Introduction To Real Analysis
Introduction To Real Analysis
Ole A. Nielsen
August 1997
Andrew D. Lewis
2017/04/06
Contents
0 Mathematical notation and logic 1
0.1 Mathematical notation . . . . . . . . . . . . . . . . . . . . . . . . . . 1
0.1.1 Sets and elementary set notation . . . . . . . . . . . . . . . . 1
0.1.2 Notation for number systems . . . . . . . . . . . . . . . . . . 2
0.2 Mathematical logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
0.2.1 Propositions and symbolic logic . . . . . . . . . . . . . . . . . 3
0.2.2 Predicates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
0.2.3 Quantifiers . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
0.2.4 Assertions and proofs . . . . . . . . . . . . . . . . . . . . . . 5
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3 Sequences in Rd 45
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4 Some topology in Rd 50
4.1 Open and closed sets . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
4.2 Cluster points, limit points, and boundary points . . . . . . . . . . . 54
4.3 Properties of closed sets . . . . . . . . . . . . . . . . . . . . . . . . . 56
iii
5 Infinite series 69
5.1 Definitions and examples . . . . . . . . . . . . . . . . . . . . . . . . 69
5.2 Convergence of nonnegative series . . . . . . . . . . . . . . . . . . . 72
5.3 Convergence of general series . . . . . . . . . . . . . . . . . . . . . . 77
5.4 Estimating remainders . . . . . . . . . . . . . . . . . . . . . . . . . . 82
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
6 Sequences of functions 91
6.1 Motivating example . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
6.2 Pointwise convergence . . . . . . . . . . . . . . . . . . . . . . . . . . 92
6.3 Uniform convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
6.4 Properties of uniform convergence . . . . . . . . . . . . . . . . . . . 98
6.5 The sup-norm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
Index 139
Chapter 0
Mathematical notation and logic
In this chapter, we provide the briefest of overviews of the basic mathematical no-
tation we shall use, and of mathematical logic, as this course is intended to introduce,
not just mathematical ideas, but also techniques of proof.
image(f ) = {f (x) | x ∈ S} ⊆ T.
1. [a, b] = {x ∈ R | a ≤ x ≤ b} 5. [a, ∞) = {x ∈ R | a ≤ x}
2. [a, b) = {x ∈ R | a ≤ x < b} 6. (a, ∞) = {x ∈ R | a < x}
3. (a, b] = {x ∈ R | a < x ≤ b} 7. (−∞, b] = {x ∈ R | x ≤ b}
4. (a, b) = {x ∈ R | a < x < b} 8. (−∞, b) = {x ∈ R | x < b}
9. (−∞, ∞) = R
0.2 Mathematical logic 3
However, we leave these details for a proper course in mathematical logic. Note
that using these rules ¬(p =⇒ q) ⇐⇒ (p ∧ ¬q).
2. Negation of an implication: An important application of the rules above occurs
when we wish to negate an implication. ¬(p =⇒ q) ⇐⇒ ¬(¬p∨q) ⇐⇒ p∧¬q.
0.2.2 Predicates
Let X be a set. A predicate p in X assigns to each x ∈ X the value true or
false. Note that, for each x ∈ X, p(x) is a proposition since it is either true or
false. Thus a predicate can be called a “propositional function.” The variable x is
the subject of the predicate. Often p(x) is thought of as being a “property” of x, as
will be clear from the examples below.
Predicates are often used to form conditional propositions, i.e., “if/then” state-
ments. Thus, if p and q are predicates in X, one has a proposition p(x) =⇒ q(x).
This means that, if x ∈ X is such that p(x) is true, then so too is q(x) true. As a
proposition itself, p(x) =⇒ q(x) may be true or false.
Here are some examples.
0.1 Examples: 1. Take X = Z>0 and consider the predicates p and q defined
by p(n)=“n is odd” and q(n)=“n is prime”. Thus p(1)=true, p(3)=true,
q(1)=false, and q(3)=true. Note that both of the propositions p(n) =⇒ q(n)
and q(n) =⇒ p(n) are false.
2. Let X be the set of maps f : [a, b] → R. Consider the two predicates p(f )=“f is
continuous” and q(f )=“there exists c ∈ (a, b) such that f (c) = 0”. The interme-
diate value theorem says that p(f ) =⇒ q(f ) is true.
3. Let p(x) be the predicate in R given by “x is such that 1 − x2 = 0.” This is true
if x = ±1 and false otherwise. •
0.2.3 Quantifiers
Associated with predicates are some useful logical constructions called “quanti-
fiers.”
1. Existential quantifier: ∃ x p(x) means that there is x such that p(x) is true. For
example, ∃ x x2 − 1 = 0 means that there is an x such that x2 − 1 = 0. We might
find it necessary to be more specific about where x is as in ∃ x (x ∈ R)∧(x2 −1 = 0),
which we shall write as ∃ x ∈ R, x2 − 1 = 0. We shall also write @x ∈ R, x2 + 1 = 0
for ¬(∃ x ∈ R, x2 + 1 = 0). Finally, we shall write ∃! x to mean there exists a
unique x, as in ∃! x ∈ R, x3 − 1 = 0.
2. Universal quantifier: ∀ x p(x) means that, for all x, p(x) is true. For example,
∀ x ∈ R x2 ≥ 0 means that the square of every real number is nonnegative.
In constructions such as p(x), x is a free variable; whereas in ∀ x p(x), x is a bound
variable. A bound variable is like a dummy variable in calculus: the proposition
∀ x p(x) is logically equivalent to ∀ y p(y).
0.2 Mathematical logic 5
The quantifiers ∀ and ∃ are related by the rule ¬(∃ x p(x)) ⇐⇒ ∀ x ¬p(x) and
¬(∀ x p(x)) ⇐⇒ ∃ x ¬p(x). Indeed, one can define ∃ in terms of ∀ and ¬.
Let us give an example that illustrates how a well-known notion, that of continuity
of a function (see Section 4.4), can be expressed using our above symbolic logic
terminology.
0.2 Example: Let f be a function and p(x) be the proposition
∀ > 0 ∃ δ > 0 ∀ y |x − y| < δ =⇒ |f (x) − f (y)| <
This is the way of saying that f is continuous at x.
Let us use our rules to find ¬p(x):
¬p(x) ⇐⇒ ¬(∀ > 0 ∃ δ > 0 ∀ y |x − y| < δ =⇒ |f (x) − f (y)| < )
⇐⇒ ∃ > 0 ¬(∃ δ > 0 ∀ y |x − y| < δ =⇒ |f (x) − f (y)| < )
⇐⇒ ∃ > 0 ∀ δ > 0 ¬(∀ y |x − y| < δ =⇒ |f (x) − f (y)| < )
⇐⇒ ∃ > 0 ∀ δ > 0 ∃ y¬(|x − y| < δ =⇒ |f (x) − f (y)| < )
⇐⇒ ∃ > 0 ∀ δ > 0 ∃ y |x − y| < δ ∧ |f (x) − f (y)| ≥
Thus ¬p(x) is how we write the proposition that f is not continuous at x. •
Direct proof
Here we wish to prove that, given propositions p and q, p =⇒ q. The method of
direct proof is that one uses p the fact that p is true, with no other assumptions, to
prove that q is true.
We illustrate this with a few simple examples.
0.3 Examples: 1. Consider the proposition p given by “4 is an even number.” To
prove that p is true, we note that 4 = 2 · 2, and so 4 is even, being an integer
multiple of 2.
2. Consider the proposition p given by “4 is a prime number.” To prove that p is
false, we note that 4 = 2 · 2, and so 4 has a factor other than 1 and 4.
3. We take p(n) to be the predicate with integer subject “n is a square of odd
numbers” and q(n) to be the predicate with integer subject “n is an odd number.”
We will show that p(n) =⇒ q(n). Suppose that n is such that p(n) is true.
Then n = m2 for an odd number m. Thus m = 2k + 1 for an integer k and
Proof by contradiction
For any proposition p, the proposition p ∧ ¬p is always false. If, in a proof, one shows
that a certain hypothesis implies the proposition p ∧ ¬p for some p, then we say we
have reached a contradiction and our hypothesis must be false. Sometimes, when
we wish to prove p is true, we assume that p is false and then reach a contradiction,
say q∧¬q. That is, we assume ¬p and then reach a contradiction; thus we have proved
¬(¬p)), i.e., p is true.
We give an example of proof by contradiction.
√ √
0.4 Example: We prove that 2 is irrational. Suppose that 2 = p/q for some
integers p and q with p and q relatively prime, i.e., having no common factor. Then
p2 = 2q 2 . So p2 is even and hence p is even and thus p is divisible by 2. Hence
p2 = (p0 )2 · 22 for p0 = p/2. We can now cancel a 2 to obtain 2(p0 )2 = q 2 and conclude
that q is also even. This contradicts our assumption that p and q are relatively prime.
Hence we have reached a contradiction. √ Thus our assumption that √ there were integers
p and q, relatively prime, such that 2 = p/q was false, i.e., 2 is irrational. •
Proof of contrapositive
We saw that p =⇒ q was logically equivalent to ¬p ∨ q, which is the same as
¬(¬q) ∨ ¬p, which finally is the same as ¬q =⇒ ¬p. This last expression is called
the contrapositive of p =⇒ q and we have just seen that (p =⇒ q) ⇐⇒
(¬q =⇒ ¬p).
Let us give an example of proof by contrapositive.
0.2 Mathematical logic 7
0.5 Example: Let X = Z>0 ×Z>0 ×Z>0 , and consider the predicates p(n1 , n2 , n3 ) and
q(n1 , n2 , n3 ) in X given by “the product of n1 , n2 , and n3 is greater than 1000” and “at
least one of n1 , n2 , and n3 is greater than 10.” We will prove that p(n1 , n2 , n3 ) implies
q(n1 , n2 , n3 ) by proving the contrapositive. Then we assume that ¬q(n1 , n2 , n3 ) is
true, meaning that we assume that n1 , n2 , n3 ≤ 10. Then
n1 · n2 · n3 ≤ 10 · 10 · 10 = 1000.
Now, for the last statement, we can give a direct proof: given x > y, let = (x−y)/2,
then > 0 and y + = (x + y)/2 < x as required. •
Proof by induction
It is common to have situations where one has an assertion that is said to hold for
every n ∈ Z>0 . That is, one has a predicate p in Z>0 , and one wishes to prove the
proposition ∀ n ∈ Z>0 p(n). Frequently such propositions are proved by induction.
The way this works is as follows. The assertion p(1) is proved, somehow, someway.
Then the proposition p(n) is assumed; this is called the induction hypothesis.
Using the fact that p(n) is true, p(n + 1) is proved to be true, somehow, someway.
In this way we prove the proposition ∀ n ∈ Z>0 p(n). Another strategy that is rather
similar is strong induction, where the induction hypothesis is replaced with the
strong induction hypothesis, which is that p(k) is true for k ∈ {1, . . . , n}, and
then it is proved that p(n + 1) is true.
Let’s look at a couple of examples.
0.7 Examples: 1. Consider the predicate p in Z>0 given by
Nonconstructive proof
In a nonconstructive proof, one proves a statement about the existence of something,
but without giving a specific instance of it, or an algorithm for finding it.
Here are two examples.
0.8 Examples: 1. If there are 367 people in a room, then at least two of the people
share the same birthday. A nonconstructive proof is immediate: if the mapping
from “person” to “birth date” is one-to-one, this would mean that there are at
least 367 days in a year, which cannot be. However, in this case a constructive
proof is also possible by interviewing all people in the room and establishing which
have the same birthday.
2. The intermediate value theorem asserts that, if f : [a, b] → R is a continuous
function for which f (a) < 0 and f (b) > 0, then there exists c ∈ (a, b) such that
f (c) = 0. While this theorem is true, a proof does not give any indication of what
c is or how to find it. •
Proof by counterexample
A counterexample is typically used to show that a proposition is not true. Normally
a counterexample is most useful in a situation where the proposition is of the form
p(x) for a predicate in a set X. To show that ∀ x ∈ X p(x) = true is true, one
must do exactly that: show that it is true for all x. To show that ∀ x p(x) is false,
one need only show that ∃ x ∈ X p(x) = false, i.e., a single x ∈ X is all that is
required, and this is said to be a counterexample.
Here are some examples.
0.9 Examples: We show that the converses to some propositions proved above are
false, by finding counterexamples.
1. We consider the predicates in Z>0 given by
and recall that in Example 0.3–3 we showed that p(n) =⇒ q(n) is true. We
now show by counterexample that q(n) =⇒ p(n) is false. Indeed, q(7) is true,
while p(7) is false, and so q(n) =⇒ p(n) is false.
2. We take X to be the set of mappings f : [a, b] → R and consider the two predicates
p(f )=“f is continuous” and q(f )=“there exists c ∈ (a, b) such that f (c) = 0”.
As remarked in Example 0.1–2, the proposition p(f ) =⇒ q(f ) is true. We will
show by counterexample that the proposition q(f ) =⇒ p(f ) is false. Indeed,
the mapping f ∈ X defined by
(
1, x = 21 (a + b),
f (x) =
0, otherwise
10 0 Mathematical notation and logic
satisfies q(f ) = true and p(f ) = false, and so the proposition q(f ) =⇒ p(f ) is
false. (Note: We should assume that a < b for this counterexample to be valid.
Indeed, if a > b, then X = ∅, and so q(f ) =⇒ p(f ) is true, since q(f ) is never
true. Also, the proposition is actually true if a = b.) •
Exercises for Chapter 0 11
Exercises
E0.1 Let A be a set of real numbers. What does the following proposition say?
Is there a set A for which this proposition is true? Is there a set A for which
it is false?
E0.2 Let A = [0, 1]. Let p(x) be the proposition
(∀ a ∈ A a ≤ x) ∧ (∀ b ∈ R (∀ a ∈ A a ≤ b) =⇒ x ≤ b)
The basic ingredient in “real analysis” is the set of real numbers. While we may
well have some understanding of what these are—we are taught about the “number
line” in school—a full and complete understanding of the crucial attributes of the
real numbers may not be entirely familiar. While we understand that the algebraic
properties of addition, subtraction, multiplication, and division are important, these
fall into the category of “algebra.” What concerns us here is “analysis,” and this
requires the property of the real numbers known of as “completeness.” This property
can be arrived at in a variety of ways, and we shall sketch two of these, the “least upper
bound property” and “convergence of Cauchy sequences.” Ultimately we arrive at
the former of these as the starting point for the properties of the real numbers that
allows one to do analysis. We do this by giving an overview, without all details,
of the manner in which one builds the real numbers “from nothing.” To do this
properly takes more space and effort than we can give here, but to not mention this
at all is rather unsatisfying. Thus we attempt to tread the line between making the
construction clear, but without making it excessively burdensome.
0=∅
1 = 0 ∪ {0} = ∅ ∪ {∅} = {∅} = {0}
2 = 1 ∪ {1} = {0} ∪ {1} = {0, 1}
3 = 2 ∪ {2} = {0, 1} ∪ {2} = {0, 1, 2}
..
.
Note that 0 is a set with zero elements, 1 is a set with one elements, and so on. By
Z≥0 we denote the set whose elements are all the results of the preceding recursive
construction. We call this the set of nonnegative integers. Note that the lan-
guage here is confusing, since it seems to presuppose that we know what the integers
are. However, keep in mind that it is the “nonnegative integers” (all one notion) we
1.2 Construction of the integers 13
Thus we define addition and multiplication only knowing addition and multiplication
in Z≥0 . Note that we can now define subtraction by
[(j1 , k1 )] − [(j2 , k2 )] = [(j1 , k1 )] + (−1) · [(j2 , k2 )].
One can verify, first of all, that this corresponds to the usual notions we have in mind
for addition and multiplication. Then one can tediously verify that all commutativity,
associativity, and distributivity properties we know about addition and multiplication
are valid.
One can also order the integers in the usual way. This is done with our abstract
definition by defining
[(j1 , k1 )] < [(j2 , k2 )] ⇐⇒ j1 + k2 < k1 + j2 ,
[(j1 , k1 )] ≤ [(j2 , k2 )] ⇐⇒ j1 + k2 ≤ k1 + j2 .
Again, one can work through the definitions to see that this corresponds to the
expected notion of “less than” and “less than or equal to,” and that it satisfies the
usual properties we know about these relations.
Finally, we define the absolute value function on Z by
k,
0 < k,
|k| = 0, k = 0,
−k, k < 0.
This is the expected thing, and has all the expected properties, although, again, one
must tediously prove them.
One can verify easily that these are the usual operations for rational numbers, and
that they satisfy the expected properties of commutativity, associativity, and dis-
tributivity. We also define subtraction in the expected way:
These are easily verified to be the usual “less than” and “less than or equal to,” and
they have the expected properties. We denote by Q>0 the set of rational numbers
greater than 0.
Finally, we define the absolute value function on Q by
[(j, k)],
[(0, 1)] < [(j, k)],
|[(j, k)]| = [(0, 1)], [(j, k)] = [(0, 1)],
−[(j, k)], [(j, k)] < [(0, 1)].
Again, one can verify that the usual properties of absolute value are valid.
Now we have at hand the set Q of rational numbers, with its expected operations
of addition, subtraction, multiplication, and division, with its expected relations <
and ≤, and with its absolute value function. Our next step is to build the real
numbers. There are two common ways of building the real numbers, and we shall
sketch both. They turn out to be equivalent, and this is something we will outline in
Chapter 2.
One can easily see that A is bounded if and only if it has both an upper and lower
bound (Exercise E1.3).
The subjects of least upper bound and greatest lower bound are dominated by
two important questions:
1. If a set A is bounded, does it always possess a greatest lower bound and a least
upper bound?
2. If a set A possesses a least upper bound or greatest lower bound, is it unique?
The answer to these questions is, “No,” and, “Yes.” Just by using properties of Q
as enumerated so far, and not presupposing the existence of the real numbers, it is
not easy to prove that a bounded set might not possess, say, a least upper bound.
However, it one presupposes the existence of the real numbers, this is easy to see. So
let us consider this for the moment.
√ √
1.3 Example: Take A = {x ∈ Q | x ≤ 2}. We know that 2 is irrational by
Example 0.4. Now√suppose √ for√A. Thus x ≤ u for every
√ that u is any upper bound
x ∈ A. Since u > 2, u − 2 > 0. Thus u − 12 (u − 2) > 2 and so is also an upper
bound. Moreover, it is smaller than u, and so there can no no least upper bound. •
It is possible to make the preceding example make sense without presupposing
the existence of the real numbers, but that would take too much effort. And, in any
case, it is something that is at least intuitively clear.
