = x− + − . . . + (−1)n + ...
k=0
2k + 1 3 5 2n + 1
R1 ≥ R3 .
(14)
So finally R1 = R2 = R3 . (15)
(a) How do you know that there is an r with the properties claimed in line
(3)?
(b) Why is the series in line (4) absolutely convergent?
(c) Why is the sequence (|an r n |) bounded, as claimed in line (5)?
(d) Show how the inequality in line (7) is obtained.
(e) Why is the series in line (8) convergent?
(f) How is line (9) obtained from the previous working?
(g) Why does line (10) follow?
(h) Show how the inequality in line (12) is obtained.
(i) Why does line (13) follow?
CHAPTER 6. SEQUENCES AND SERIES OF FUNCTIONS 178
Integration
One of our aims in this course has been to put first-year calculus on firmer founda-
tions. We have done this by giving proper definitions for the basic concepts encoun-
tered in the calculus: convergence, limits, continuity, differentiability. There is still
one key concept in the calculus that we have not dealt with, namely integration. In
your first-year course you have seen how the definite integral can be defined as a
limit of sums, and you were told that the definite integral of a continuous function
over a bounded interval exists. It seems reasonable to expect that now that we have
a better understanding of limits and continuity we should be able to treat integra-
tion a little more rigorously, and prove results like these. In this chapter we look
very briefly at a proper definition for definite integrals and show that such integrals
exist, not only for continuous functions, but in fact for a wider class of functions.
Perhaps rather surprisingly, we make extensive use of suprema and infima in the
process.
179
CHAPTER 7. INTEGRATION 180
Figure 7.1: Approximating the area under the graph of a positive continuous func-
tion f .
region under the graph of f by summing the areas of a number of rectangles. These
rectangles are obtained by subdividing the interval [a, b] into a number (say n) of
subintervals using the division points x0 = a, x1 , x2 , . . . , xn = b. Each subinterval
[xi−1 , xi ] then becomes the base of a rectangle. The height of the rectangle is chosen
to be the value of the function at some point ti in the interval [xi−1 , xi ]. The sum
of the areas of the rectangles obtained in this way
n
X
f (ti )(xi − xi−1 )
i=1
then gives a (rough) approximation to the area under the graph of f . A sum like
this is called a Riemann sum.
To improve on this approximation, the argument continues, we have to increase
the number n of subintervals. It is clear (always take this phrase with at least a
pinch of salt!) that as n tends to infinity, the approximations will tend to the area
under the curve, at last if we ensure that the length of the largest subinterval also
goes to 0 as n tends to infinity. The area under the curve is therefore equal to the
limit n
X
lim f (ti )(xi − xi=1 ).
n→∞
i=1
Rb
At this stage the integral a f (x)dx is introduced as shorthand for this limit.
Note that we could (and in fact we do) formally write down the same kind of sums
and limits in the case where f is no longer positive on the whole of [a, b]. In this
case the integral can no longer be interpreted as an area.
CHAPTER 7. INTEGRATION 181
It is not difficult to expose the gaps in the argument above. Here are a few
uncomfortable questions:
• Does the limit appearing in the definition of the integral always exist (i.e for
all sequences of partitions of [a, b] and all choices of the points ti ∈ [xi−1 , xi ])?
• If the limit always exists, are all these limits equal (i.e. is the limit independent
of the choice of the sequence of partitions and the points ti ) in the subintervals?
We could get round the first question by admitting that we have not defined
what we mean by the area of a region in the plane that is not a polygon, and then
solving the problem by defining the area to be equal to the integral. This will agree
with our intuitive idea of area, but will only make sense if we know that the integral
exists, and has a unique value. So we are forced to consider the remaining two
questions.
It this stage it will pay to be more precise about definitions. We are going to
formulate the definitions in such a way that we can consider integrals of functions
that are not necessarily continuous.
Summary:
In this introductory section we raised some questions about the definition of the
definite integral usually given in introductory calculus courses.
The next step is to associate with each partition certain special approximating
sums.