The main point, however, is that the quite intuitive property of existence of least
upper bounds is not true in Q. It is true, however, for the set of real numbers.
in the preceding section, this is easy to imagine if one presupposes the real numbers,
but is not so easy to prove otherwise. We thus consider an example of a nonconvergent
Cauchy sequence in Q.
1.5 Example: For n ∈ Z>0 , consider the rational number qn = (1 + n1 )n . Perhaps
the reader has run into the sequence (qn ), and indeed it converges to the real number
e ≈ 2.718281828 . . ., i.e., to Euler’s constant. Since it is a convergent sequence of
real numbers, it is also a convergent sequence of rational numbers (convergence for
real numbers is defined in Definition 2.4, but, anyway, is exactly like for rational
numbers). Therefore, it is a Cauchy sequence of rational numbers. However, since
it’s limit in the set of real numbers is e, which is not rational, it cannot converge in
Q. •
It is possible to come up with nonconvergent Cauchy sequences without making
reference a priori to the set of real numbers, but it is simply difficult to do this.
This lack of necessity of Cauchy sequences converging is problematic, since Cauchy
sequences “seem like” they want to converge., and it would be ever so convenient if
they did.
A = {q ∈ Q | q · q < 2}, B = Q \ A.
A = {r ∈ Q | r < q}, B = Q \ A. •
Note that a Dedekind cut is effectively defined by one of the sets A and B, and
in what follows we will work with the set A.
We need to define the standard constructions on R using this definition. First we
define the order relations “less than” and “less than or equal to” as follows:
A1 < A2 ⇐⇒ A1 ⊂ A2 , A1 ≤ A2 ⇐⇒ A1 ⊆ A2 .
A1 + A2 = {a1 + a2 | a1 ∈ A1 , a2 ∈ A2 },
{a1 · a2 | a1 ∈ A1 , a2 ∈ A2 , a1 , a2 ≥ 0} A1 , A2 ≥ 0,
∪{q ∈ Q | q ≤ 0},
A1 · A2 = −(A1 · (−A2 )), A1 ≥ 0, A2 < 0,
−((−A1 ) · A2 ), A1 < 0, A2 ≥ 0,
(−A ) · (−A ),
1 2 A1 , A2 < 0.
One can show that all these operations satisfy the expected properties. We leave
these rather tedious exercises to the reader.
Two Cauchy sequences (qn ) and (rn ) of rational numbers are equivalent if, for
each ∈ Q>0 , there exists N ∈ Z>0 such that |qn − rn | < for every n ≥ N . The idea
is that equivalent Cauchy sequences have tails that get close to one another. Given a
Cauchy sequence (qn ), we denote by [(qn )] the set of all Cauchy sequences equivalent
to (qn ). We then define R, the set of real numbers, to be the set
This construction, then, includes in R all Cauchy sequences, even those that do not
converge. In this way, it is a natural sort of definition.
One now has to show that R has all of the attributes one is familiar with. First,
we indicate how we add and multiply in R:
One has many things to verify here. First of all, the definitions need to be shown to be
independent of the choice of equivalent Cauchy sequences. Second, one should show
that the righthand sides of the preceding definitions are, in fact, Cauchy sequences.
After this, one should show that addition and multiplication have the expected prop-
erties.
We can define the usual order relations on R as well. Specifically, we write [(qn )] <
[(rn )] if there exists N ∈ Z>0 such that qn < rn for n ≥ N . We also write [(qn )] ≤ [(rn )]
if either [(qn )] < [(rn )] or [(qn )] = [(rn )]. Of course, these order relations have all of
the expected properties.
Finally, we define the absolute value function on R. For this, we note that, if (qn )
is a Cauchy sequence, then (|qn |) is a Cauchy sequence. We then define
The absolute value function has all of the properties we expect. One of the most
important of these is the following:
real numbers we construct in each case amount to the same thing. The way one
normally proceeds is by selecting one of the constructions as the starting point, and
proves everything with this as a starting point. That is to say, one works with
either (1) all bounded sets in R have a least upper bound or (2) all Cauchy sequences
in R converge. In this text we will be assuming that all bounded sets have a least
upper bound, and proving facts with this as a starting point.
However, our construction of the reals by Dedekind cuts above is precisely designed
so that the following theorem holds.
1.9 Theorem: Every nonempty subset of R that is bounded above has a least upper
bound, i.e., an upper bound that is smaller than any other upper bound. Similarly,
every nonempty subset of R that is bounded below has a greatest lower bound, i.e., a
lower bound that is greater than any other lower bound.
This theorem justifies the following definition.
1.10 Definition: Suppose that A is a nonempty subset of R.
(i) If A is bounded above then the smallest upper bound for A is called the least
upper bound or supremum of A and is denoted by sup A.
(ii) If A is bounded below then the largest lower bound for A is called the greatest
lower bound or infimum of A and is denoted by inf A. •
The proof of the following result is both easy and instructive.
1.11 Proposition: If A and B are two nonempty subsets of R that are bounded
above and if A ⊆ B then sup A ≤ sup B.
Proof: If a is any element of A then a ∈ B since A ⊆ B, and thus a ≤ sup B since
sup B is an upper bound for B. But this means that sup B is an upper bound for A,
and therefore sup A ≤ sup B by the definition of sup A as the smallest upper bound
of A.
The next result gives a useful characterization of the supremum of those sets that
are bounded above.
1.12 Theorem: If A is a nonempty subset of R that is bounded above and if s ∈ R
is an upper bound for A, then s = sup A if and only if
(i) a ≤ s for all a ∈ A and
(ii) A ∩ (s − , s] 6= ∅ for all > 0.
Proof: Suppose first that s = sup A. Then s is an upper bound for A and thus (i)
holds. If (ii) were false there would be an > 0 such that A ∩ (s − , s] = ∅. Then,
since s is an upper bound for A and since A ∩ (s − , s] = ∅, s − would also be an
upper bound for A. But this contradicts the fact that s is the least upper bound for
A, and therefore (ii) must be true.
Now suppose that, conversely, (i) and (ii) are true. Then (i) implies that s is an
upper bound for A. If s were not the least upper bound for A then there would be an
upper bound t for A such that t < s. Thus a ≤ t for all a ∈ A and so if = (s − t)/2
then t + = s − and a < t + for all a ∈ A, and thus A ∩ (s − , s] = ∅. This
contradicts (ii) and shows that s = sup A.
Theorem 1.12 implies the following assertion that is sometimes useful.
1.13 Corollary: If A is a nonempty subset of R that is bounded above and if s =
sup A then either s ∈ A or else A ∩ (s − , s) 6= ∅ for all > 0.
22 1 The real numbers
Exercises
E1.1 For j1 , j2 , k ∈ Z, prove the distributive rule (j1 + j2 ) · k = j1 · k + j2 · k.
E1.2 Show that the relations < and ≤ on Z have the following properties:
1. [(0, j)] < [(0, 0)] for all j ∈ Z>0 ;
2. [(0, j)] < [(k, 0)] for all j, k ∈ Z>0 ;
3. [(0, j)] < [(0, k)], j, k, ∈ Z≥0 , if and only if k < j;
4. [(0, 0)] < [(j, 0)] for all j ∈ Z>0 ;
5. [(j, 0)] < [(k, 0)], j, k ∈ Z≥0 , if and only if j < k;
6. [(0, j)] ≤ [(0, 0)] for all j ∈ Z≥0 ;
7. [(0, j)] ≤ [(k, 0)] for all j, k ∈ Z≥0 ;
8. [(0, j)] ≤ [(0, k)], j, k, ∈ Z≥0 , if and only if k ≤ j;
9. [(0, 0)] ≤ [(j, 0)] for all j ∈ Z≥0 ;
10. [(j, 0)] ≤ [(k, 0)], j, k ∈ Z≥0 , if and only if j ≤ k.
E1.3 Show that a subset A ⊆ Q is bounded if and only it is has a lower bound and
an upper bound.
E1.4 For each of the following subsets of R, find the least upper bound and greatest
lower bound if they exist. n n o
(d) n ∈
Z >0
n+1
(a) {1, 3} n
n (−1) o
(b) [0, 4] (e) n+ n ∈ Z>0
n1 o n
∞
(c) ∈ Z>0
\ 1 1
n (f) 1 − ,1 +
n=1
n n
E1.5 Find the least upper bound and the greatest lower bound of the sets below.
For any given > 0 find a number in the set that is greater than sup A − and
a number in the set that is smaller than inf A + .
n4 + x o
(a) A = x≥1
x
n √x − 1 o
(b) A = x ≥ 2
x
2
(c) A = {x | x − x < 6}
E1.6 If S = {x ∈ R | x2 + x < 3}, find sup S and inf S.
Hint: Draw a graph.
n n o
E1.7 Let A = n ∈ Z>0 and notice that sup A = 1. For = 0.01 and for
n+1
= 0.0001 find an element of A larger that 1 − .
E1.8 Suppose A is a nonempty bounded subset of R and let −A denote the set
{−x | x ∈ A}. Show that sup(−A) = − inf A.
E1.9 For each nonempty subset S of R and each real number k, let kS = {ks | s ∈
S}.
Exercises for Chapter 1 23
(a) Suppose that S = {−1, 2, 4} (so that 3S = {−3, 6, 12} and −2S =
{2, −4, −8}). Find sup S, sup 3S, sup −2S, inf S, inf 3S, and inf −2S.
Do you notice a pattern?
(b) If S is any nonempty bounded subset of R show that:
1. If k ≥ 0 then sup kS = k · sup S and inf kS = k · inf S.
2. If k < 0 then sup kS = k · inf S and inf kS = k · sup S.
E1.10 Prove that, if a subset S of R has a maximal element s (that is, there exists
an s ∈ S such that x ≤ s for all x in S), then s = sup S.
E1.11 Prove that x = sup{q ∈ Q | q < x} for each x ∈ R.
E1.12 Suppose that A and B are two nonempty bounded subsets of R and let
A + B = {x + y | x ∈ A and y ∈ B}.
Show that
sup(A + B) = sup A + sup B.
E1.13 Let A and B be two nonempty sets of real numbers that are bounded above
and put a = sup A, b = sup B, and
E1.17 Let S denote the set consisting of all those numbers with decimal expansions
of the form x = 0.a1 a2 a3 . . ., where all but finitely many of the digits a1 , a2 , . . .
are 5 or 6. Find sup S and inf S.
E1.18 Suppose that P is a subset of [0, ∞) with the property that, for any integer k,
there is an xk ∈ P such that kxk ≤ 1. Prove that 0 = inf P .
E1.19 If P and Q are two subsets of R such that sup P = sup Q and inf P = inf Q
does it follow that P = Q? Support your answer with either a proof or a
counterexample.
E1.20 For any two nonempty subsets of R let us write P ≤ Q if, for each x ∈ P ,
there is a y ∈ Q satisfying x ≤ y.
(a) Show that, if P ≤ Q, then sup P ≤ sup Q.
(b) Give an example to show that, if P ≤ Q, then it does not follow that
inf P ≤ inf Q.
(c) Give an example to show that, if P ≤ Q and if Q ≤ P , then it does not
follow that P = Q.
E1.21 Let S be a nonempty subset of R that is bounded below. Show that inf S =
sup{x ∈ R | x is a lower bound for S}.
E1.22 Prove that every nonnegative real number has a nonnegative square
root, i.e., prove that, if x is any nonnegative real number, then there is non-
negative number y such that y 2 = x.
Hint: Consider the set sup{u ∈ R | u2 ≤ x}.
E1.23 One of the important consequences of Theorem 1.9 is the Archimedean prop-
erty of R: The set Z>0 of positive integers is not bounded above in R. Although
this property may seem obvious, its proof actually depends on Theorem 1.9.
Prove the Archimedean property of R.
E1.24 Suppose that D is a nonempty subset of R and that f and g are two functions
from D to R such that f (x) ≤ g(y) for all x, y ∈ D. Show that f (D) is
bounded above, that g(D) is bounded below, and that sup f (D) ≤ inf g(D).
(Recall that f (D) = {f (x) | x ∈ D}.)
E1.25 Let D be a nonempty subset of R and suppose that f and g are two real-valued
functions defined on D. Let f + g denote the function on D defined by the
formula (f + g)(x) = f (x) + g(x) for all x ∈ D.
(a) If f (D) and g(D) are bounded above prove that (f + g)(D) is bounded
above and that sup(f + g)(D) ≤ sup f (D) + sup g(D).
(b) Show by means of an example that the inequality in (a) may be strict.
(c) State and prove the analog of (a) for infima.
Chapter 2
Sequences of real numbers
Sequences arise in a variety of ways in real analysis. They show up, for example,
when considering convergence of infinite series, such as we shall discuss in Chapter 5.
What is less clear to a newcomer, however, is that sequences in some way characterize
the very fabric of the set R of real numbers, and also its multi-dimensional version
Rd .
obtained by joining every pair of points among n points in general position on the
circumference of a circle for n = 2, 3, 4, 5, . . .).
Beware of the difference between a sequence and the range of that sequence. The
range of a sequence is analogous to the range of a function, i.e., the set of values
attained by the sequence (or function). The following two examples illustrate this.
2.2 Examples: 1. The sequence 1, 2, 3, 1, 2, 3, 1, 2, 3, . . . has an infinite number of
terms, but its range contains only the three points 1, 2, 3.
2. The range of the sequence (sin nπ
2
) is the set {−1, 0, 1}. •
2.3 Definition: A sequence (an ) is said to be bounded if there exist an M > 0 such
that |an | ≤ M for all n. •
Notice that, according to this definition, a sequence is bounded if and only if its
range is a bounded set in the sense of Definition 1.7.
2.4 Definition: A sequence (an ) is said to have a limit L if, for any > 0, there
exists a number N (depending on ) such that
|an − L| <
for every integer n ≥ N . If (an ) has a limit L, then we say that the sequence is
convergent and that it converges to L and we write limn→∞ an = L. If (an ) is not
convergent, it is said to diverge or to be divergent. If, for every c > 0, there exists
a number N such that an > c for all n ≥ N , we will write limn→∞ an = ∞ and say
that (an ) diverges to ∞. Similarly, if, for every c > 0, there exists a number N
such that an < −c for all n ≥ N , we will write limn→∞ an = −∞ and say that (an )
diverges to −∞. •
minimum and a maximum; let A and B be the minimum and maximum, respectively.
So if m = min{A, L − 1} and if M = max{B, L + 1}, then clearly m ≤ an ≤ M for
all n and so (an ) is bounded.
Example 2.5–2 shows that the converse of Proposition 2.6 is false and so bound-
edness is necessary but not sufficient for convergence. To obtain a sufficient condition
for convergence, we introduce the following definition.
2.7 Definition: A sequence (an ) is said to be nondecreasing if am ≤ an whenever
m < n and to be strictly increasing if am < an whenever m < n. Similarly, the
sequence is said to be nonincreasing if am ≥ an whenever m < n and to be strictly
decreasing if am > an whenever m < n. Finally, (an ) is said to be monotone if it
is either nondecreasing or nonincreasing. •
Note that a strictly increasing sequence is nondecreasing and a strictly decreasing
sequence is nonincreasing. The point of this definition is to have a formal way to
avoid the ambiguous term “increasing,” which sometimes means strictly increasing
and, at other times, nondecreasing.
If (an ) is a nondecreasing sequence and if A = {an | n ∈ Z>0 }, then clearly
a1 = inf A and (an ) is bounded if and only if A is bounded above.
The following example shows that a strictly increasing sequence may be bounded
or unbounded. The terms in part 1 of this example are the partial sums of the
exponential series while those in part 2 are the partial sums of the harmonic series,
cf. Example 5.2–4.
2.8 Examples: 1. If
1 1 1
an = 1 + 1 + + + ··· +
2! 3! n!
n−1
then an < an+1 for all n and, since n! > 2 if n ≥ 2,
1 1 1 1
an < 1 + 1 + + + + · · · + n−1
2 4 8 2
1−n
=3−2 <3
by the familiar formula for the sum of a geometric series. The sequence (an ) is
therefore strictly increasing and bounded.
2. If
1 1 1
an = 1 + + + · · · +
2 3 n
then an+1 > an for all n and
1 1 1 1 1 1 1
a2n = 1 + + + + + ··· + + ··· + + ··· + n
2 3 4 5 8 2n−1 + 1 2
1 1 1 1 1 1 1 1 1
>1+ + + + + + + +··· + n
+ ··· + n
2 4 4 8 8 8 8 2 2
| {z } | {z } | {z }
2 terms 4 terms 2n−1 textrmterms
1 1 n
>1+ + ··· + = 1 + .
2 2 2
28 2 Sequences of real numbers
2.2 Subsequences
A subsequence of a sequence is, roughly speaking, a sequence that is obtained by
taking only certain terms from the given sequence in the given order. Thus
2, 4, 6, 8, . . .
is a subsequence of 1, 2, 3, . . . whereas
2, 1, 8, 7, 3, 2, 31, . . .
2.11 Example: Suppose that (an ) is a sequence that is bounded above and put t =
sup{an | n ∈ Z>0 }. We consider two cases.
1. If an = t for some n, then t may or may not be the limit of a convergent subse-
quence whereas, if an < t for all n, then t is the limit of a convergent subsequence.
To see this, first notice that, if (an ) is the sequence 1, 0, 0, . . ., then t = 1 and no
subsequence of (an ) converges to 1.
2.2 Subsequences 29
2. Now suppose that an < t for all n. We will show that, in this case, there is a
subsequence (ank ) that converges to t.
First we make an observation. For any number > 0 and any positive integer m,
consider the number
0 = min{/m, t − a1 , t − a2 , . . . , t − am }.
Then 0 > 0 and so, by Theorem 1.12, there will be an integer n such that
t − 0m < an < t, and n > m by the definition of 0 . We shall use this observation
repeatedly to construct the subsequence (nk ).
Let n1 = 1 and let 1 = t − a1 . Then 1 > 0 and there is an n2 such that
t − 1 < an2 < t and n2 > n1 . Now, if
then there is an n3 such that t − 2 < an3 < t and n3 > n2 . Next, if
then there is an n4 such that t − 3 < an3 < t and n4 > n3 . Continuing in this
manner will give a subsequence (ank ) that converges to t. •
The following lemma (whose proof is easy and will be omitted) asserts that every
subsequence of a convergent sequence converges to the same limit. What turns out to
be much more interesting is that, as several of the above examples show, a subsequence
of a divergent sequence may well converge.
2.12 Lemma: If (ank ) is a subsequence of a sequence (an ) and if limn→∞ an = L,
then limk→∞ ank = L.
The next theorem turns out to be crucial for the subsequent development of the
theory and is even equivalent to the least upper bound property.
2.13 Theorem: (Bolzano–Weierstrass theorem) Every bounded sequence has a
convergent subsequence.
Proof: Suppose that (an ) is a bounded sequence and let m and M be numbers sat-
isfying m ≤ an ≤ M for all n. Put c = (m + M )/2, the midpoint of [m, M ]. If the
inequality m ≤ an ≤ c holds for infinitely many n, let m1 = m and M1 = c. On the
other hand, if the inequality m ≤ an ≤ c holds for only finitely many n, then the
inequality c ≤ an ≤ M must hold for infinitely many n, and in this case put m1 = c
and M1 = M . So, in either case [m1 , M1 ] ⊆ [m, M ], M1 − m1 = 2−1 (M − m), and
the inequality m1 ≤ an ≤ M1 holds for infinitely many n. Now let n1 be any integer
for which m1 ≤ an1 ≤ M1 .
The argument just described can be repeated. Let c1 be the midpoint of [m1 , M1 ].
Since the inequality m1 ≤ an ≤ M1 holds for infinitely many n, it follows that
m1 ≤ an ≤ c1 for infinitely many n or c1 ≤ an ≤ M1 for infinitely many n or both.
30 2 Sequences of real numbers
2.14 Example: Here are six sequences and, for each one, the corresponding set of all
limits of convergent subsequences:
Sequence Set of limits of all convergent subsequences
1, 12 , 13 , . . . {0}
1, −1, 1, −1, . . . {−1, 1}
1, 2, 3, . . . ∅
1, 23 , 13 , 54 , 15 , 76 , . . . {0, 1}
0, 1, 21 , 41 , 34 , 18 , . . . , 78 , 16
1
,... [0, 1]
(an ) is an enumeration of Q R
Let us characterize the convergent subsequences of a sequence.
2.15 Lemma: Suppose that (an ) is a sequence and let T denote the set of all limits
of convergent subsequences.
(i) Suppose that T 6= ∅.
(a) If sup T < ∞, then sup T ∈ T and, if inf T > −∞, then inf T ∈ T .
(b) If sup T = ∞, there is a subsequence of (an ) diverging to ∞ and, if inf T =
−∞, there is a subsequence of (an ) diverging to −∞.
(ii) If T = ∅ then no subsequence of (an ) converges and there is a subsequence
diverging to ∞ or to −∞ (it is possible that some subsequences of (an ) diverge
to ∞ and others diverge to −∞).
Proof: (i) Suppose first that T is not empty and that sup T < ∞. Let t = sup T and
let > 0. Theorem 1.12 implies that T ∩ (t − , t] 6= ∅ and hence there is a convergent
subsequence of (an ) whose limit lies in the interval (t − , t]. Now if (ank ) is such a
subsequence then t − < ank < t + for infinitely many values of k and, therefore,
an must lie in the interval (t − , t + ) for infinitely many values of n.
Since this is true for each > 0 and it follows that, for each positive integer
k, the inequality t − 1/k < an < t + 1/k must hold for infinitely many values of
n. We can use this to build a subsequence whose limit is t as follows. Let n1 be
any positive integer satisfying t − 1 < an1 < t + 1, let n2 be any integer satisfying
n2 > n1 and t − 1/2 < an2 < t + 1/2, let n3 be any integer satisfying n3 > n2
and t − 1/3 < an3 < t + 1/3, etc. This construction gives a subsequence (ank ) that
converges to t and therefore t ∈ T .
This proves the assertion in (i a) dealing with the case sup T < ∞, and the proof
of the one dealing with the case inf T > −∞ is similar and will be omitted.
Now suppose that T is not empty and that sup T = ∞. Then, for any positive
integer p, there is a convergent subsequence of (an ) whose limit lies in the interval
(p, ∞); if q is any other integer, there will be an integer n > q such that an > p. This
means that, given any two integers p and q, there is an integer n such that n > q and
an > p. We can use this observation to build a subsequence of (an ) that diverges to
∞ as follows. Let n1 be any positive integer satisfying 1 < an1 , let n2 be any integer
32 2 Sequences of real numbers
satisfying n2 > n1 and 2 < an2 , let n3 be any integer satisfying n3 > n2 and 3 < an3 ,
etc.
This proves the assertion in (i b) dealing with the case sup T = ∞, and the proof
of the one dealing with the case inf T = −∞ is similar and will be omitted.
(ii) Now suppose that T = ∅. The definition of T then implies that no subsequence
of (an ) can converge to a real number. Consider a positive integer k and the inequality
−k < an < k. If this inequality holds for infinitely many n, then (an ) would have
a bounded subsequence and this subsequence would, by the Bolzano–Weierstrass
theorem, have a convergent subsequence. But, since a subsequence of a subsequence
is a subsequence, this is impossible by the assumption that T = ∅. This shows that,
for each integer k, the inequality −k < an < k holds for only finitely many integers n.
So, for each integer k, the inequality |an | ≥ k must hold for infinitely many integers
n, and hence at least one of the inequalities an ≥ k and an ≤ −k must hold for
infinitely many values of n. We can use this observation to build a subsequence of
(an ) that diverges to ∞ or to −∞ as follows.
First suppose that, for each positive integer k, the inequality an ≥ k holds for
infinitely many values of n. Let n1 be any integer satisfying an1 ≥ 1, let n2 be any
integer satisfying n2 > n1 and an2 ≥ 2, let n3 be any integer satisfying n3 > n2 and
an3 ≥ 3, etc. Then (ank ) is clearly a subsequence diverging to ∞.
Now suppose that there is a positive integer k such that the inequality an ≥ k holds
for only finitely many values of n. Then, for each positive integer k, the inequality
an ≤ −k must hold for infinitely many values of n, and a similar argument will lead
to a subsequence that diverges to −∞.
Consider a sequence (an ) and let T be as in the preceding lemma. We want to
associate with this sequence two elements of the set R ∪ {−∞, ∞} to be denoted
by lim inf n→∞ an and lim supn→∞ an with the following desiderata in mind. If the se-
quence converges to a real number, say L, then both lim inf n→∞ an and lim supn→∞ an
are to be equal to L. More generally, lim supn→∞ an is to be the largest element of
the set R ∪ {−∞, ∞} to which a subsequence of (an ) converges or diverges, and
lim inf n→∞ an is to be the smallest element to which a subsequence converges or di-
verges (where, of course, we regard −∞ < x < ∞ for all x ∈ R). The point of the
above lemma is that this really does make sense. For example, if, say, T 6= ∅ and
sup T < ∞, then sup T belongs to T and so is the limit of a convergent subsequence.
In this example, then, lim supn→∞ an will be ∞ if a subsequence of (an ) diverges to
∞ and will be sup T otherwise. As a second example, if T 6= ∅ and sup T = ∞, then
there is a subsequence diverging to ∞ and lim supn→∞ an will be ∞. And as a final
example, if T = ∅, then no subsequence converges to a real number and there are
subsequences diverging to at least one of ±∞; so lim supn→∞ an will be ∞ if there
is a subsequence diverging to ∞ and will be −∞ otherwise. The next definition
makes this intuitive discussion of the definitions of lim supn→∞ an and lim inf n→∞ an
precise.
2.16 Definition: Suppose that (an ) is a sequence and let T be the set of all limits of
2.2 Subsequences 33
convergent subsequences.
(i) The limit superior of the sequence (an ) is the element lim supn→∞ an of the
set R ∪ {−∞, ∞} defined as follows:
(a) If (an ) is not bounded above, then lim supn→∞ an = ∞.
(b) If (an ) is bounded above and if T 6= ∅, then lim supn→∞ an is the largest
number that is the limit of a convergent subsequence of (an ).
(c) If (an ) is bounded above and if T = ∅, then lim supn→∞ an = −∞.
(ii) The limit inferior of the sequence (an ) is the element lim inf n→∞ an of the
set R ∪ {−∞, ∞} defined as follows:
(a) If (an ) is not bounded below, then lim inf n→∞ an = −∞.
(b) If (an ) is bounded below and if T 6= ∅, then lim inf n→∞ an is the smallest
number that is the limit of a convergent subsequences of (an ).
(c) If (an ) is bounded below and if T = ∅, then lim inf n→∞ an = ∞. •
The idea that the limit superior and limit inferior are in some sense replacements
for the limit is borne out by the following result.
2.17 Proposition: Suppose that (an ) is a sequence.
(i) If the sequence (an ) converges to a number or diverges to ±∞ then
lim inf an = lim an = lim sup an . (2.1 )
n→∞ n→∞ n→∞
(ii) If lim inf n→∞ an = limn→∞ an then either the sequence (an ) converges or di-
verges to ±∞; in either case (2.1) holds.
Proof: It is possible to prove this directly from the above definitions, but it is more
easily deduced from the properties of the limit superior and inferior contained in
Propositions 2.19 and 2.20 (cf. Exercise E2.44).
2.18 Example: This example is a continuation of Example 2.14.
Sequence lim inf n→∞ an lim supn→∞ an
1, 21 , 13 , . . . 0 0
1, −1, 1, −1, . . . -1 1
1, 2, 3, . . . ∞ ∞
1, 23 , 31 , 54 , 15 , 76 , . . . 0 1
0, 1, 21 , 41 , 34 , 18 , . . . , 78 , 16
1
,... 0 1
(an ) is an enumeration of Q −∞ ∞
The next two propositions contain important properties of the limit superior; the
corresponding properties of the limit inferior can be deduced from these propositions
and the easily verified formula
lim inf an = − lim sup(−an ) (2.2)
n→∞ n→∞
an > L0 + = L −
holds for only finitely many integers n, contradicting the assumption that L satis-
fies (i b). Now consider the case in which L0 > L and put = (L0 − L)/2. Then (just
as in the first case) the definition of L0 and part (i) implies that the inequality
an > L0 − = L +
holds for only infinitely many integers n, contradicting the assumption that L satis-
fies (i b).
This means that both of the assumptions L > L0 and L0 > L lead to contradictions,
and thus L = L0 .
Definition 2.16 together with Proposition 2.19 give us two different ways of char-
acterizing the limit superior and limit inferior of a sequence. We get a third charac-
terization by considering the limit of the sequence formed from (an ) by taking the
suprema and infima of its tails.
2.2 Subsequences 35
and
lim inf an = lim inf{ak | k ≥ n}.
n→∞ n→∞
ak > L + holds for only finitely many k. Now suppose that the inequality ak > L −
holds for only finitely many k. Then there would exist n ∈ Z>0 such that ak ≤ L− for
all k ≥ n. But then Mn ≤ L − , contradicting the fact that L = inf{Mn | n ∈ Z>0 }.
So L satisfies conditions (i a) and (i b) of Proposition 2.19 and so equals lim supn→∞ an .
The assertion dealing with the limit inferior can be proven in a similar manner or
can be deduced from what was just proven and equation (2.2).
for any two positive integers k and n. If is some positive number, then, since
L = limk→∞ ank , it follows that there will be an integer N1 such that |ank − L| < /2
whenever k ≥ N1 . Next, the assumption that (an ) is a Cauchy sequence implies
that there is an integer N2 such that |an − am | < /2 whenever m, n ≥ N2 . Now let
N = max{N1 , N2 } and recall that nk ≥ k for all k by the definition of a subsequence.
So, if n ≥ N , then
|an − L| ≤ |an − anN | + |anN − L| < ,
and this shows that L = limn→∞ an .
2.24 Example: We can use the preceding theorem to show that the sequence whose
nth term is
1 1 1 1
an = 1 − + − + · · · + (−1)n−1
2 3 4 n
is convergent. To do this, we must examine the expression |am − an | and, in doing
so, we may as well assume that m > n. The fact that n1 − n+1 1
> 0 for any positive
integer n implies that, if m − n is even, then
n 1 n+1 1 m+1 1
|am − an | = (−1) + (−1) + · · · + (−1)
n+1 n+2 m
1 1 1 1 1 1
= − + − + ··· + −
n+1 n+2 n+3 n+4 m−1 m
1 1 1 1 1 1
= − + − + ··· + −
n+1 n + 2 n + 3 n + 4 m−1 m
1 1 1 1 1 1
= − − − ··· − − −
n+1 n+2 n+3 m−2 m−1 m
1
<
n+1
whereas, if m − n is odd, then
1 1 1 1 1
|am − an | = − + ··· − +
n+1 n+2 m−2 m−1 m
1 1 1 1 1
= − − − ··· − −
n+1 n+2 n+3 m−1 m
1
< .
n+1
Now suppose that an > 0 is given and let N be an integer satisfying N + 1 > 1 .
1
Then |an − am | ≤ n+1 < whenever m ≥ n ≥ N and, therefore, (an ) is a Cauchy,
and hence a convergent, sequence.
The value of limn→∞ an will be determined in Example 5.13–1 to be ln 2. •
38 2 Sequences of real numbers
2.25 Example: Let (an ) be as in Example 2.24. Recall from this example that the
limit limn→∞ an exists and that, for any integer m ≥ 5, we have |a4 − am | < 1/5. So,
if L = limn→∞ an , then |a4 − L| ≤ 1/5, and since
1 1 1 7
a4 = 1 − + − = ,
2 3 4 12
7
we can conclude that | 12 − L| < 14 . This is consistent with the assertion made in the
previous example that L = ln 2. •
Exercises for Chapter 2 39
Exercises
E2.1 Determine limn→∞ an if:
n+1
(a) an =
3n
n3 + n
(b) an = 2
n + n
(c) an = 1 − 21 1 − 13 1 − 14 · · · 1 − n+1 1
1 2 n
(d) an = 2 + 2 + · · · 2
n n n
E2.2 Find limn→∞ an in the following cases:
n2
(a) an = , n ≥ 1.
n + 1000
1 1 1 1
(b) an = 1·2 + 2·3 + 3·4 + · · · + n(n+1) , n ≥ 1.
In each case, you must give a proof using the definition of limn→∞ an .
1
Hint: For (b), express k(k+1) as the difference of two fractions.
E2.3 Using the definition of a limit, find limn→∞ an , where
( 2
n +1
sin(n2 +1)
, if n < 21000 ,
an = sin(n 2 +1)
n2 +1
, if n ≥ 21000 .
n
E2.4 It is easy to see that limn→∞ n+1 = 1. For each number > 0 find an integer
N such that
n
n + 1 − 1 <
for all n ≥ N .
E2.5 Prove, using the definition of convergence, that
n−1
lim = 1.
n→∞ n + 1
√ √
E2.6 Define a sequence (an ) by the two conditions a1 = 2 and an+1 = 2 + an for
n ≥ 1. Show that this sequence converges and find its limit.
Hint: Show that the sequence is monotone and bounded.
n
E2.7 Assuming that limn→∞ 1 + n1 = e, compute:
1 n
(a) limn→∞ 1 + 3n ;
n 2
(b) limn→∞ 1 − n12 .
(a) limn→∞ an + bn = A + B
(b) limn→∞ an bn = AB
Hint: For part (b), see Exercise E2.20.
E2.22 Show that, if (an ) and (bn ) are two sequences such that an ≤ bn for every
n ∈ Z>0 and limn→∞ an = ∞, then limn→∞ bn = ∞.
E2.23 Show that, if (an ) is a sequence such that limn→∞ an = ∞, then limn→∞ a1n = 0.
E2.24 Suppose that limn→∞ an = a, with a 6= 0. Prove that limn→∞ 1/an = 1/a.
√
E2.25 (a) Show that xy ≤ (x + y)/2 for any two nonnegative numbers x and y.
(The left and right sides of this inequality are the geometric and arithmetic
means of x and y, respectively.)
(b) Let x1 and y1 be two numbers satisfying 0 < x1 < y1 and put xn+1 =
√
xn yn and yn+1 = (xn + yn ) /2 for n ≥ 1. Prove that the sequences (xn )
and (yn ) are convergent and that they have the same limit.
E2.26 Let (an ) be a sequence and define a sequence (bn ) by putting
a1 + a2 + · · · + an
bn =
n
for n ≥ 1.
(a) Show that, if limn→∞ an = 0, then limn→∞ bn = 0.
(b) Show that, if limn→∞ an = L, then limn→∞ bn = L, where L is a real
number. (This result is important in the study of probability theory and
Fourier series.)
(c) Give an example for which (an ) diverges but (bn ) converges.
E2.27 (a) Use induction to prove Bernoulli’s inequality: (1 + h)n > 1 + nh or all
integers n > 1 and all numbers h > 0.
(b) Use this inequality to prove that if r > 1 then limn→∞ rn = ∞.
(c) Prove that if |r| < 1 then limn→∞ rn = 0.
E2.28 In this exercise, we will prove the following statement, known as Fekete’s
Lemma:
Let a1 , a2 , a3 , . . . be a sequence of nonnegative real numbers with the
property that am+n ≤ am + an for all m, n. Then, the sequence ann
converges to inf{ ann | n ≥ 1}.
So, let (an ) be a sequence as in the hypothesis of the above statement. The
set { ann | n ≥ 1} is bounded below by 0, so its infimum exists. Set L =
inf{ ann | n ≥ 1}.
(a) For positive integers k, m, prove that akm ≤ mak .
(b) Show that, for any > 0, there exists N such that, for all n ≥ N ,
an
n
≤ L + .
Hint: Let > 0 be arbitrary. Pick a k such that akk < L + /2 (why
can this be done?). Now, any n can be expressed as n = km + r, with
42 2 Sequences of real numbers
0 ≤ r < k (why?). Part (a) shows that an ≤ mak + ar . Deduce from this
that there exists N such that, for all n > N , ann < akk + /2.
(c) Deduce from (b) that limn→∞ ann = L.
E2.29 Suppose that (ank ) is a subsequence of a monotone sequence (an ). Is it true
that, if the limit limk→∞ ank exists, then so does limn→∞ an and that these
limits are equal? Why?