CHAPTER 7. INTEGRATION 182
Definition 7.2.2 Let f be a bounded real-valued function on the interval [a, b], and
P = {x0 , x1 , . . . , xn } a partition of [a, b]. For i = 1, 2, . . . , n, put
mi = inf{f (x) : x ∈ [xi−1 , xi ]}, Mi = sup{f (x) : x ∈ [xi−1 , xi ]}.
The lower sum L(P, f ) associated with the partition P is given by
n
X
L(P, f ) = mi (xi − xi−1 )
i=1
Remarks:
1. Since the function f is bounded on [a, b], it is also bounded on each of the
intervals [xi−1 , xi ]. The sets {f (x) : x ∈ [xi−1 , xi ]} are therefore bounded, and
this ensures that mi and Mi exist.
CHAPTER 7. INTEGRATION 183
4. Since the function f is bounded on [a, b], we can find constants m and M such
that m ≤ f (x) ≤ M for all x ∈ [a, b]. It follows easily from this that (see
Figure 7.2) that for any partition P we have
It follows from this that the set of all upper sums is a non-empty set of real
numbers which is bounded below, and therefore has an infimum. In the same
way it follows that the set of all lower sums has a supremum.
Definition 7.2.3 Let f be a bounded real-valued function on the interval [a, b]. The
lower integral of f on [a, b] is defined by
Z b
f (x) dx = sup{L(P, f ) : P is a partition of [a, b]}
a
To find out what the relationship between the lower and upper integral is, we
first have to look at the relationship of upper and lower sums for different partitions.
In what follows we’ll assume that f is a bounded real-valued function on [a, b].
(a) If P2 is a refinement of P1 ,
L(P1 , f ) ≤ U(P2 , f ).
Proof:
(a) We first look at the situation where P2 contains exactly one point more than
P1 , say P1 = {x0 , x1 , . . . , xn } and P2 = {x0 , x1 , . . . , xi−1 , y, xi , . . . , xn }, with
xi−1 < y < xi (see Figure 7.3). Let mi be defined as before, and put
This shows that L(P1 , f ) ≤ L(P2 , f ) in this case. The general case (P2 contains
m more points than P1 ) can be proved by induction. The proof that U(P2 , f ) ≤
U(P1 , f ) is similar.
Theorem 7.2.5 Let f be a bounded real-valued function on the interval [a, b]. Then
Z b Z b
f (x) dx ≤ f (x) dx.
a a
Proof: If P and Q are any two partitions of [a, b], it follows from the previous
proposition that L(P, f ) ≤ U(Q, f ). Hence L(P, f ) is a lower bound for the set
{U(Q, f ) : Q is a partition of [a, b]} and so
Z b
L(P, f ) ≤ inf{U(Q, f ) : Q is a partition of [a, b]} = f (x) dx.
a
Then
Z b Z b
f (x) dx = sup{L(P, f ) : P is a partition of [a, b]} ≤ f (x) dx.
a a
The following example shows that the inequality in the theorem above can be a
strict one.
Summary:
In this section we made precise the notion of a partition of a closed bounded interval
[a, b], and also that of a refinement of a partition. With each partition and each
bounded function defined on the interval we can associate a lower and upper sum.
Refining a partition increases the lower sum and decreases the upper sum. The
supremum of the set of all lower sums and the infimum of the set of all upper sums
are respectively called the lower integral and the upper integral. The lower integral
is always less than or equal to the upper integral, but there are bounded functions
for which the two are not equal.
CHAPTER 7. INTEGRATION 186
Definition 7.3.1 Let f be a bounded real-valued function on the interval [a, b].
Then f is said to be Riemann integrable over [a, b] if
Z b Z b
f (x) dx = f (x) dx.
a a
Note that if the function f is integrable then it follows from the definitions that
Z b
L(P, f ) ≤ f (x) dx ≤ U(P, f )
a
Rb
for every partition P of [a, b], and that in fact the real number a
f (x) dx is the
unique real number with this property.
The Dirichlet function is an example of a bounded function that is not Riemann
integrable over [0, 1] (or over any interval on the real line, for that matter).