E2.30 Suppose that f is a continuous function whose domain is the real line and that
(an ) is a convergent sequence. Show that limn→∞ f (an ) = f (limn→∞ an ).
E2.31 A sequence (an ) of real numbers is said to be contractive if there exists a
constant β with 0 < β < 1 such that
for all n ≥ 1. Show that any contractive sequence is a Cauchy sequence, and,
therefore, is convergent.
Hint: Obtain an estimate for |an − am | in terms of β and |a1 − a0 |.
E2.32 A function f : R → R is said to be a contraction on R if there exists a
constant β with 0 < β < 1 such that
E2.34 Suppose that (an ) is a sequence such that L = lim supn→∞ an is a real number.
Then we know that, for any number M > L, there are only finitely many
integers n for which an > M . Show by means of an example that it is possible
to have an > L for infinitely many n.
E2.35 Prove equation (2.2).
E2.36 Suppose that (an ) and (bn ) are two bounded sequences.
(a) Show that, if an ≤ bn for all n then lim supn→∞ an ≤ lim supn→∞ bn .
(b) Show that
lim sup (an + bn ) ≤ lim sup an + lim sup bn .
n→∞ n→∞ n→∞
(c) Give an example for which equality holds in (a) and an example for which
strict inequality holds.
E2.37 Can you find two distinct sequences (an ) and (bn ) such that each is a subse-
quence of the other? If yes, give an example; if no, give a proof.
E2.38 Find the limit superior, the limit inferior, and the set of limits of convergent
subsequences of the sequence 1, 2, 1, 4, 1, 6, 1, 8, . . ..
E2.39 For each of the sequences (an ) below, identify the set of all limits of convergent
subsequences. Also, determine lim inf n→∞ an and lim supn→∞ an .
n!
(a) an = n , for n ≥ 1
n
(−1)n
n
(b) an = (−1) 1 + , for n ≥ 1
n
(c) 1, 1, 2, 1, 2, 3, 1, 2, 3, 4, 1, 2, 3, 4, 5, 1, 2, . . . (The first term is 1, the next two
terms are 1, 2, the next three terms are 1, 2, 3, and so on.)
(d) 21 , 31 , 23 , 14 , 42 , 34 , 15 , 25 , 53 , 45 , . . . (The first term is 12 , the next two terms are
1 2
, , the next three terms are 14 , 42 , 34 , and so on.)
3 3
E2.40 For each of the sequences (an ) below, identify the set of all limits of convergent
subsequences. Also, determine lim inf n→∞ an and lim supn→∞ an .
1
(a) an = 2 − (−1)n , for all n ≥ 1
n
1 1 1
(b) an = 1 + + + · · · + , for all n ≥ 1
2 3 n
(c) 1, 1, 12 , 1, 12 , 13 , 1, 21 , 13 , 14 , 1, 12 , 13 , 14 , 15 , . . . (The first term is 1, the next two
terms are 1, 12 , the next three terms are 1, 21 , 13 , and so on.)
E2.41 Let
−n,
n ≡ 0 mod 3,
an = 1 + 1/n, n ≡ 1 mod 3,
−1 − 1/n, n ≡ 2 mod 3
for n ∈ Z>0 .
(a) Show that (an ) is bounded above, is not bounded below, and that the set
of limits of convergent subsequences is {−1, 1}.
44 2 Sequences of real numbers
(b) Prove that lim supn→∞ an = 1 and lim inf n→∞ an = −∞.
E2.42 (a) Suppose that (an ) and (bn ) are two bounded sequences. Show that
lim supn→∞ (an + bn ) ≤ lim supn→∞ an + lim supn→∞ bn .
(b) Give an example in which the inequality above is strict, and another in
which the inequality holds with equality.
E2.43 Prove that, if a sequence diverges to ∞, then it has no convergent subsequences.
E2.44 Prove that, if L is a real number and that (an ) is a sequence, then limn→∞ an =
L if and only if lim inf n→∞ an = L and lim supn→∞ an = L (cf. Proposi-
tion 2.17).
Chapter 3
Sequences in Rd
Sequences and their limits can be defined in Rd for d ≥ 2 almost exactly as they
were defined in R. (It will be convenient to use d for the dimension of the Euclidean
space because of the ubiquitous use of n as a subscript.) As is customary in linear
algebra, if x is a vector in Rd then xk denotes the kth-component of x for k = 1, . . . , d,
so that x = (x1 , . . . , xd ).
The norm of a vector x ∈ Rd is the number
q
kxk = x21 + · · · + x2d .
Notice that, if d = 1, then the norm is just the absolute value. The norm has the
following properties:
1. kxk ≥ 0 for all x ∈ Rd with equality if only if x = 0;
2. kaxk = |a|kxk for all a ∈ R and x ∈ Rd ;
3. kx + yk ≤ kxk + kyk for all x, y ∈ Rd .
The first two of these properties are obvious, and the third, called the triangle
inequality , is most easily verified by first proving the Cauchy–Schwartz inequality
|x · y| ≤ kxkkyk,
which is valid for all x, y ∈ Rd . Both this inequality and property 3 above should be
familiar from linear algebra.
If (xn ) is a sequence in Rd the kth-component of xn is denoted by xn,k and so xn
itself is (xn,1 , . . . , xn,d ).
3.1 Definition: A sequence (xn ) in Rd is said to have a limit L, or to converge to
L if, for any > 0, there exists an integer N such that
kxn − Lk <
hand, and it is necessary to change notation and replace the original sequence (xn )
by the subsequence (xnk ). Now the sequence (xn,1 ) converges, and we can apply the
Bolzano–Weierstrass theorem to the sequence (xn,2 ) to find a convergent subsequence,
say, (xnk ,2 ). The sequence (xnk ,1 ) still converges by Lemma 2.12, and the sequence
(xnk ) is a subsequence of the original sequence. Let us now change notation once
again, replacing the sequence (xn ) by the subsequence (xnk ); this gives a sequence
which is a subsequence of the original sequence and with the property that both of
the sequences (xn,1 ) and (xn,2 ) converge. We can now repeat the argument, applying
the Bolzano–Weierstrass theorem to the sequence (xn,3 ).
After carrying out this argument d times, we will get a sequence (xn ) which is a
subsequence of the original sequence, and such that the sequence (xn,i ) converges for
i ∈ {1, . . . , d}, and hence, by Theorem 3.2, such that limn→∞ xn exists.
The notion of Cauchy sequence in Rd can be made entirely analogously to the same
notion in R. Moreover, it is still true that a sequence in Rd converges if and only if
it is a Cauchy sequence (cf. Exercise E3.3). The fact that this sort of equivalence is
not true for infinite-dimensional R-vector spaces with norms is a launching point for
a great deal of interesting and useful analysis.
A contraction on Rd is a function f from Rd to Rd such that there is a number
K < 1 with the property that
for all u, v ∈ Rd . The important property of contractions is they have unique fixed
points, i.e., if f is a contraction on Rd , there is exactly one point x ∈ Rd with the
property that f (x) = x (cf. the following theorem). This fact (albeit in a more
general context) can be used to prove a number of important theorems, such as the
differentiability of the inverse of a one-to-one differentiable function or the existence
of solutions of ordinary differential equations.
3.5 Theorem: (Contraction mapping theorem) A contraction on Rd is contin-
uous and has a unique fixed point.
Proof: The proof will be left as an exercise (cf. Exercise E3.9).
48 3 Sequences in Rd
Exercises
E3.1 Prove the equivalence of the two possible definitions of boundedness in Rd .
E3.2 Find the limits, if they exist, of these sequences in R2 :
n 1
(a) (−1) , ;
n
1
(b) 1, ;
n
1 1 π
(c) (cos(nπ)), sin nπ + ;
n n 2
1 −n
(d) ,n .
n
E3.3 (a) Give a definition of a Cauchy sequence in Rd .
(b) Show that a sequence (xn ) in Rd is Cauchy if and only if (xn,k ) is a Cauchy
sequence of real numbers for k = 1, . . . , d.
(c) Prove that every convergent sequence in Rd is Cauchy.
(d) Prove that every Cauchy sequence in Rd converges.
E3.4 Suppose that (xn ) is a Cauchy sequence of points in Rd .
(a) Prove that (xn ) is bounded.
(b) Prove that, if (an ) is a sequence of real numbers such that limn→∞ an = 0
then ((1 + an )xn ) is a Cauchy sequence in Rd .
E3.5 (a) Prove that every subset of R that is not bounded above contains a se-
quence that diverges to ∞.
(b) Prove that every unbounded subset of Rd contains a sequence (xn ) with
the property that limn→∞ kxn k = ∞.
E3.6 Let S = {(x, y) ∈ R2 | xy > 1} and B = {k(x, y)k | (x, y) ∈ S}. Find inf B.
E3.7 For n ≥ 1, let an = (−1)n , and let
(
1, if n < 10100 ,
bn =
1/n otherwise.
E3.9 Suppose that f is a contraction on Rd and let K denote the associated constant
(so that 0 < K < 1). Prove Theorem 3.5 as follows. Let x1 be any point in
Rd and let x2 = f (x1 ), x3 = f (x2 ), . . ..
(a) Prove that f is continuous.
(b) Prove that f has at most one fixed point.
Hint: Suppose there were two.
(c) Obtain an estimate for kxm+1 − xm k in terms of K and kx2 − x1 k for any
integer m.
(d) Obtain an estimate for kxn − xm k in terms of K and kx2 − x1 k for any
two integers m, n.
(e) Show that (xn ) is a Cauchy sequence in Rd .
(f) Show that the vector x = limn→∞ xn is a fixed point of f .
E3.10 For each number α put fα (x) = αx for all x ∈ R.
(a) For which numbers α is the function fα a contraction on R?
(b) For those numbers α for which fα is a contraction, what is the fixed point
of fα ?
E3.11 For what intervals [0, r], r ≤ 1, is
f : [0, r] → [0, r]
x 7→ x2
contraction?
Chapter 4
Some topology in Rd
Topology is, roughly speaking, that part of mathematics that studies open and
closed sets and continuous functions, and the purpose of this section is to give an
introduction to these ideas and some related ones in the context of Euclidean space.
In particular, we shall prove that a continuous function defined on a closed and
bounded set has a bounded range and attains both a maximum and a minimum
value. This result should be familiar from first-year calculus, where it is often stated
but rarely proven.
These are the open ball of radius r and center x, the closed ball of radius r and
center x, and the punctured ball of radius r and center x, respectively.
If d = 1, then Br (x) and B r (x) are just the familiar open and closed intervals
(x − r, x + r) and [x − r, x + r], respectively, and Bbr (x) = (x − r, x) ∪ (x, x + r).
If d = 2, then Br (x) and B r (x) are disks in R2 with center at x and radius r,
with the former not including its rim and the latter including its rim. The set B br (x)
is the “punctured” disk with center x and radius r that includes neither its center
nor its rim.
If d = 3, the three sets Br (x), B r (x), and B
br (x) are balls (in the familiar sense)
with center x and radius r.
Balls will be essential to phrase many of the notions we will introduce to charac-
terize subsets of Rd . We begin this process now.
4.1 Open and closed sets 51
A ⊆ B if and only if B { ⊆ A{ .
4.2 Definition: A subset A ⊆ Rd is said to be open if, for each point x ∈ A, there
is an r > 0 such that Br (x) ⊆ A, and to be closed if A{ is open. •
Open and closed sets are important building blocks for the structure of Rd . It is
obvious that Rd is an open set and hence that the empty set is a closed set. The empty
set is also an open subset of Rd , and to see this it is best to argue by contradiction.
Indeed, if a subset A of Rd is not open then, by definition, it must contain a point x
with the property that Br (x) 6⊆ A for all numbers r > 0. But the empty set does not
have this property since it does not contain any points. The fact that the empty set
is open now implies that Rd is closed. So the empty set and Rd are both open and
closed subsets of Rd , and it is true (but not so easy to prove) that these are the only
subsets of Rd that are both open and closed.
Points in open sets have a useful property.
4.3 Definition: A point x ∈ Rd is said to be an interior point of a subset A ⊆ Rd
if there is an r > 0 such that Br (x) ⊆ A. By int(A) we denote the set of interior
points of A, which is the interior of A. •
Sometimes the notation A◦ is used for the interior.
4.4 Proposition: A set A is open if and only if every point of A is an interior point
of A. Equivalently, A is open if and only if int(A) = A.
Proof: This follows immediately from Definitions 4.2 and 4.3.
Informally speaking, we can say that a set is finite if it is empty, or contains 1
point, or contains 2 points, or . . . and is infinite if it is not finite. More formally, a
set is usually said to be finite if every one-to-one function from the set into itself is
onto and to be infinite if there is a one-to-one function from the set into itself that is
not onto. For example, the function that multiplies each integer by 2 is a one-to-one
function from Z>0 into itself that is not onto, and so Z>0 is infinite. It is important
to not confuse the adjectives “finite” and “bounded”: Every finite set is bounded but
a bounded set need not be finite.
Let us consider some examples of open and closed sets.
4.5 Examples: 1. Every nonempty finite subset of Rd is closed and not open. Indeed,
suppose that A is a nonempty finite subset of Rd . To show that A is closed it is
necessary to show that A{ is open, so consider a point x ∈ A{ . Then the number
r = min{kx − yk | y ∈ A}
52 4 Some topology in Rd
that Br (x) is an open set, consider a point y ∈ Br (x) and put s = r − ky − xk.
Then s > 0 by the definition of Br (x) and we can show that Bs (y) ⊆ Br (x) as
follows. If z ∈ Bs (y), then
kz − xk ≤ kz − yk + ky − xk < s + ky − xk = r,
4.6 Proposition: (i) The union of any number of open subsets of Rd is open.
(ii) The intersection of any finite number of open subsets of Rd is open.
(iii) The union of any finite number of closed subsets of Rd is closed.
(iv) The intersection of any number of closed subsets of Rd is closed.
Proof: (i) Suppose that a set A is the union of open sets and let x be a point in A.
Then x must belong to one of the open sets, say B, whose union is A. But then
Br (x) ⊆ B for some r > 0 since B is open, and thus Br (x) ⊆ A since B ⊆ A. This
shows that A is open.
(ii) Suppose that A = m
T
i=1 Bi , where each of the sets B1 , . . . , Bm is open, and
let x be a point in A. Then, for each i ∈ {1, . . . , m}, there is an ri > 0 such that
Bri (x) ⊆ Bi . If r = min{r1 , . . . , rm }, then r > 0 and
The following example will show that one cannot omit the adjective “finite” in
parts (ii) and (iii) of the above proposition.
k−1
4.7 Example: Consider a point x ∈ Rd . For any k ∈ Z>0 it is clear that k
< 1 and
hence that B (k−1)/k (x) ⊆ B1 (x), and, therefore,
∞
[
B (k−1)/k (x) ⊆ B1 (x).
k=1
On the other hand, if y ∈ B1 (x), then there is a k ∈ Z>0 such that kx−yk ≤ (k−1)/k.
Thus y ∈ B (k−1)/k (x) and this gives
∞
[
B (k−1)/k (x) = B1 (x),
k=1
54 4 Some topology in Rd
and so the union of an infinite number of closed sets may well be open and not closed.
A similar analysis will show that
∞
\
B(k+1)/k (x) = B 1 (x),
k=1
and so the intersection of an infinite number of open sets may be closed and not
open. •
then r > 0 since A is finite, and A ∩ Br (x) is either ∅ or {x}. In either case,
A∩B br (x) = ∅, and thus x is not a cluster point of A.
2. It is easy to see that, for each x ∈ Rd and any r > 0, each of the sets der(Br (x)),
der(B r (x)), and der(B br (x)) is equal to B r (x). •
4.2 Cluster points, limit points, and boundary points 55
We are now ready to prove one of the fundamental theorems about continuous
functions, one that has to do with functions with a compact domain.
4.24 Theorem: Suppose that A is a compact subset of Rd and that f is a continuous
Rm -valued function with domain A.
(i) The image image(f ) = {f (x) | x ∈ A} of f is a compact subset of Rm .
(ii) If m = 1, then f attains a minimum value and a maximum value on
A, i.e., there are points u and v in A such that
for all x ∈ A.
Proof: (i) To show that image(f ) is compact it is only necessary to show that it
satisfies condition (ii) of Theorem 4.21. Thus suppose that (y n ) is a sequence in
image(f ). Then the definition of image(f ) implies that there must be a sequence
(xn ) in A such that f (xn ) = y n for all n ∈ Z>0 . Theorem 4.21 now implies that
there is a subsequence (xnk ) of (xn ) that converges to a point, say x, in A. But then
Theorem 4.23 implies that limk→∞ y nk = limk→∞ f (xnk ) = f (x), and so (y n ) does
indeed have a subsequence that converges to a point in image(f ).
(ii) Now suppose that m = 1. Since image(f ) is compact by (i), both inf image(f )
and sup image(f ) belong to image(f ). If u and v are points in A such that f (u) =
inf image(f ) and f (v) = sup image(f ), then it is clear that f attains a minimum
value at u and a maximum value at v.
The proof of Theorem 4.24 used the Bolzano–Weierstrass theorem and thus de-
pended on the completeness of R, i.e., the fact that all Cauchy sequences in R con-
verge. The intermediate value theorem is another theorem that is familiar from
first-year calculus and that depends on the completeness of R (cf. Exercise E4.29).
4.26 Example: The set {(1/n, 1) | n ∈ Z>0 } of open intervals is an open covering of
the interval (0, 1) that has no finite subcover. •
The point of introducing open coverings and finite subcovers is that they can be
used to give an third condition equivalent to the two conditions in Theorem 4.21, and
to give another proof of Theorem 4.24. Why is it desirable to do this since we already
know these theorems are true? The reason is that, in more advanced analysis, it is
sometimes necessary, when proving properties of closed and bounded sets, to think of
them in terms of open coverings and finite subcovers. There are also many situations
in analysis and topology where the three conditions are no longer equivalent and the
correct (or at least the most useful) one is the one involving open coverings and finite
subcovers.
4.27 Theorem: (Heine–Borel theorem) For a subset A ⊆ Rd , the following three
conditions are equivalent:
(i) A is closed and bounded;
(ii) every sequence in A has a subsequence converging to a point in A;
(iii) every open covering of A has a finite subcover.
Proof: We already know that (i) and (ii) are equivalent.