To show that there is in fact a large class of functions that are Riemann inte-
grable, we need the following criterium for integrability:
Let > 0. By our assumption we can find a partition P such that U(P, f )−L(P, f ) <
. Then Z b Z b
f (x) dx ≤ U(P, f ) ≤ L(P, f ) + ≤ f (x) dx + .
a a
using the fact that the upper and lower integrals are equal since f is integrable.
Theorem 7.3.3 If the real-valued function f is continuous on the interval [a, b], it
is Riemann integrable over [a, b].
Proof: Since the function f is continuous on the closed and bounded interval [a, b],
it is uniformly continuous on [a, b], by Theorem 5.3.26. We use Proposition 7.3.2
to show that f is Riemann integrable. Let > 0. Since f is uniformly continuous
on [a, b], there is a δ > 0 such that |f (x) − f (y)| < b−a whenever x, y ∈ [a, b] and
b−a
|x − y| < δ. Choose n ∈ N such that n < δ and let Pn be the partition of [a, b] into
n subintervals of equal length. If Mi and mi are defined as usual, for i = 1, 2, . . . , n,
then
n
X b−a
U(Pn , f ) − L(Pn , f ) = (Mi − mi )
i=1
n
n
X b−a
≤
i=1
b−a n
= .
Proof: It suffices to give a proof for a bounded increasing function; the proof for
decreasing functions is similar. For such functions it is easy to see that
(f (b) − f (a))(b − a)
U(Pn , f ) − L(Pn , f ) = ,
n
where Pn is the subdivision of [a, b] into n subintervals of equal length. It follows
from Proposition 7.3.2 that f is Riemann integrable.
Since there are bounded monotone functions that are not continuous everywhere,
it follows that the class of Riemann integrable functions is genuinely larger than the
class of continuous functions.
Summary:
A bounded function is Riemann integrable on an interval if the lower sum of the
function is equal to its upper sum. In this section we derived a necessary and
sufficient condition for a bounded function to be Riemann integrable, and use it to
show that continuous and also monotone functions are integrable.
Historical notes
Georg Friedrich Bernhard Riemann (1826 – 1866)
Riemann’s father was a Lutheran min-
ister who took responsibility for the ini-
tial schooling of his six children him-
self. He taught Bernhard until he
was ten. At school Bernhard quickly
showed an interest in mathematics. In
1846 Riemann went to the University
of Göttingen, where he at first stud-
ied theology since he was encouraged
to do so by his father. His interest
was in mathematics, however, and he
obtained permission from his father to
change course. Although Gauss taught
at Göttingen at the time, the math-
ematics department was not all that
good, and Riemann moved to Berlin in
1847. There he came under the influ-
ence of Dirichlet. Riemann liked the
latter’s intuitive way of reasoning and adopted the same style. During this time Rie-
CHAPTER 7. INTEGRATION 189
Exercises
1. Let f (x) = x2 and Pn be the partition of the interval [0, 1] into n subintervals
of equal length.
(a) For each n ∈ N, find L(Pn , f ) and U(Pn , f ).
R1
(b) Find 0 f (x) dx.
R1
(c) Find 0 f (x) dx.
(d) Use only your answers to (a), (b) and (c) to determine whether f Riemann
integrable over [0, 1]. If so, find the value of the integral.
2. Let f : [1, 4] → R be defined by
√
1≤x< √
2 if √ 2
f (x) = 1 if √2 ≤ x < 5
3 if 5 ≤ x ≤ 4
and let Pn be the partition of the interval [1, 4] into n subintervals of equal
length.
(a) Calculate L(P3 , f ) and U(P3 , f ).
(b) Calculate L(P6 , f ) and U(P6 , f ).
(c) Let n ≥ 3. Find U(Pn , f ) − L(Pn , f ) in terms of n.
(d) Is f Riemann-integrable on [1, 4]? If so, what is the value of the Riemann
integral? Give reasons for your answers.
3. Letf be a bounded real-valued function on [a, b]. Prove or disprove the follow-
ing statements:
(a) If there is a partition P of [a, b] such that L(P, f ) = U(P, f ), then f is
Riemann integrable on [a, b].
(b) If for every partition P of [a, b] we have L(P, f ) < U(P, f ), then f is not
Riemann integrable on [a, b].