(i) =⇒ (iii) For this, the more difficult part of the proof, we first prove a couple
of lemmata.
1 Lemma: If K1 ⊆ Rm is compact and if K2 ⊆ Rn is compact then K1 ×K2 ⊆ Rm+n
is compact.
Proof: Let us denote points in Rm+n by (x, y) ∈ Rm × Rn . For x ∈ Rm , denote
K2,x = {(x, y) | y ∈ K2 }.
Ux = {U ∈ U | U ∩ K2,x 6= ∅}.
For U ∈ Ux , define
VU = {y ∈ U | (x, y) ∈ K2,x }.
We claim that VU is open. Indeed, let y ∈ VU so that (x, y) ∈ U . Since U is open,
there exists r > 0 such that Br (x, y) ⊆ U . Therefore, Br (y) ⊆ VU , and so VU is
open as claimed. Therefore, (VU )U ∈ Ux is an open cover of K2 . Thus there exists
Ux,1 , . . . , Ux,kx ∈ Ux such that K2 ⊆ ∪kj=1
x
VUx,j .
Now, for U ∈ U, denote
We claim that WU is open. To see this, let x ∈ WU and let y ∈ Rn be such that
(x, y) ∈ U . Since U is open, there exists r > 0 such that Br (x, y) ⊆ U . Therefore,
Br (x) ⊆ WU , giving WU as open, as desired. Now define Wx = ∩kj=1 x
WUx,j and note
4.5 Compact sets and open coverings 61
that, by Proposition 4.6, it follows that Wx is open. Thus (Wx )x∈K1 is an open cover
for K1 . By compactness of K1 , there exists x1 , . . . , xm ∈ K1 such that K1 = ∪m
l=1 Wxl .
Therefore,
kx
K1 × K2 = ∪x∈K1 K2,x = ∪x∈K1 ∪kj=1
x
Ux,j = ∪m
l=1 ∪j=1 Uxl ,j ,
l
and so U is an open covering of A. Sn But then{ (iii) implies that there are positive
numbers r1 , . . . , rn such that A ⊆ j=1 B rj (x) . Now, if r = min{r1 , . . . , rn }, then
r > 0 and n n
\ [
{
B r (x) ⊆ B rj (x) =⇒ B r (x) = B rj (x){
j=1 j=1
and, therefore, A ⊆ B r (x){ . But this means that B r (x) ⊆ A{ and so Br (x) ⊆ A{ .
Since x was an arbitrary point in A{ , it follows that A{ is open.
The more difficult part of Theorem 4.27 is the implication (i) =⇒ (iii). One part
of the proof we pull out as the following example, since this may help us to understand
this implication.
4.28 Example: We shall show that any open covering of [0, 1] has a finite subcover.
Indeed, suppose that U is an open covering of [0, 1] and consider the set
The first step is to show that C is not empty. There is a set U in U with 0 ∈ U and
hence with (−r, r) ⊆ U for some r > 0. But then [0, r/2] ⊆ U and so there is finite
subcover (consisting of just one set, in fact) of [0, r/2] and thus r/2 ∈ C.
To prove that U contains a finite subcover of [0, 1], it is enough to show that
1 ∈ C and we will do this by putting s = sup C and showing that s = 1 and that
s ∈ C.
Since s ∈ [0, 1] there will be a set U ∈ U with s ∈ U and hence with (s−, s+) ⊆
U for some > 0. Now, by Theorem 1.12, the interval (s − , s] must contain a point
in C, sayS x. Then there are a finite number of sets U1 , . . . , Un in U such that
[0, x] ⊆ ni=1 Ui . But then the sets U, U1 , . . . , Un are a finite subcover of [0, s] and
so s ∈ C. Moreover, if s < 1 and if r = min{1 − s, /2}, then it is even true that
s + r ≤ 1 and these same n + 1 sets cover [0, s + r], and hence s + r ∈ C. But this
contradicts the choice of s and therefore s = 1. •
Let us give a couple of examples of the use of the finite subcover characterisation
of compactness. For the first result, we shall make use of the following definition.
4.29 Definition: Let A ⊆ Rd and let f be a real-valued function on A. The function
f is uniformly continuous if, for each > 0, there exists δ > 0 such that |f (x) −
f (y)| < if x, y ∈ A satisfy kx − yk < δ. •
Note carefully the difference between “continuous” and “uniformly continuous.”
The next theorem shows that these notions align for compact domains.
4.5 Compact sets and open coverings 63
{Bδx /2 (x) | x ∈ K}
Now let x, y ∈ K satisfy kx − yk < δ. We must have x ∈ Bδxj /2 (xj ) for some
j ∈ {1, . . . , k}. This, in turn, gives
{Bδx (x) | x ∈ A}
Exercises
E4.1 Determine the interior points, the cluster points, the limit points, the boundary
points, and the isolated points of each of the following sets (note that the sets
in (a)–(c) are subsets of R2 whereas the set in (d) is a subset of R):
(a) {(q, r) | q and r are rational and 0 ≤ q, r < 1}
(b) {(x, y) | x ∈ R and − 1 < y < 1}
n1 1 o
(c) , n, m ∈ Z>0
n m
∞
[ 1 1
(d) ,
n=1
n + 1 n
E4.2 Determine the interior points, the cluster points, the limit points, the isolated
points, and the boundary points of each of the following sets. Also, classify
each of the sets as open, closed, neither or both.
1
(a) n ∈ Z>0 , n ≥ 1
n
(b) {x ∈ Rd | kxk = 1}
E4.3 Determine the interior points, the boundary points, the cluster points, the
limit points, and the isolated points of each of the following subsets of R2 .
Also, classify each of the sets as open, closed, neither or both.
(a) {(m, n) | m, n ∈ Z}
(b) {(x, y) ∈ R2 | 0 < x ≤ 1}
E4.4 Determine the interior points, the cluster points, the limit points, the boundary
points, and the isolated points for each of the following sets. Also, state
whether the sets are open or closed or neither. (Note that the sets in (a)–
(e) are subsets of R.)
(a) (1, 2)
(b) [2, 3]
(c) (1, 4]
\∞
(d) [−1, 1/n)
n=1
(e) (0, 1) ∩ Q
(f) Rd
(g) A hyperplane in Rd .
(h) {(x, y) ∈ R2 | 0 < x ≤ 1}
(i) {x ∈ Rd | kxk = 1}
E4.5 Let S = {(x, y) ∈ R2 | |x| ≤ 1 and |y| < 1}.
(a) Show that S is not closed by finding a sequence in S that converges to a
point not in S.
Exercises for Chapter 4 65
S = {(x, y) ∈ R2 | y > x2 }.
{x ∈ S | f (x) ≤ a}
is a closed subset of Rd .
E4.27 Let f : Rd → Rm be a function. For any B ⊆ Rm , the preimage under f , of B
is defined to be the set
Show that f is a continuous function if and only if, for each open set U ⊆ Rm ,
the preimage f −1 (U ) is an open set in Rd .
E4.28 Let S be a nonempty subset of R.
68 4 Some topology in Rd
(a) Give an example to show that, if S is not closed, then there is a continuous
real-valued function with domain S that does not attain a maximum value.
(b) Give an example to show that, if S is not bounded, then there is a contin-
uous real-valued function with domain S that does not attain a maximum
value.
(c) Give an example of a continuous bounded function with domain R that
attains neither a maximum or minimum value.
E4.29 (Intermediate value theorem) Suppose that f is a continuous real-valued
function whose domain includes a closed interval [a, b]. Show that, if f (a) < 0
and f (b) > 0, then there is a point c ∈ [a, b] such that f (c) = 0.
Hint: Consider the set {x ∈ [a, b] | f (x) < 0}.
E4.30 Show that every cubic polynomial has a root.
Hint: Use Exercise E4.29.
E4.31 Show that every continuous function from the unit interval to itself has a fixed
point.
Hint: Apply the intermediate value theorem to the function x 7→ f (x) − x.
E4.32 Suppose that f is a continuous function whose domain is [1, 2] and that satisfies
0 ≤ f (x) ≤ 3 for all x ∈ [1, 2]. Show that f has a fixed point, i.e., that there
is a point x ∈ [1, 2] satisfying f (x) = x.
E4.33 Let f be a function from Rm to Rn and recall that we can write f (x) =
(f1 (x), . . . , fn (x)) for x ∈ Rm , where f1 , . . . , fn are real-valued functions on
Rm . Show that f is continuous if and only if fi is continuous for i ∈ {1, . . . , n}.
Chapter 5
Infinite series
for a sequence (ak ). The natural way to do this would be to consider the limit (as n
tends to ∞) of the sum of the first n terms, and this is, in fact, what will be done.
5.1 Definition: Consider a sequence (ak ) of real numbers defined for k ≥ 1. The
formal expression
X∞
ak (5.1)
k=1
is called the infinite series corresponding to the sequence (ak ). For n ∈ Z>0 the
sum n
X
Sn = ak
k=1
is called the n partial sum of (5.1) and the sequence (Sn ) is called the sequence
of partial sums corresponding to (5.1).
If the sequence (Sn )P
converges then the series (5.1) is said to converge
P∞and one
∞
says that the value of k=1 ak is the limit limn→∞ Sn , and one writes k=1 ak =
limn→∞ Sn . On the other hand, if the sequence (Sn ) diverges, then the series
P (5.1) is
said to diverge or to not exist and no meaning is assigned to the symbol ∞ k=1 ak . •
The point of this definition is that, for some sequences (ak ), the infinite series (5.1)
is defined to be a real number, while for other sequences (5.1) is not assigned a
meaning. Namely, if the limit limn→∞ Sn exists, then (5.1) is defined to be the value
70 5 Infinite series
of this limit, whereas if this limit does not exist then the expression (5.1) is not
assigned a value. In the latter case, then, we cannot “talk” about and manipulate
the “sum” as though it were a number—it is merely a “formal expression.1 ” Students
sometimes ignore the importance of the sequence (Sn ) of partial sums and thereby
miss the essential feature of infinite series. Partial sums are finite sums, so they can
be dealt with and manipulated according to all the familiar rules for addition; the
corresponding statement for infinite series is not true. All statements about infinite
series must be understood as abbreviations for statements about limits of partial
sums.
Notice that, if two sequences are identical except for a finite number of terms, then
the two corresponding infinite series will either both converge or both diverge. Indeed,
for n sufficiently large the two nth partial sums will have a constant difference, and
hence will either both converge or both diverge. A particularly important case of this
arises when we “neglect the first few terms of aP series:” If (aP
k ) is a sequence defined
for k ∈ Z>0 and if N ∈ Z>0 , then the two series k=1 ak and ∞
∞
k=1 ak will either
P both
converge or both diverge. To see this,Plet Sn denote the nth partial sum nk=1 ak ,
and let Tn denote the nth partial sum nk=N ak for n ≥ N . Then Sn = SN −1 + Tn for
n ≥ N and, since SN −1 is fixed, limn→∞ Sn exists if and only if limn→∞ Tn exists.
It seems somewhat ironic that, while it is generally difficult to calculate the value
of a convergent infinite series, it is often relatively easy to decide whether an infinite
series converges or diverges. What this really means, when rephrased in terms of
partial sums, is that, while it is generally difficult to calculate the limit of a convergent
sequence, it is often relatively easy to decide whether a sequence converges or diverges.
(Of course, if a series is known to converge one can, in principle, calculate its sum as
accurately as desired by simply calculating its nth partial sum for large values of n.
But this is not the same as finding the value of the infinite series.) Before going on to
describe these so-called “convergence tests,” it will be useful to give a few examples.
5.2 Examples: 1. It is easy to see that both of the infinite series
∞
X ∞
X
1 and (−1)k
k=1 k=1
diverge. Indeed, for the first series the nth partial sum is n and so the sequence
of partial sums diverges to ∞. And for the second the sequence of partial sums
is −1, 0 − 1, 0, . . . and hence is divergent.
2. Let r be a real number and consider the infinite series ∞ k
P
k=0 r . This series is
called a geometric series; let (Sn ) be its sequence of partial sums. It is obvious
that, if r = 1, then Sn = n + 1 and hence the geometric series diverges in this
case. Now suppose that r 6= 1. Then
(1 − r)(1 + r + · · · + rn ) = 1 − rn+1 ,
1
Something which can be made precise, but which will definitely not be made precise here.
5.1 Definitions and examples 71
and hence
1 − rn+1
Sn = .
1−r
This makes it clear that, if |r| < 1, the geometric series converges with
∞
X 1
rk = ,
k=0
1−r
Proof: The two parts of this theorem are, of course, logically equivalent and so it is
only
P∞ necessary to prove (i). Let (Sn ) be the sequence of partial sums of the series
k=1 ak and let be a given positive number. Then limn→∞ Sn exists and, if L =
limn→∞ Sn , there is an integer N such that |L − Sn | < 2 whenever n ≥ N . But then,
for any integer n ≥ N , we have
|an+1 | = |Sn+1 − Sn |
≤ |Sn+1 − L| + |L − Sn |
< + = .
2 2
This shows that limn→∞ an = 0.
Notice that Example 5.2–4 shows that the converse of this theorem is false. It is
very important to realize that this theorem gives a necessary, but not a sufficient,
condition for convergence of a series. This theorem can sometimes be used to establish
the divergence of a series but cannot be used to establish convergence.
P∞One way to understand the preceding example and theorem is that, if a series
k=1 ak is to converge, then it is necessary that ak → 0 quite quickly as k → ∞, and
that the terms of the harmonic series simply approach zero too slowly.
Proof: (i) Let Sn = nk=1 ak and Tn = nk=1 bk for n ∈ Z>0 . Then the assumption
P P
that P0 ≤ ak ≤ bk for all k implies that 0 ≤ Sn ≤ Tn for all n, and the assumption
that ∞ k=1 bk converges impliesPthat (Tn ) is a bounded sequence. But then (Sn ) too
is a bounded sequence, and so ∞ k=1 ak converges by Theorem 5.4.
(ii) This assertion is logically equivalent to (i).
(iii) The hypotheses implies that, if L = limk→∞ ak /bk , then there exists an integer
N such that ak /bk ≤ L + 1 for all k ≥ N , and hence such that ak ≤ (L + 1)bk for
all k ≥ N . The desired conclusion now follows from (i), together with our earlier
observation that we may “neglect the first few terms” in showing convergence or
divergence (cf. page 70).
(iv) In this case, let 0 < L = limk→∞ ak /bk . Let > 0 be such that L − > 0.
Then there exists N ∈ Z>0 such P that L − < ak /bk for all k ≥ N . Thus, if (Sn ) are
the partial sums for the sum ∞ k=1 an , ak > (L − )bk for k ≥ N and so, for n > N ,
n
X n
X
Sn − SN = an > (L − ) bn .
k=N +1 k=N +1
Therefore n
X
lim Sn > lim (L − ) bn + SN = ∞,
n→∞ n→∞
k=N +1
as desired.
In applying parts (i) and (ii) of this theorem, it is sufficient that there be an n0
such that the inequalities 0 ≤ ak ≤ bk hold for all k ≥ N (and not for all k as in the
statement of the theorem). This follows from our earlier remark (cf. page 70) about
“neglecting the first few terms” in proving convergence or divergence.
5.6 Examples: 1. Since k 1/2 ≤ k for k ≥ 1, it follows that 1/k 1/2 ≥ 1/k for
P∞k ≥ 11.
The comparison test and Example 5.2–4 then imply that the series k=1 k1/2
diverges.
2. Since
1 1
>
2n + 1 2(n + 1)
for n ∈ Z>0 , the comparison test and Example 5.2–4 imply that the series
P ∞ 1
n=1 2n+1 diverges. Notice that this series consists of odd terms of the harmonic
sequence.
3. Since
1 2
2
≤
n n(n + 1)
for n ∈ Z>0 , the comparison test and Example 5.2–3 imply that the series ∞ 1
P
n=1 n2
converges.
74 5 Infinite series
P∞ 3
4. Does the series n=1 2n2 −n1/2 converge? Since the terms are positive and
. 1 3
3
lim = ,
n→∞ 2n2 − n1/2 n2 2
the series converges by 3 and the comparison test. Note that it would be more
awkward to use part (i) of the comparison test for this particular example. •
5.7 Theorem: (Integral test for positive series) Suppose that (ak ) is a sequence
and that f is a function which is continuous,2 nonnegative-valued, and nonincreasing
half-line [1, ∞), and satisfies ak = f (k) for all k ∈ Z>0 . Then the infinite
on the P
series ∞ k=1 ak and the improper integral
Z ∞
f (x) dx
1
or
n+1
X Z n+1 n
X
ak ≤ f (x) dx ≤ ak (5.3)
k=2 1 k=1
P∞
for all n ∈ Z>0 by (5.3), and so the series k=1 ak converges by 2.9.
Under the hypothesis of the integral test, it is sufficient to assume that there
is a positive integer N such that ak = f (k) for all k ≥ N , and f is nonincreas-
ing, nonnegative-valued, and continuous on [N, ∞). This remark is similar to one
made following the proof of the comparison test and follows from our earlier remark
(cf. page 70) about “neglecting the first few terms” in proving convergence or diver-
gence.
Let us consider some applications of the integral test.
5.8 Example: The two infinite series
∞ ∞
X 1 X 1
p
,
n=1
n k=2
k(log k)p
both converge for p > 1 and diverge for p ≤ 1. To see this, it is enough to apply the
integral test with the functions f (x) = x1p and f (x) = x(log1 x)p . In carrying out this
integration is is necessary to distinguish between p = 1 and p 6= 1. •
P∞ 1
The infinite series k=1 kp is called the p-series. Notice that, if p = 1, then the
p-series is just the harmonic series.
Let us turn to another convergence test.
5.9 Theorem: (Ratio test for positive series) Suppose that (ak ) is a sequence of
positive terms and assume that limk→∞ ak+1 ak
= L exists.
P∞
(i) If L < 1, then k=1 ak converges.
(ii) If L > 1, then ∞
P
k=1 ak diverges.
(iii) If L = 1, then this test is inconclusive.
Proof: (i) Suppose that L < 1 and let r be any number satisfying L < r < 1. Then
there must be N ∈ Z>0 such that 0 < ak+1ak
< r for all k ≥ N . Thus ak+1 < rak for
all k ≥ N and, therefore,
aN +1 ≤ aN r
aN +2 ≤ aN +1 r ≤ aN r2
aN +3 ≤ aN +2 r ≤ aN r3
..