4. Show that if f is Riemann integrable over [a, b] and |f (x)| ≤ M for all x ∈ [a, b],
Rb
then | a f (x) dx| ≤ M(b − a).
5. Let a < c < b and let f be a function on [a, b] defined by putting f (x) = 1 if
x = c and f (x) = 0 for all other x ∈ [a, b]. Prove that f is Riemann integrable
Rb
and that a f (x) dx = 0. Deduce that if g is Riemann integrable on [a, b] and
the function h is obtained from g by changing
Rb the value
R b of g at one point in
[a, b], then h is Riemann integrable and a h(x) dx = a g(x) dx.
CHAPTER 7. INTEGRATION 191
6. Show that if f is Riemann integrable over [a, b] and a < c < b, then f is
Riemann integrable over [a, c] and over [c, b], and
Z b Z c Z b
f (x) dx = f (x) dx + f (x) dx.
a a c
[Hint: Show that it suffices to consider partitions of [a, b] containing the point
c.]
10. Let f be Riemann integrable over [a, b] and define the function F : [a, b] → R
by Z x
F (x) = f (t) dt (x ∈ [a, b]).
a
11. Let f be Riemann integrable over [a, b] and suppose that F is a differentiable
function on [a, b] such that F 0 (x) = f (x) for every x ∈ [a, b]. Prove that
Z b
f (x) dx = F (b) − F (a).
a
[Hint: For every > 0, choose a partition P as in Proposition 7.3.2. Then use
the Mean Value Theorem.]
Appendix A
192
APPENDIX A. LOGIC AND PROOFS 193
The examples in (b), (c) and (d) show that in mathematics we quite naturally
come across sentences that contain variables, and that this disqualifies such sentences
from being statements. You may quite rightly feel unhappy about not taking such
sentences seriously. They seem to be saying something that makes sense. Perhaps
we should have a second look at what they seem to be saying. But let’s first agree on
a name for such problematical sentences: we’ll call a sentence containing variables
an open sentence. Whether such a sentence is true or false depends on the values
of the variables.
There is another bit of terminology associated with such open sentences. When
we write down an open sentence, we usually have in mind a set of elements that can
be substituted in the place of the variable(s) in the open sentence. We call this set
the universe for the sentence. Ideally we should specify the universe for the open
sentences we use; you will find that this is sometimes omitted if it is clear from the
APPENDIX A. LOGIC AND PROOFS 194
context. We have already suggested the set of real numbers as the universe for the
open sentence “x2 > x”. For the open sentence “n2 is an odd number” the universe
could be taken to be the set of integers, and for the sentence “n is a prime number”
we could use the set of natural numbers.
Associated with an open sentence and a universe there is a truth set: the set of
all the elements of the universe which when substituted in the open sentence makes
it a true statement. So, for example, 2 is in the truth set of x2 > x, but 21 is not.
The truth set in this case is the union of the intervals (−∞, −1) and (1, ∞).
But what is the use of open sentences if we cannot even decide whether they
are true or false? We need something that will change an open sentence into a
statement. If we change the open sentence “x2 > x” into
“ For all real numbers x, x2 > x”
we do get a statement, since we can decide whether it is true or false. (It is false,
since we have an example of a real number x for which it fails.)
We can also change the open sentence “x2 > x” into
“ There is a real number x such that x2 > x.
If we do this, we again get a statement, since we can decide whether it is true or
false. (It is true.)
Phrases such as “For all” and “There is (are)” are called quantifiers and they
are used to bind the variables in an open sentence to change it into a statement.
The examples above introduce the two most important kind of quantifiers. Phrases
such as for all, for every, for any are called universal quantifiers and denoted
by the symbol ∀, while phrases such as there is, there exists, for some are called
existential quantifiers and denoted by the symbol ∃.
The example (c) is a little more difficult to analyse. You will probably agree
that when we say
“If n2 is an odd number, then n is an odd number”
we really mean
“It does not matter what the integer n is,
as long as n2 is an odd number, then n is an odd number.”
Another way to say this is:
“For all integers n, if n2 is an odd number, then n is an odd number.”
Here the quantifier “For all” is used to bind the variable n in the open sentence
“if n2 is an odd number, then n is an odd number.”