.
aN +k ≤ aN rk
76 5 Infinite series
for k ≥ 1. Now ∞
P k
P∞
k=1 r is a convergent geometric series and, therefore, k=1 ak
converges by the comparison test (Theorem 5.5(i)).
(ii) Suppose that L > 1. Then there exists an N ∈ Z>0 such that ak+1
ak
≥ 1 for all
k ≥ N . This means that ak+1 ≥ ak for all k ≥ N and, therefore,
aN +1 ≥ aN
aN +2 ≥ aN +1 ≥ aN
aN +3 ≥ aN +2 ≥ aN
..
.
series whose behaviors we understand, then it is possible to make good use of the
comparison test.
The other three tests are more straightforward to use. If either the ratio or root
tests predicts the convergence of a series, then that series converges at least as rapidly
as a geometric series (this is clear from the proofs of these two tests). For series that
converge relatively slowly (such as the p-series), the ratio and root tests are necessarily
inconclusive and there is no alternative but to use the comparison or integral test, or
possibly some other device.
In applying the ratio and root tests the following simple lemma is sometimes
useful.
√
5.11 Lemma: limn→∞ n n = 1.
Proof: If f (x) = x1/x then
ln x 1/x
lim ln f (x) = lim = lim =0
x→∞ x→∞ x x→∞ 1
√
by l’Hôspital’s rule. Thus limx→∞ f (x) = e0 = 1 and so limn→∞ n
n = 1.
Let us consider an example that makes use of this fact.
5.12 Example: Consider the series ∞ n
P
n=0 nr , where r ≥ 0. Here
√ √
lim n nrn = lim r n n = r
n→∞ n→∞
by Lemma 5.11, and so the root test implies that the series converges if r < 1 and
diverges if r > 1. The root test is inconclusive if r = 1 but it is clear that this series
diverges in this case. •
These examples will show that this series belongs to case 4 and that it is very much
affected by permutations of its terms.
5.13 Examples: 1. The key observation (or trick) in studyingR 1 n−1the convergence of
the alternating harmonic series is to notice that 1/n = 0 x dx for all n ∈ Z>0
and thus that
m m Z 1
X (−1)n+1 X
n+1
= (−1) xn−1 dx
n=1
n n=1 0
Z 1X m
= (−1)n+1 xn−1 dx
0 n=1
Z 1Xm
= (−x)n−1 dx
0 n=1
1
1 − (−x)m
Z
= dx
0 1+x
Z 1 Z 1
dx (−x)m
= − dx
0 1+x 0 1+x
Z 1 m
m x
= ln 2 − (−1) dx.
0 1+x
2. This example continues the analysis of the alternating harmonic series in part 1.
In the notation introduced at the beginning of this section, we have
(
1/n, n odd,
pn =
0, n even
and (
1/n, n even,
qn =
0, n odd
and thus ∞
X 1 1
pn = 1 + 0 + + 0 + + 0 + ···
n=1
3 5
and ∞
X 1 1 1
qn = + 0 + + 0 + + 0 + ··· .
n=1
2 4 6
So each of n=1 pn and ∞
P∞ P
n=1 qn is more or less one half the sum of the harmonic
series, and hence they both diverge.
Now let us consider the series
1 1 1 1 1 1 1 1 1
1+ − + + − + + − + − ··· . (5.4)
3 2 5 7 4 9 11 6 13
Notice that this series is a permutation of the alternating harmonic series and
that, if (Sn ) denotes its sequence of partial sums, then
1 1 1 1 1 1
S3n+1 = 1 + − + + ··· + − + >1
3 2 5 4n − 1 2n 4n + 1
since
1 1 1
− + >0
2n − 1 2n 2n + 2
for all n ≥ 1. This inequality means that, if the series (5.4) converges, then its
limit exceeds 1 and hence does not converge to ln 2. The alternating harmonic
series and the series (5.4), therefore, behave very differently. •
80 5 Infinite series
The next theorem is very important because it sometimes allows us to use the
comparison, integral, ratio, or root test to establish convergence for series with pos-
itive and negative terms. Note, however, that the converse of this theorem is false
(cf. Example 5.13–1) so that tests for series with positive terms cannot be used to
prove divergence of series with terms of both signs.
P∞ P∞
5.14 Theorem: If n=1 |an | converges, then so does n=1 an .
Pn Pn
Proof: Put Sn = k=1 ak and Tn = k=1 |ak | for n ≥ 1. The hypothesis is that
the limit limn→∞ Tn exists and hence (Tn ) is a Cauchy sequence by Theorem 2.23.
So, given > 0, there will be an integer N ∈ Z>0 such that |Tn − Tm | < for all
m, n ≥ N . But then, for any two integers m and n satisfying n > m > N , we have
|Sn − Sm | = |am+1 + am+2 + · · · + an |
≤ |am+1 | + |am+2 | + · · · + |an |
= Tn − Tm < .
This means that (Sn ) is a Cauchy sequence and then 2.23 implies that it converges.
5.15 Definition: The series ∞
P
n=1 an is said
(i) to converge absolutely if the series ∞
P
n=1 |an | converges and
(ii) to converge conditionally if n=1 an converges but ∞
P∞ P
n=1 |an | diverges. •
This is, at first sight, a very strange definition. It is, in fact, one of the (few?)
situations in mathematics where terminology is a hindrance rather than an aid to
understanding. The first thing to note is that a conditionally convergent series is
convergent, the second is that (by Theorem 5.14) an absolutely convergent series
is convergent, and the third is that there are series that converge conditionally but
not absolutely (the alternating harmonic series, for example). When Theorem 5.14,
in conjunction with the comparison, integral, ratio or root test is used to establish
the convergence of a series with positive and negative terms, it is actually absolute
convergence which is being established. To emphasise this, we state the corresponding
versions of these theorems for general series.
5.16 Theorem: (Integral test for general series) Suppose that (ak ) is a sequence
and that f is a function which is continuous,3 nonnegative-valued, and nonincreasing
half-line [1, ∞), and satisfies |ak | = f (k) for all k ∈ Z>0 . Then the infinite
on the P
series ∞ k=1 ak converges if the improper integral
Z ∞
f (x) dx
1
converges.
Proof: By applying Theorem 5.7 to the series ∞
P P∞
k=1 |ak |, we see that the series k=1 ak
is absolutely convergent, and so convergent by Theorem 5.14.
3
The assumption that f is continuous is made merely to ensure that it is integrable over any
interval of finite length.
5.3 Convergence of general series 81
5.17 Theorem: (Ratio test for general series) Suppose that (ak ) is a sequence
of nonzero terms and assume that limk→∞ | ak+1 ak
| = L exists.
P∞
(i) If L < 1, then k=1 ak converges.
(ii) If L > 1, then ∞
P
k=1 ak diverges.
(iii) If L = 1, then this test is inconclusive.
P∞ P∞
Proof: (i) If L < 1 then the series k=1 |ak | converges. Thus k=1 ak converges
absolutely and so converges by Theorem 5.14.
(ii) Referring to the proof of the corresponding part of Theorem 5.9, we see that, if
L > 1, then there exists N ∈ Z>0 such that |ak | ≥ |aN | for all k ≥ N . This prohibits
limk→∞ ak = 0, and so prohibits convergence of the series by Theorem 5.3.
(iii) This follows from the corresponding part of Theorem 5.9.
5.18 Theorem: (Root test for general p series) Suppose that (an ) is a sequence of
nonzero terms and let L = lim supn→∞ |an |.
n
S2n = a1 + · · · + a2n
= a1 + · · · + a2n−2 + a2n−1 + a2n
= S2n−2 + a2n−1 + a2n
> S2n−2
for all n ∈ Z>0 , and so the sequence (S2n ) is strictly increasing. Moreover,
and so the sequence (S2n ) is bounded above by a1 . But then (S2n ) must be a conver-
gent sequence by Theorem 2.9; let S denote its limit. Now S2n+1 = S2n + a2n+1 for
all n ∈ Z>0 , and, since limn→∞ an = 0, it follows that
lim S2n+1 = lim (S2n + a2n+1 ) = S
n→∞ n→∞
P∞
and, therefore, limn→∞ Sn = S and so the series n=1 an converges.
Let us point out the significance of the three hypothesis. Condition (iii) says that
the terms are nonzero and alternate in sign and condition (ii) says that, in magnitude,
the terms are strictly decreasing. Of course, (i) is necessary for convergence by
Theorem 5.3.
We have already remarked that a series converges conditionally if the sum of its
positive terms diverges to ∞ and the series of its negative terms diverges to −∞, yet
the sum itself converges and has a finite value. This phenomenon can occur because of
a careful “balancing” or “near-cancellation” of the positive and negative contributions
within any particular partial sum. A special case of such “near-cancellation” occurs in
Example 5.13–2 and in the proof of Theorem 5.19. We might expect that permuting
the terms in a conditionally convergent series (a “rearrangement” of the series) would
alter this balancing act and hence change the sum of the series. This is indeed the
case and is exactly the point of the second part of Example 5.13–2.
5.20 Theorem: By permuting the terms of a conditionally convergent series we can
(i) make it converge to any number we wish or diverge to ±∞ or
(ii) make its partial sums oscillate between any two numbers.
Proof: We will only show how to rearrange the terms of a conditionally convergent
series so as to make it converge to a given positive number S; rearranging the terms
to make the series converge to a negative number or diverge to ±∞ is similar or to
make its partial sums oscillate is similar.
First we can reorder the sequence, which we denote by (an ), so that (|an |)
is nonincreasing, this being possible since conditional convergence ensures that
limn→∞ an = 0. First add together the positive terms from the beginning of the
sequence to make a sum that is larger than S (this is possible since the sum of the
positive terms diverges). Now add enough negative terms from the beginning of the
sequence to make the sum less than S. Next, add more positive terms, until the sum is
again larger than S, and then add negative terms to make the sum less than S. Since
the magnitudes of the terms is decreasing (this was the point of initially reordering
so that (|an |) is nonincreasing), the oscillations about S are becoming smaller and
smaller and so the rearranged series converges to S.
for any n ≥ 1.
Proof: First notice that an + an+1 has the same sign as an and the opposite sign as
an−1 by assumptions (ii) and (iii). Thus
∞
X
ak = |(an+1 + an+2 ) + (an+3 + an+4 ) + · · ·|
k=n+1
for all n ≥ 1.
84 5 Infinite series
Proof: If m and n are any two integers satisfying m > n ≥ 1, then the inequality (5.2)
implies
Z m m−1
X Z k+1
f (x) dx = f (x) dx
n+1 k=n+1 k
m−1
X m−2
X
≤ ak = ak+1
k=n+1 k=n
m−2
X Z k+1 Z m−1
≤ f (x) dx = f (x) dx.
k=n k n
The desired inequality now follows by taking the limit as m tends to infinity.
5.24 Example: What can we say about the error when the series ∞ 1
P
P10 1 P∞ n=1 n2 is approx-
imated by the sum n=1 n2 ? The error here is n=11 n12 and Theorem 5.23 implies
that Z ∞ ∞ Z ∞
1 1 X 1 1 1
= 2
dx ≤ 2
≤ 2
dx = .
11 11 x n=11
n 10 x 10
We thus have ! ∞
10 10
!
X 1 1 X 1 X 1 1
+ ≤ ≤ + ,
n=1
n2 11 n=1 n2 n=1
n2 10
and, since
10
X 1
1.5497 < < 1.5498,
n=1
n2
it follows that
∞
X 1
1.6406 ≈ 1.5497 + .0909 < < 1.5498 + 0.1000 ≈ 1.6498.
k=1
n2
Exercises
E5.1 Determine
P∞ whether the following series converge or diverge:
2
(a) √ 2n +3 P∞ n!
n=1 n2 +3n+2
P∞ 1·2·3·...·n
(e) n=1 nn
(b) n=1 3·5·7·...·(2n+1) P∞ n2 (n+1)n
P∞ (f) n=1 (2n)n
−n2
(c) n=1 ne P∞ e1/n
P∞ (n!)2 1000n (g) n=1 n2
(d) n=1 2n 2
is convergent.
2 −b2
Hint: Use the identity a − b = aa+b .
P∞
E5.5 Determine whether the series n=1 k + k1 (k 5 + ln k)−1/2 converges.
E5.9 In each of the following cases, determine all the values of x ∈ R for which the
series converges.
P∞ n2 n
(a) n=0 n+1 x
P∞ 2
x n
(b) n=1 1 + n
E5.10 Determine all the values of x for which the following series converge:
P∞ (n+3)x2n
(a) nn
Pn=1
∞ (n+1)(2x−1)n
(b) n=1 2n
E5.11 Determine all the values of x for which the following series converge:
P∞ (n+3)x2n
(a) nn
Pn=1
∞ (n+1)(2x−1)n
(b) n=1 2n
2
x n
E5.12 Determine all the values of x for which the infinite series ∞
P
n=1 1 + n con-
verges.
E5.13 Recall that the binomial coefficient nk is defined as k!(n−k)!
n!
. Determine all
P∞ 2n n
values of x ≥ 0 for which the series n=0 n x converges. What can you say
about the series when x < 0?
Hint: You may find the following bounds on 2n
n
to be useful. Use the grouping
to deduce that 2n 1
n
≥ ( 2n )4n .
E5.14 If x > y > 0, does the series ∞ 1
P
n=1 xn −y n converge or diverge? Why?
E5.15 (a) PProve that, if (an ) is a sequence of positiveP numbers such that the series
∞ ∞ 2
a
n=1 n converges, then so does the series n=1 an .
(b) Show by example that the converse of (a) is false.
P∞
(c) Give an example of
P∞ 2 a sequence (a n ) such that the series n=1 an converges
but the series n=1 an diverges.
E5.16 Let (ak ) be a sequence of nonnegative real numbers.
P∞ 2
(a) PShow that, if the series k=1 ak converges, then so does the series
∞ ak
k=1 k .
Hint: Apply the Cauchy–Schwartz inequality
v v
Xn u n
X
u n
X
u u
xk yk ≤ t x2k t yk2
k=1 k=1 k=1
P∞ ak
to the partial sums of k=1 k .
88 5 Infinite series
√
(a) Show that limn→∞ n an does not exist.
√
(b) Find lim supn→∞ n an .
(c) Does the series converge or diverge?
E5.19 (a) Show that, if (an ) is a sequence of positive numbers, then
√ an+1
lim sup n
an ≤ lim sup .
n→∞ n→∞ an
√
Hint: Show that, if L > lim supn→∞ an+1an
, then L ≥ lim supn→∞ n an .
(b) Explain the significance of this inequality with respect to the relative
strengths of the ratio and root tests.
n ∈ Z>0 and if the series ∞
P
E5.20 If an >
P0∞ for aall n=1 an , diverges prove that the
series n=1 1+an diverges.
n
converges and Z n Z ∞
lim f (x) dx = f (x) dx.
n→∞ 1 1
E5.33 Suppose we have an inexhaustible supply of uniform rectangular cards. Show
that, given any distance d, a finite number of these cards can be stacked on a
table so that the outside edge of the top card projects a distance d beyond the
edge of the table.
Chapter 6
Sequences of functions
In this chapter we are going to examine what it means for a sequence of functions
to converge to a function. This turns out not to be as straightforward as one might
think but, at the same time, is a very useful and powerful concept in analysis, applied
mathematics, and geometry. We will begin with an example that illustrates both
the use of convergence of functions and points out some of subtleties involved in a
discussion of such convergence.
y 0 = f (x, y)
y(x0 ) = y0
(This equivalence is easy to check using the fundamental theorem of calculus.) The
point of replacing the original initial value problem by the integral equation is simply
that integration is much easier to deal with than differentiation.
We will solve this integral equation by finding a sequence of approximate solutions
that converges to the actual solution. The initial approximation is the constant
function y0 satisfying y0 (x) = y0 for all x, and the successive approximations y1 , y2 , . . .
92 6 Sequences of functions
The question now, of course, is whether this sequence of functions converge (in some
sense) to a function y and, if so, whether this limit function is a solution of (6.1). If
we believe that the sequence (yn ) converges to a function y, it is tempting to argue
as follows:
There are two questionable steps here, in the fourth and fifth lines. The fourth
line involves interchanging the limit and the integral and is not always valid. The
conditions under which this is valid is one of the topics to be studied in this chapter
(cf. Theorem 6.9). The fifth line is also contentious although it is clearly related to
the continuity of f . But, at any rate, if this calculation can be justified, then y is a
solution of the given initial value problem.
Many of the classical methods of solving ordinary and partial differential equations
involve series of functions (for example, power, Fourier, and Bessel series). To under-
stand and justify the use of such methods, it is necessary to understand convergence
of sequences and series of functions.
infinite length.) Then, for each point x ∈ I, we have a sequence (fn (x)) of numbers.
In fact, we can think of fn (x) as a function of the two variables n ∈ Z>0 and x ∈ I,
and so examining this sequence for a fixed value of x is not unlike examining a
function of two real variables along, say, a vertical line in the plane. For each fixed
value of
P∞ x ∈ I, we can ask whether the limit limn→∞ fn (x) exists or whether the
series n=1 fn (x) converges. The point here is that, since x is fixed, the discussion
of convergence of sequences in Chapter 2 and of series in Chapter 5 can be brought
to bear on these questions. Since this sequence and series are being considered for
a fixed value of x, the adjective “pointwise” is attached to these questions and their
answers to distinguish them from a related situation we will consider later in this
chapter.
6.1 Definition: Let I ⊆ R be an interval and suppose that (fn ) is a sequence of
real-valued functions defined on I. We will say that
(i) the sequence (fn ) converges pointwise on I if the limit limn→∞ fn (x) exists
for each point x ∈ I and that
(ii) the series ∞
P P∞
n=1 fn converges pointwise on I if the series n=1 fn (x) con-
verges for each point x ∈ I. •
Suppose that (fn ) is a sequence of functions defined on an interval I ⊆ R. If
this sequence converges pointwise on I, we can define a function f on I by letting
f (x) = limn→∞ fn (x) for all x ∈ I,P and, in this case, we will write “limn→∞ fn = f
∞
pointwise”. Similarly, if the series P n=1 fn converges pointwise on I, we can define
a function fPon I by letting f (x) = ∞ n=1 fn (x) for all x ∈ I, P
and in this case, we
“ ∞
will write P f
n=1 n = f pointwise”. Notice that, if we let gn = n
k=1 fk for n ∈ Z>0 ,
∞
then f = n=1 fn pointwise if and only if f = limn→∞ gn pointwise.