Once we have done this, it becomes a statement; we can in this case show it is true
(try to do this, if you have not done so already).
We can treat the example (d) in the same way. The open sentence “If n is a
prime number, then n + 1 is a prime number” can be changed into a statement
APPENDIX A. LOGIC AND PROOFS 195
by using the universal quantifier to get “For all natural numbers n, if n is a prime
number, then n + 1 is a prime number.” It is a statement, because we can decide
whether it is true or false. In this case it is false (why?). We could also use the
existential quantifier: “There exists a natural number n such that if n is a prime
number, then n + 1 is a prime number.” In this case we get a true statement (why?).
It is unfortunately true that we tend to be somewhat sloppy about inserting
quantifiers when we state results. So, for example, you may well find that
“If n2 is an odd number, then n is an odd number.”
is stated as a result.
What is passed off as a statement when this is done is strictly speaking only an open
sentence. The result, as we have seen, should read
”For every integer n, if n2 is an odd number, then n is an odd number.”
Such sloppiness is widespread, and you will come across it often in these notes
and many mathematics texts you may read. It is assumed in such cases that it is
clear what the quantifier is that is necessary to turn the open sentence into a true
statement; almost invariably it will be some form of universal quantifier. Here is
another example:
Example A.1.2 Suppose you are asked to prove or disprove the statement “The
sum of two odd numbers is an even number.” We can write this as “If n and m
are odd numbers, then n + m is an even number.” This looks quite acceptable, but
strictly speaking it is an open sentence (containing the variables n and m), rather
than a statement. The intended statement is clearly: “For every integer n and every
integer number m, if n is an odd number and m is an odd number, then n + m is
an even number.” In this form it is clear that if we want to prove the statement, we
will have to give an argument that is valid for every odd number n and every odd
number m; an example of two odd numbers with a sum which is an even number
will not prove the statement.
You may feel that the insistence on putting in quantifiers is bordering on splitting
hairs. What makes matters worse is that we are not going to adhere to our own
strict standards in future. The reason that we are so petty now is that one should
always be aware of the intended quantifier, even when it is not there explicitly. This
becomes crucial when trying to prove that a statement is false, as we’ll see later.
Two of the connectives listed earlier need some further comment. The word “not”
is not strictly speaking a connective, since it is not used to link two statements, but
rather to change the meaning of a statement by negating it. Thus applying “not”
to the statement “x = 0” we get the statement “x 6= 0”, and applying “not” to the
statement “n is is even”, we get the statement “n is not even” (or, equivalently. “n
is odd”).
APPENDIX A. LOGIC AND PROOFS 196
The connective “if and only if” (which is often abbreviated to “iff”) is used to
indicate that two implications are both true. As an example, the sentence
When you are asked to prove a statement involving the connective “if and only if”,
keep in mind that there are two things you need to prove.
When it is made clear that what is given is a definition, the less precise first version
is usually given (and we’ll do this in these notes as well).
Most of the theorems you’ll see in this course (and elsewhere) will be statements
of the form “For every x, if P (x), then Q(x)”, where P (x) and Q(x) are open
sentences depending on the variable x. To prove that such a statement is true, we
assume that the statement P (x) is true, and show that it follows from this that the
statement Q(x) is true, making sure that our argument is valid for every x in the
universe for the open sentences P (x) and Q(x). Write out a proof of the statement
APPENDIX A. LOGIC AND PROOFS 197
“For every integer n, if n is an even number, then n2 is an even number” and make
sure that your proof satisfies this requirement.
When working with implications, the order in which we write down things is
very important. Here is an example.
Look at the statements:
(a) For every natural number n, if n is an odd number, then n2 + 1 is not a prime
number.
(b) For every natural number n, if n2 + 1 is not a prime number, then n is an odd
number.
These two statements clearly do not say the same thing. We can prove that
statement (a) is true. (Do this, using the fact that if n is odd, n2 will be odd.)
However, we cannot assume that because statement (a) is true, statement (b) will
also be true (in fact, it is false).
More generally, if P and Q are statements, then P ⇒ Q is a statement again.