According to Definition 2.4, this means that the following condition holds.
6.2 Lemma: A sequence (fn ) of functions on an interval I ⊆ R converges pointwise
if and only if the following condition holds:
for each > 0 and for each x ∈ I, there is an N ∈ Z>0 such that
|fn (x) − f (x)| < for all n ≥ N .
Note that here N is chosen after and x are selected, and hence that N depends
on and x, i.e., N = N (, x).
Let us apply these definitions in an example.
6.3 Example: It is easy to see that
(
1, x = 1,
lim xn =
n→∞ 0, 0 ≤ x < 1.
then limn→∞ fn = f pointwise on [0, 1]. Notice that each fn is continuous and that
the limit function f is discontinuous.
If and x are two numbers in the open interval (0, 1), then |fn (x) − f (x)| = xn ,
ln
and thus |fn (x) − f (x)| < if and only if n > ln x
. So the smallest possible value of
ln
N satisfying Lemma 6.2 is the smallest integer larger than ln x
. Thus N does indeed
depend on both and x and, for a fixed ∈ (0, 1), limx→1− N (, x) = ∞. •
Pointwise convergence is a very weak notion of convergence and consequently not
a very useful one. That is to say, one can draw very few conclusions from the assertion
that limn→∞ fn = f pointwise on I. For example, the preceding example shows that
1. f need not be continuous, even if each of the fn is C∞ ,
and the next few examples will show that
2. f need not be integrable, even if each of the fn is integrable,
3. if f and each of the fn is continuous, then it is not necessarily true that
Z b Z b
lim fn (x) dx = f (x) dx,
n→∞ a a
6.4 Example: For each n ∈ Z>0 , let an be a positive number and let fn be the
function with domain [0, 1] which is zero on [1/n, 1] and whose graph over [0, 1/n] is
an isosceles triangle with base [0, 1/n] and height an . Formally,
2nan x,
0 ≤ x ≤ 1/(2n),
fn (x) = 2an − 2nan x, 1/(2n) ≤ x ≤ 1/n,
0, 1/n ≤ x ≤ 1.
The sequence (fn ) converges pointwise to the zero function on [0, 1], regardless of the
choice of the sequence (an ). Indeed, it is certainly true that fn (0) = 0 for all n. And,
if x ∈ (0, 1] and if N ∈ Z>0 such that 1/N < x, then fn (x) = 0 for all n ≥ N , and,
therefore, limn→∞ fn (x) = 0. Notice that, just as in the preceding example, the N of
Lemma 6.2 appears to depend on both x and .
It is easy to see that, by choosing the sequence (an ) appropriately, the limit
Z 1
an
lim fn (x) dx = lim
n→∞ 0 n→∞ 2n
The example just presented, where pointwise convergence of functions does not
imply the convergence of their integrals, is perhaps artificial in that the functions are
piecewise linear, but not C∞ . The next example exhibits this same phenomenon with
C∞ functions.
n
6.5 Examples: 1. If fn (x) = nx(1 − x2 ) and gn (x) = nfn (x) for x ∈ [0, 1], and
n ∈ Z>0 , then
lim fn (x) = 0, lim gn (x) = 0
n→∞ n→∞
4. Notice that, in the preceding example, the partial sums of the series are C∞
functions, but that the pointwise limit of the series is not even continuous. Let’s
explore this a little more deeply.
Let fn (x) = sin√nx
n
for n ∈ Z>0 and x ∈ R. Then limn→∞ fn = 0 pointwise on R
since |sin nx| ≤ 1. In fact, if > 0 is given and if N is an integer with 1/N < ,
then |fn (x)| < for all n ≥ N and all x. So, in this example, the N of Lemma
√ 6.2
depends only on , in contrast to several earlier examples. Here fn0 (x) = n cos nx,
and thus limn→∞ fn0 (x) = ∞ for every x, and hence the sequence (fn0 ) does not
converge pointwise. •
for all x ∈ R and all n > −2 and, therefore, limn→∞ sin√nx n
= 0 uniformly on R and
hence on any interval.
4. This example is a continuationP∞ of Example 6.5–3. If f is as in that example and if
x2
I is a closed interval, then n=0 (1+x2 )n = f (x) uniformly on I if and only if 0 6∈ I.
To see that this it is necessary, in view of the definition of uniform convergence
of a series, to estimate the partial sums. First of all, if x 6= 0 we have
n n
!
X x2 X
2 )k
= (1 + x2 )−k x2
k=0
(1 + x
k=0
1 − (1 + x2 )−n−1
= x2
1 − (1 + x2 )−1
= 1 + x2 − (1 + x2 )−n
and, therefore,
(
n 2
X x 0, x = 0,
− f (x)=
2
(1 + x ) k 2 −n
k=0
6 0.
(1 + x ) , x =
Suppose first that 0 ∈ I and assume that 0 is not √ the right end point of I. Then
I = [a, b], where a ≤ 0 < b, and the number xn = 21/n − 1 belongs to I at least
for sufficiently large values of n. Now
n
X x2n −n 1
− f (x ) = 1 + x2n − 1 + x2n − (1 + x2n ) =
2 k n
(1 + xn ) 2
k=0
98 6 Sequences of functions
for all sufficiently large values of n by the above calculation, and the formula for
f contained in Example 6.5–3. A similar argument will show that the series does
not converge uniformly if 0 is not the left end point of I.
Now suppose that 0 6∈ I and assume, to be definite, that I consists of positive
numbers. Then I = [a, b], where 0 < a < b, and
n 2
X x
− f (x) = (1 + x2 )−n ≤ (1 + a2 )−n
2
(1 + x ) k
k=0
for all x ∈ I and all n ∈ Z>0 . Now limn→∞ (1 + a2 )−n = 0 and so, given an > 0,
there will be an integer N ∈ Z>0 such that (1 + a2 )−N < . But this means that
n
2
X x
− f (x)<
(1 + x2 )k
k=0
for all n ≥ N and all x ∈ I and, therefore, the series converges uniformly on I. A
similar argument will evidently show that the series converges uniformly on I if I
consists of negative numbers.
Finally, note that, if 0 ∈ I, then it also follows from Theorem 6.8 (below), and
the fact that f is discontinuous at 0, that the convergence is not uniform on I. •
for all x ∈ I and all n ∈ Z>0 . Now, since limn→∞ fn = f uniformly on I, there is an
integer N ∈ Z>0 such that |f (x) − fn (x)| < for all x ∈ I and all n ≥ N and then,
since fN is continuous, there is a number δ > 0 such that |fN (x) − fN (x0 )| < /3 for
all x ∈ I satisfying |x − x0 | < δ. But then, if x is any point in I satisfying |x − x0 | < δ
and if n = N , each of the three summands in the last line of the above estimate is
less than /3 and, therefore,
|f (x) − f (x0 )| <
for all such x.
While Example 6.5–1 showed that we cannot interchange a pointwise limit and
an integral, the next result shows that we can interchange a uniform limit and an
integral.
6.9 Theorem: Suppose that (fn ) is a sequence of continuous functions defined on an
interval [a, b] and assume that there is a function f on I such that limn→∞ fn = f
uniformly on [a, b]. Then
Z b Z b
lim fn (x) dx = f (x) dx.
n→∞ a a
Proof: Recall from Theorem 6.8 that f is continuous and hence that all of the integrals
in the statement of the theorem exist. It is necessary to show that, given an > 0,
there is an integer N ∈ Z>0 such that
Z b Z b
fn (x) dx − f (x) dx <
a a
for all n ≥ N . Now, since limn→∞ fn = f uniformly on [a, b], it follows that there is
an integer N ∈ Z>0 such that |fn (x) − f (x)| < 2(b−a) for all x ∈ [a, b] and all n ≥ N .
But then
Z b Z b Z b
fn (x) dx − f (x) dx ≤
|fn (x) − f (x)| dx
a a a
Z b
≤ dx
a 2(b − a)
= <
2
for all n ≥ N .
We already know from Example 6.5–4 that, if limn→∞ fn = f pointwise, then it
does not follow that limn→∞ fn0 = f 0 pointwise and, in analogy with the preceding
theorem, we might hope to be able to prove that, if limn→∞ fn = f uniformly, then
limn→∞ fn0 = f 0 uniformly. But Exercise 6.7 and the next example both show that this
is false in general, although Theorem 6.11 shows it is true under certain hypotheses.
The point here is that Z x
g(x) = f (t) dt
a
100 6 Sequences of functions
is a more smooth function than f , but that f 0 is a less smooth function. (For instance,
if f is, say, C4 , then g is C5 , whereas f 0 is C3 .) The effect of this is that, in the
above theorem on integration, it was not necessary to make assumptions beyond
those involving the functions themselves, whereas in the theorem on differentiation
(see below) it is necessary to make assumptions about both the functions and their
derivatives.
x
6.10 Example: Let fn (x) = for n ∈ Z>0 and x ∈ [−1, 1]. It follows easily
1 + n 2 x2
from elementary calculus1 that
1 1
− ≤ fn (x) ≤
2n 2n
for all n ∈ Z>0 and all x ∈ [−1, 1], and, therefore, limn→∞ fn = 0 uniformly on [−1, 1].
On the other hand,
0 1 − n 2 x2
fn (x) =
(1 + n2 x2 )2
and so (
1, x = 0,
lim fn0 (x) =
n→∞ 0, 0 < |x| ≤ 1.
So here limn→∞ fn0 6= (limn→∞ fn )0 and, moreover, since limn→∞ fn0 is not continuous,
the sequence (fn0 ) does not converge uniformly by Theorem 6.8. •
The derivative of a function defined on an interval is generally only considered at
interior points of that interval and, for this reason, a function is usually only described
as being C1 if its domain is an open interval. For the purpose of the next theorem and
corollary, it will be convenient to modify this terminology a little. Namely, suppose
that h is a function defined on a closed interval [a, b]. Then, for this theorem and
corollary only, we will let
the right derivative of h at a, x = a,
0
h (x) = the derivative of h at x, a<x<b
the left derivative of h at b, x = b.
(iii) there is a point c ∈ [a, b] for which the limit limn→∞ fn (c) exists.
Then there is a function f on [a, b] such that limn→∞ fn = f uniformly on [a, b] and
f 0 = g.
Proof: The function f that is asserted to exist must satisfy the equation
Z x Z x
f (x) = f (a) + g(t) dt = lim fn (a) + g(t) dt
a n→∞ a
by the fundamental theorem of calculus. (Notice that g is continuous by assump-
tions (i) and (ii), and Theorem 6.8 and hence is integrable.) This suggests that we
attempt to carry out the proof by
1. first showing that the limn→∞ fn (a) exists and defining a function f on [a, b] by
the formula Z x
f (x) = lim fn (a) + g(t) dt, (6.2)
n→∞ a
2. showing that limn→∞ fn = f uniformly on [a, b], and
3. showing f 0 = g.
The proof of 1 is easy. We know that
Z c
fn (c) − fn (a) = fn0 (t) dt
a
for all n ∈ Z>0 by the fundamental theorem of calculus and that limn→∞ fn0 = g
uniformly on [a, c], and so
Z c
0
lim fn (a) = lim fn (c) − fn (t) dt
n→∞ n→∞ a
Z c
= lim fn (c) − g(t) dt
n→∞ a
by Theorem 6.9.
We can now use (6.2) to define a function f on [a, b].
Since g is continuous, the definition of f and the fundamental theorem of calculus
imply that f satisfies 3. To show that f also satisfies 2, suppose that > 0 is given.
Then (by (ii)) there will be an integer N ∈ Z>0 such that |fn0 (x) − g(x)| < 2(b−a)
for all x ∈ [a, b] and all n ≥ N and such that |fm (a) − limn→∞ fn (a)| < /2 for all
m ≥ N . It now follows from the fundamental theorem of calculus and the definition
of f that
Z x Z x
0
|fm (x) − f (x)| = fm (a) +
fm (t) dt − lim fn (a) + g(t) dt
a n→∞ a
Z x
0
≤ fm (a) − lim fn (a) + (fm (t) − g(t)) dt
n→∞ a
Z x
0
< + |fm (t) − g(t)| dt
2 a
≤ + =
2 2
102 6 Sequences of functions
for all x ∈ I. Now, since (Tn ) is a Cauchy sequence, this calculation implies that
(Sn (x)) is a Cauchy sequence for each x ∈ I and, therefore, that the sequence (Sn )
converges pointwise by Theorem 2.23. So, if we put f (x) = limn→∞ Sn (x) for each
6.5 The sup-norm 103
x ∈ I, then f = ∞
P
n=1 fn pointwise on I, and it remains to show that this convergence
is actually uniform. Now, if x is any point in I, then
Xn
f (x) − fn (x) = |f (x) − Sm (x) + Sm (x) − Sn (x)|
k=1
≤ |f (x) − Sm (x)| + |Sm (x) − Sn (x)|
≤ |f (x) − Sm (x)| + Tm − Tn
Now, if a number > 0 is given, there will be an integer N ∈PZ>0 such that
n
limm→∞ Tm − Tn < whenever n ≥ N , and, therefore, P∞ |f (x) − k=1 fn (x)| <
for all n ≥ N and all x ∈ I. This makes it clear that n=1 fn = f uniformly on I.
the series ∞ 1
P
and sinceP n=1 n2 converges by Example 5.8.
∞ 1
Now n=1 n3 sin(nx) converges to 0 when x = 0. Hence, by Theorem 6.11, we
conclude that the series may be differentiated term-by-term, and
∞ ∞
d X 1 X d 1
sin(nx) = sin(nx) . •
dx n=1 n3 n=1
dx n3
these sets of functions is a subset of the other, although if I is closed and bounded
then C(I) ⊆ B(I) by Theorem 4.24(i). Both of these sets of functions are (infinite-
dimensional) vector spaces in the sense that, if f and g are functions in one of these
sets and if a is a number, then the functions f + g and af defined by the equations
and
(af )(x) = af (x)
also belong to that set. This is known in the case of C(I) (since linear combinations
of continuous functions are again continuous) and in the case of B(I) can be shown as
follows. If f and g are in B(I) then there are numbers M and N satisfying |f (x)| ≤ M
and |g(x)| ≤ N for all x ∈ I, and consequently
and
|(af )(x)| = |af (x)| = |a||f (x)| ≤ |a|M
for all x ∈ I. The origin in both of these vector spaces is the zero-function.
For any function f ∈ B(I), the quantity
kf k = sup{|f (x)| | x ∈ I}
is a finite number called the sup-norm of f on I. (Notice that the sup-norm k·k
depends on I. but that I is not included in the notation.) Thus k·k is a function on
B(I) and it has the three properties:
1. kf k ≥ 0 for all f ∈ B(I) with equality if and only if f = 0;
2. kaf k = |a| kf k for all a ∈ R and all f ∈ B(I);
3. kf + gk ≤ kf k + kgk for all f, g ∈ B(I).
Only the third of these properties is not obvious and it can be proven as follows. If
f and g are two functions in B(I) and if x is a point in I, then
and, therefore,
The sup-norm on B(I), therefore, has the same basic properties as the usual norm
has on Rd (cf. the beginning of Chapter 4). Now a cursory examination will reveal
that all of the ideas presented in Chapter 4 depended ultimately on the balls Br (x)
and that the definitions of these balls depended only on the norm. This suggests
that it should be possible to carry over to B(I) all of the definitions that were made
for Rd in Chapter 4 and most (but definitely not all) of the results that were proven
6.5 The sup-norm 105
there. In doing this we must replace Rd by B(I) and regard this as the set we are
studying and we must replace the vectors in Rd by the functions in B(I) and regard
these as the points in the set we are studying. Moreover, the functions that were
studied in Section 4.4 (which all had subsets of Rd as their domains) must now be
replaced by functions defined on subsets of B(I), i.e., functions whose argument is
itself a function. (We already know many examples of such functions: evaluation,
differentiation, and integration are just three such examples).
The definitions of distance between points in B(I) and convergence of sequences
in B(I) are completely analogous to those in Rd .
6.15 Definition: Let I be an interval in R.
(i) The distance between two functions f and g in B(I) is the number |f − g|.
(ii) A sequence (fn ) in B(I) is said to converge to a function f ∈ B(I) if, for each
> 0, there is an integer N ∈ Z>0 such that kfn − f k < whenever n ≥ N . •
Consider two functions f and g in B(I) and an > 0. Now
and so, if kf − gk < , then |f (x) − g(x)| < for all x ∈ I and, conversely, if
|f (x) − g(x)| < for all x ∈ I, then kf − gk ≤ . In particular, if (fn ) is a sequence in
B(I), this observation takes the following form: If kfn − f k < for all n ≥ N , then
|fn (x) − f (x)| < for all x ∈ I and all n ≥ N and, conversely, if |fn (x) − g(x)| <
for all x ∈ I and all n ∈ N , then kfn − f k ≤ for all n ∈ Z>0 . This just means that
convergence in B(I) in the sense of Definition 6.15 is the same as uniform convergence
on I. This observation can be summed up as follows:
6.16 Proposition: A sequence (fn ) in B(I) converges to a function f ∈ B(I) (in the
sense of Definition 6.15) if and only if limn→∞ fn = f uniformly on I.
Consider again an interval I in R. A sequence (fn ) of functions in B(I) is said to be
pointwise Cauchy if the numerical sequence (fn (x)) is Cauchy for all x ∈ I. With
this definition, it follows from Theorem 2.23 that the sequence (fn ) converges point-
wise on I if and only if it is pointwise Cauchy. The following definition and theorem
capture the corresponding definition and characterization for uniform convergence.
6.17 Definition: Suppose that I is an interval in R. A sequence (fn ) of functions in
B(I) is said to be uniformly Cauchy on I if, for each > 0, there exists an integer
N such that kfn − fm k < for all m, n > N . •
Exercises
E6.1 Show that each of the following sequences of functions converges pointwise on
[0, 1] to the zero function and determine whether the convergence is uniform
on that interval.
(a) fn (x) = nxe−nx
2
(b) fn (x) = nxe−n x
2
(c) fn (x) = nxe−nx
E6.2 Determine if the following sequences (fn ) of functions on [0, 1] converge point-
wise and/or uniformly.