We call the statement Q ⇒ P the converse of the statement P ⇒ Q.
WARNING:
is not true. How do we prove that a statement of this form is false? The statement
claims that for every natural number something is true. To show that the claim is
false, it is enough to give an example of one natural number for which that something
is not true. The “something” in this case is the implication “if n2 + 1 is not a prime
number, then n is an odd number”. We therefore need to find a natural number n,
such that n2 + 1 is not a prime number, and n is not an odd number, that is, n is
an even number. This means that to disprove the statement (∗) (that is, to show
APPENDIX A. LOGIC AND PROOFS 198
that it is false) we have to find one example of an even number n such that n2 + 1
is not a prime number. A little experimentation shows that n = 8 will do (there are
others as well).
An example like this that is used to disprove a statement containing a universal
quantifier is called a counterexample.
We claimed a little earlier that the statement
is true, and asked you to prove it. You may have discovered that it is not immediately
clear how one could start a proof. A direct proof would start by assuming that n is
a natural number such that n2 is an odd number. This means that we can find some
natural number k such that N 2 = 2k − 1. But where does one go from there? There
is a bit of logic that comes to the rescue here. First recall that ¬ Q is shorthand for
“not Q”. A statement of the form
P ⇒Q
¬ Q ⇒ ¬ P.
P ⇒ Q ≡ ¬ Q ⇒ ¬ P.
This means that if we can prove that one of the statements P ⇒ Q and ¬ Q ⇒ ¬ P
is true, the other will be true as well. We call ¬ Q ⇒ ¬ P the contra-positive of
P ⇒ Q. There are cases where it turns out to be easier to prove the contrapositive
than the statement itself.
Let’s return to the statement we are trying to prove, and write down its contra-
positive. We have not looked at what is meant by the contrapositive of a statement
that contains a quantifier. We’ll say that the contrapositive of a statement of the
form
For every x, P (x) ⇒ Q(x)
is
For every x, ¬ Q(x) ⇒ ¬ P (x) .
The contrapositive of the statement (∗∗) above then becomes
APPENDIX A. LOGIC AND PROOFS 199
or, equivalently,
This statement is in fact true. You will find this easier to prove; do this!
A closely related way of proving statements of the form P ⇒ Q is the method
of proof by contradiction. To use this method, we assume that the statement P
is true, and suppose that Q is false. Using these two assumptions, we then try to
prove a statement that is clearly false (this is the contradiction). It the only way
to explain this contradiction is that the assumption that P is true and Q is false
was wrong. Hence if P is true, Q must be true as well. There is a good example
in Chapter 1 of this method of proof: the proof that there is no positive rational
number x such that x2 = 2.
We often need to negate statements when doing a proof by contradiction, or
the contrapositive of a statement. These statements often contain implications and
quantifiers. We first look at the negation of an implication.
To say that the the statement P ⇒ Q is not true means that P is true and Q is
not true. In symbols:
¬(P ⇒ Q) ≡ P ∧ (¬Q).
• A statement of the form “ For every x, P (x)” says that for every x in some
universal set the statement P (x) is true. If this is not true, that means that
there is an x in the universal set for which P (x) is not true. The negation of
the statement “For every x, P (x)” is therefore the statement “There exists an
x such that ¬P (x)”. In symbols:
¬(∀x)(P (x))) ≡ (∃x)(¬(P (x)).
• A statement of the form “There exists an x for which P(x)” says that there is
some x in the universal set for which the statement P (x) is true. If this is not
true, that means that for every x in the universal set P (x) is not true. The
negation of the statement “There exists an x for which P (x)” is therefore the
statement “For every x, ¬P (x)”. In symbols:
¬(∃x)(P (x))) ≡ (∀x)(¬(P (x)).
APPENDIX A. LOGIC AND PROOFS 200
Example A.2.1 (a) The negation of the implication “if n2 is an odd number, then
n is an odd number” is “n2 is an odd number and n is not an odd number” or “n2
is an odd number and n is an even number”.
(b) The negation of the statement “For every n ∈ N+ , if n2 is an odd number, then
n is an odd number” is “There exists an n ∈ N+ such that n2 is an odd number and
n is an even number”.
(c) Our final example looks at the negation of a statement with more than one
quantifier. We use the definition of convergence of a real sequence to 0. If you have
not seen it before, do not worry. Simply look at the structure of the statement in
the definition, even if you do not yet understand it.
A real sequence (xn ) converges to 0 iff for every ∈ (0, ∞), there exists
an N ∈ N+ such that for every n ∈ N+ , if n ≥ N, then |xn | < .
In symbols:
(xn ) converges to 0 ⇐⇒ (∀ ∈ (0, ∞))(∃N ∈ N+ )(∀n ∈ N+ )[n ≥ N ⇒ |xn | < ].
To say what it means for the sequence (xn ) not to converge to 0, we have to negate
the statement following the “iff”. If we do this we get:
(xn ) does not converge to 0 iff there exists an ∈ (0, ∞) such that for
every N ∈ N+ , there exists an n ∈ N+ such that n ≥ N and |xn | ≥ .
In symbols:
(xn ) does not converges to 0 ⇔ (∃ ∈ (0, ∞))(∀N ∈ N+ )(∃n ∈ N+ )(n ≥ N)∧(|xn | ≥ ].
Appendix B
Rational numbers
We have initially defined the set of rational numbers as the set {p/q : p ∈ Z, q ∈ N+ },
and addition and multiplication by:
One could raise two objections to this. The first is that it is not clear what p/q
means. The second is that we would like, for example, 1/2 and 2/4 to be regarded
as equal. We can get around the first difficulty by replacing p/q by the ordered pair
(p, q), with p ∈ Z and q ∈ N+ . Let us write Q for this set, i.e.
Q = {(p, q) : p ∈ Z, q ∈ N+ }.
The way to get around the second difficulty is to introduce a new definition for
equality in the set of all such pairs:
We can then check that ∼ is an equivalence relation. It follows that the equivalence
classes of this equivalence relation is then a partition of Q (i.e. every element (p, q)
of Q belongs to exactly one equivalence class; we’ll denote this equivalence class by
201
APPENDIX B. RATIONAL NUMBERS 202
[(p, q)]. The set of rational numbers Q is then defined to be the set of all these
equivalence classes:
[(p, q)] + [(r, s)] = [(ps + qr, qs)], [(p, q)][(r, s)] = [(pr, qs)].
Note that we are adding and multiplying equivalence classes here. The first thing
that needs to be checked now is that these definitions of addition and scalar mul-
tiplication do not depend on the representatives that we choose from equivalence
classes. For example, we must show that if [p, q)] = [(p1 , q1 )] and [(r, s)] = [(r1 , s1 )],
then [(ps + qr, qs)] = [(p1 s1 + q1 r1 , q1 s1 )]. Once we have done this, it is then possible
to check the properties A1 – A4, M1 – M4 and D for Q using the properties of the
integers.
We can define an order relation on Q (i.e. an order relation on equivalence classes
of ordered pairs) by
We will have to check again that this definition does not depend on the representa-
tives we choose from the equivalence classes. The properties O1 – O4 can then be
checked, using the order properties of the integers.
The above gives a very brief sketch of the construction of the rational numbers,
using the integers and their properties. There is a fair amount of work to be done
to check all the properties, but the checking is not unduly complicated.
One could of course go even further back, and ask how the integers are con-
structed. It is possible to start with a set of axioms for the natural numbers N
(the Peano axioms) and use these to define addition and multiplication on N, and
then to prove the usual properties of the natural numbers. Then the integers can
be constructed as equivalence classes of ordered pairs of natural numbers, in a fash-
ion somewhat similar to the method we indicated above for the construction of the
rational numbers. If you are interested, you will find the detail in many algebra
books.
Index
203
INDEX 204
telescoping sum, 90
test
comparison, 104
integral, 112
Leibniz, 95
limit ratio, 108
ratio, 111
root, 111
Weierstrass M, 172
theorem
mean value, 156
Bolzano-Weierstass, 72
Cauchy mean value, 159
intermediate value, 150
Rolle’s, 155
truth set, 194
unbounded, 27
uniform convergence, 164
uniformly continuous, 152
universe, 193