1
2nx,
x ∈ [0, 2n ],
1 1
(a) fn (x) = 2 − 2nx, x ∈ ( 2n , n ],
0, x ∈ ( n1 , 1]
sin(1/n2 x)
(
√
n
, x ∈ (0, 1],
(b) fn (x) =
0, x=0
n
X cos(kx)
(c) fn (x) =
k=1
k
E6.3 Let fn (x) = x1 + n1 and let gn (x) = 1 + n1 for n ∈ Z>0 and x ∈ (0, 1).
(a) Show that limn→∞ fn = 1/x uniformly on (0, 1).
(b) Show that limn→∞ gn = 1 uniformly on (0, 1).
(c) Show that limn→∞ fn gn = 1/x pointwise, but not uniformly, on (0, 1).
E6.4 Determine the pointwise limit limn→∞ fn of the given sequence of functions on
the interval [0, 1], and also determine whether the convergence is uniform on
that interval.
(a) fn (x) = 1/(1 + nx)
(b) fn (x) = x/(1 + nx)
E6.5 Compute the pointwise limits of the following sequences of functions on the
indicated interval and explain why the convergence is not uniform.
n 2 xn
(a) Sn (x) = on [0, 2]
1 + n2 xn
1 − n2 x2
(b) Sn (x) = on [−1, 1]
1 + n2 x2
nx2
E6.6 Let fn (x) = 1+nx 2 for x ∈ R and n ∈ Z>0 .
E6.16 Suppose that f is a real-valued function with domain R and with the property
that |f (x) − f (y)| ≤ K|x − y| for all x, y ∈ R and some K > 0. Suppose also
that (gn ) is a sequence of functions from R to R which converges uniformly
on R to a function g. Prove that the sequence (f ◦ gn ) converges uniformly to
f ◦ g on R.
E6.17 Suppose that (fn ) is a sequence of differentiable functions on [0, 1] that con-
verges uniformly to some function. Must the sequence fn0 converge uniformly?
Hint: Consider the functions fn (x) = (sin(n2 x))/n or gn (x) = xn+1 /(n + 1)
on [0, 1].
E6.18 Recall the series ∞ 2n
P n
n=0 n x from Exercise E5.13. It was determined there
that the series converges pointwise on (−1/4, 1/4).
P∞ 2n n
(a) Show that the series n=1 n x converges uniformly on the interval
[−a, a] for any real number a satisfying 0 < a < 1/4.
(b) Set f (x) = ∞ 2n
P n
n=1 n x forP x ∈ (−1/4,
1/4). Explain why the derivative
0 ∞ 2n n−1
of f (x) is given by f (x) = n=1 n n x for any x ∈ (−1/4, 1/4).
P∞
E6.19 Show that, if the series n=1 un (x) converges uniformly on an interval I, then
limn→∞ un = 0 uniformly on I.
E6.20 (a) Use the Weierstrass M -test to show that the series ∞ −3/2
P
n=1 n cos(x/n)
converges uniformly on R. Let S denote its sum.
(b) Deduce from (a) and Theorem 6.11 that the series ∞ −1/2
P
n=1 n sin(x/n)
converges uniformly on all closed intervals of finite length, and that its
derivative Pis the function S(x) of (a). This theorem does not imply that
the series ∞ n=1 n
−1/2
sin(x/n) converges uniformly on R. Why?
(c) Use the Weierstrass M -test and the inequality P∞ |sin−1/2
u| ≤ |u| (which is valid
for all u ∈ R) to prove that the series n=1 n sin(x/n) converges
uniformly on all closed intervals of finite length.
(d) Does the series ∞ −1/2
P
n=1 n sin(x/n) converges uniformly on R?
E6.21 Let Sn (x) = nx/(nx + 1) for n ≥ 1 and x ∈ [0, 1].
(a) Compute the pointwise limit limn→∞ Sn (x) on [0, 1]. Is the convergence
uniform on this interval?
R1
(b) Compute limn→∞ 0 Sn (x) dx and
Z 1
lim Sn (x) dx.
0 n→∞
E6.22 (a) Consider functions defined on a subset S of R2 . How should the definition
of “sup-norm of a function with respect to S” be defined?
(b) Suppose that (gn ) is a sequence of functions defined a compact subset S
of R2 which converges uniformly on S to a function g. Prove that
ZZ ZZ
lim gn (x, y) dA = g(x, y) dA.
n→∞ S S
Exercises for Chapter 6 111
P∞ sin nx
E6.23 If f (x) = n=1 n3
, show that
Z π ∞
X 1
f (x) dx = 2 .
0 n=1
(2n − 1)4
for all x > 0 and deduce that f is strictly decreasing on (0, ∞).
(d) Let fn (x) = nk=1 1+k1 2 x be the nth partial sum of f . Use the fact that
P
limx→∞ fn (x) = 0 for each fixed n ∈ Z>0 to deduce that limx→∞ f (x) = 0.
(e) If fn is as in (d), show that fn (1/n2 ) > n/2 and deduce that
limx→0+ f (x) = ∞.
(f) Use the information in parts (a)–(e) to make a rough sketch of the graph
of f .
E6.25 Show that the series ∞ n−1 1
P
n=1 (−1) x2 +n
is uniformly convergent on R, but is
not absolutely convergent for any x ∈ R.
Hint: Use the Cauchy criterion for uniform convergence.
E6.26 For each n ∈ Z>0 , let fn be the function with domain R defined as follows:
(
x/n, |x| ≤ n
fn (x) =
1, |x| > n.
(a) Determine all the values of x for which the limit limn→∞ fn (x) exists.
(b) Determine the sup-norm kfn k of fn on the interval [−a, a] for each n ∈ Z>0
and each a > 0.
Hint: Consider separately the two cases n > a and n ≤ a.
(c) For which numbers a > 0 does the sequence (fn ) converge uniformly on
the interval [−a, a]?
112 6 Sequences of functions
(b) Show that if, for t = 0, the series in (a) is uniformly convergent for
x ∈ [0, π], then u(x, t) is continuous on the set
Remark: The partial differential equation uxx = ut is the heat equation and
can be derived from conservation of heat and Gauss’s theorem. The function
u is the expansion in a Fourier series of the solution of the heat equation.
Chapter 7
Power series and Taylor series
An important class of series of functions are the power series, such as arise in the
theory of Taylor series. The theory of Taylor series, then, plays an important role in
complex analysis.1 In this chapter we shall consider the special properties of power
series.
and so, by the ratio test (Theorem 5.9), the power series converges for |x| < 1
and diverges for x > 1. As we shall see in the proof of Theorem 7.3 below, the
1
While we consider here only real-valued functions of a real variable, the theory for complex-
valued functions of a complex-variable is hardly any different.
7.2 Convergence of power series 115
series also diverges for x < −1, although this does not follow immediately from
the ratio test as stated in Theorem 5.9 since the series is alternating. In addition,
the series clearly diverges for x = 1 and converges for x = −1, since then it is just
the alternating harmonic series.
P∞ an
For any a ∈ (0, 1), the series n=1 n converges and it follows from this, the
n
Weierstrass M -test, and the inequality |x|n ≤ an , which is valid for |x| ≤ a,
n
that the power series in question converges uniformly on the interval [−a, a]. On
the other hand, it is not hard to show that the power series does not converge
uniformly on the interval (−1, 1). In fact, if it did, then (taking = 1 in the
definition P
of uniform convergence) there would be an integer N with the prop-
erty that ∞ n=N
xn
n
< 1 for all x ∈ (−1, 1), and hence with the property that
PM xn P∞ 1
n=N n < 1 for all integers M > N . Now the series P n=N n diverges to ∞, and
so there will be an integer M satisfying M > N and M 1
n=N n > 1. Since this is
and 1/n
n2 x2n n2/n x2 x2
lim = lim =
n→∞ 2n n→∞ 2 2
by
P∞Example 6.5–1, and both the √ratio and the root tests assert
√ that the series
n2 x2n
converges for |x| < 2 and diverges for |x| > 2. It is obvious
n=0 2n √
that this power series diverges for x = ± 2. Finally, the argument using the
Weierstrass M -test given in 1 will work here, and leads to the conclusion
√ that this
series converging uniformly on the interval [−a, a] for any a ∈ [0, 2). •
by Lemma 5.11 and so R = 1. For the power series considered in Example 7.1–2, we
have 2 1/(2n)
p
n n 1
lim sup |an | = lim sup n
=√
n→∞ n→∞ 2 2
√
and so R = 2. •
Then
p
n |x − c|
L = |x − c| lim sup |an | =
n→∞ R
and so, if |x − c| < R, then L < 1 and the series converges absolutely by the root
test (Theorem 5.10). On the other hand, if |x − c| > R, then L > 1, and we conclude
divergence of the series by the general version of the root test (Theorem 5.18).
7.2 Convergence of power series 117
Now consider a numberP r ∈ (0, R). Then |an (x−c)n | ≤ |an |rn for all x ∈ [c−r, c+r]
and all n, and the series ∞ n
n=0 |an r | converges by the first part of the proof, and,
therefore, the power series converges uniformly on the interval [c − r, c + r] by the
Weierstrass M -test (Theorem 6.13). This completes the proof of the theorem.
Let us consider a few interesting facts about radii of convergence of power series.
To do this, let us begin by proving a generalization of Lemma 5.11.
7.4 Lemma: If P is a polynomial function, then
p
lim n |P (n)| = 1.
n→∞
We have
p
n ln(|pk |)
lim |pk | = lim exp
n→∞ n→∞ n
ln(|pk |)
= exp lim = 1.
n→∞ n
By Lemma 5.11 we have
√
n √ k √ k
lim nk = lim n
n = lim n n = 1.
n→∞ n→∞ n→∞
R
has radius of convergence .
|b|
Proof: (i) and (ii) We shall prove that the power series
∞
X P (n)
(x − c)n
n=0
Q(n)
p
= lim sup n |an |,
n→∞
R
is , as stated.
|b|
We next calculate
n/(n−1) n
lim sup|an−1 | = lim sup exp ln(|an−1 |)
n→∞ n→∞ n−1
n
= exp lim sup ln(|an−1 |)
n→∞ n + 1
= exp lim sup ln(|an−1 |)
n→∞
Therefore,
1/n
p
lim sup n |an−1 | = lim sup|an−1 |
n→∞ n→∞
n/(n−1)
= lim sup|an−1 |
n→∞
In similar manner, p p
n n
lim sup |an+1 | = lim sup |an |.
n→∞ n→∞
and
∞
! ∞
d X
n
X
an (x − c) = nan (x − c)n−1
dx n=0 n=1
or ∞
x2 x3 X (−1)n+1
ln(1 + x) = x − + − ··· = xn . (7.2)
2 3 n=0
n
The two power series (7.1) and (7.2) both have radius of convergence of 1.
We can combine the above power series for ln(1 + x) and ln(1 − x) to obtain
1+x
ln = ln(1 + x) − ln(1 − x)
1−x
1 2 1 3
= x − x + x − ···
2 3
1 2 1 3
+ x + x + x + ···
2 3
3 5
x x
=2 x+ + + ···
3 5
∞
X x2n+1
=2 .
n=0
2n + 1
1 1
arctan x = x − x3 + x5 − · · · ,
3 5
and this
√ power series too has a radius of convergence of 1. In particular, letting
x = 1/ 3 in this series gives
π 1 1 1 1
= √ 1− + − + ··· . •
6 3 3 · 3 5 · 32 7 · 33
The theorem on differentiation and integration of power series has the following
two consequences.
7.4 Taylor series and Taylor polynomials 123
P∞ n
7.10 Corollary: A power series n=0 an (x − c) whose radius of convergence R is
∞
positive is a C -function on the interval interval(c − R, c + R) and
∞
!
dk X
n
an (x − c) = k!ak
dxk n=0
x=c
and ask whether it converges on some interval of positive radius and, if so, whether
it converges to f on that interval. This power series is called the Taylor series of
f about c. Now the question of the convergence of this power series is that of the
convergence of its sequence of partial sums, and this suggests that we consider its
partial sums, i.e., the polynomials
n
X f (k) (c)
Pn (x) = (x − c)k (7.3)
k=0
k!
for n ∈ Z>0 . The polynomial (7.3) is called the nth Taylor polynomial of f at
(k)
c and has the property that f (k) (c) = Pn (c) for k ∈ Z≥0 . This property suggests
that Pn should be a good approximation to f near c since the two functions f and
Pn have the same value, the same slope (and hence the same tangent line), the same
concavity, etc., at c. In fact, if we let Rn = f − Pn , then f = Pn + Rn , and Rn is just
the error when we regard Pn as an approximation to f . The question of how good an
approximation Pn is to f , therefore, becomes the question of how small Rn is and, in
particular, the question of whether limn→∞ Pn (x) = f (x) for a particular value of x
is the question of whether limn→∞ Rn (x) = 0. To answer this question we evidently
need an expression for Rn and the next theorem contains two such expressions. Notice
that this theorem does not require that f be a C∞ -function, but merely of class Cn+1
on I.
7.12 Theorem: Suppose that f is a function defined on an open interval I, that c is
a point in I, and that n is a positive integer. If f (n+1) exists and is continuous on I
and if Rn is as above, then
1 x
Z
Rn (x) = (x − u)n f (n+1) (u) du (7.4 )
n! c
and there is a point ξ ∈ I such that
f (n+1) (ξ)
Rn (x) = (x − c)n+1 . (7.5 )
(n + 1)!
Proof: The best way to prove this theorem is probably to first prove a lemma and
then deduce the theorem from the lemma. The reason for is two-fold: the lemma is of
independent interest and doing it this way will make the proof easier to comprehend.
Part (ii) of the lemma is a special case of the generalized mean value theorem for
integrals (cf. Exercise E7.35).
7.4 Taylor series and Taylor polynomials 125
Proof: (i) This follows easily from the formula for integration by parts:
Z x Z x
j −1 d j+1
(x − u) g(u) du = (x − u) g(u) du
c j+1 c dx
−1 x
= (x − u)j+1 g(u)
j+1 c
Z x
j+1 0
− (x − u) g (u) du
c
Z x
g(c) 1
= j+1
(x − c) + (x − u)j+1 g 0 (u) du.
j+1 j+1 c
(ii) Fix a point x ∈ I \ {c} and assume that x > c (the proof in the case x < c is
similar). If
m = inf{g(u) | c ≤ u ≤ x}
and
M = sup{g(u) | c ≤ u ≤ x},
then it is certainly true that m ≤ g(u) ≤ M for all u ∈ [c, x]. Since x − u ≥ 0 for all
u ∈ [c, x], it follows that
hence Z x
m M
(x − c)n+1 ≤ (x − u)n g(u) du ≤ (x − c)n+1 ,
n+1 c n+1
126 7 Power series and Taylor series
and, therefore, Z x
n+1
m≤ (x − u)n g(u) du ≤ M. (7.6)
(x − c)n+1 c
Now let R denote the range of the restriction of g to the closed interval [c, x]. Since g
is continuous on [c, x] it follows from Theorem 4.24(i) that m, M ∈ R, and then (7.6)
and the intermediate value theorem (see Exercise E4.29) imply that
Z x
n+1
(x − u)n g(u) du
(x − c)n+1 c
Since
f (x) = P1 (x) + R1 (x), P1 (x) = f (c) + f 0 (c)(x − c)
by the definition of P1 and R1 , this verifies (7.4) for n = 1. Next,
Z x
f 00 (c) 1 x
Z
1 00
(x − u) f (u) du = 2
(x − c) + (x − u)2 f 000 (u) du
c 2 2 c
The first three terms on the right side are just P2 (x), and so the fourth term must
be R2 (x) and this verifies (7.4) for n = 2. Let’s play this game one more time:
Z x
f 000 (c) 1 x
Z
2 000
(x − u) f (x) du = + (x − u)3 f (4) du
c 3 3 c
7.4 Taylor series and Taylor polynomials 127
and hence
f 00 (c) f 000 (c)
f (x) = f (c) + f 0 (c)(x − c) + (x − c)2 + (x − c)3
2 3!
1 x
Z
+ (x − u)3 f (4) (x) du
3! c
1 x
Z
= P3 (x) + (x − u)3 f (4) (x) du.
3! c
Just as before, this verifies (7.4) for n = 3. Continuing in this manner (or, better
still, starting over and proceeding by induction) will lead to (7.4) for a general value
of n. In doing so, the assumption that f (n+1) is continuous on I is sufficient to justify
the last application of integration by parts in this calculation.
Next we verify (7.5). According to Lemma 1(ii), there is a point ξ ∈ I such that
Z x
(n+1) n+1
f (ξ) = (x − u)n f (n+1) du
(x − c)n+1 c
and, therefore,
1 (x − c)n+1 (n+1) f (n+1) (ξ)
Rn (x) = f (ξ) = (x − c)n+1
n! n + 1 (n + 1)!
by (7.4).
We can use either (7.4) or (7.5) in trying to decide whether the Taylor series of a
given function converges to that function.
7.13 Examples: 1. The derivatives of the sine function at the origin are easy to
calculate and the Taylor series of the sine function at 0 is
∞
X (−1)n 2n+1 x 3 x5 x7
x =x− + − + ··· .
n=0
(2n + 1)! 3! 5! 7!
The sine function is therefore equal to its Taylor series for all real numbers and
(from our knowledge of power series) it follows that
∞
X (−1)n 2n+1
sin x = x
n=0
(2n + 1)!
It is easy to verify (using the ratio test) that this power series has a radius of
convergence equal to ∞. To show that it actually converges to ex , it is sufficient
to prove that limn→∞ Rn (x) = 0 for each number x.
Now, for each number x and positive integer n, there is a number ξ such that
dn+1 ex xn+1 eξ
1 n+1
Rn (x) = x =
(n + 1)! dxn+1 x=ξ (n + 1)!
and, therefore,
|x|n+1 e|x|
0 ≤ Rn (x) ≤ .
(n + 1)!
|x|n
We saw in 1 that limn→∞ n!
= 0, hence limn→∞ Rn (x) = 0, and, therefore,
∞
x
X xn
e =
n=0
n!
The radius of convergence of this power series is best calculated by the ratio test:
r r n
n+1 |r − k|
lim
x x = |x| lim
n→∞ n+1 n n→∞ (k + 1)!
= |x|,
for |x| < 1, then we know that g is a C∞ -function defined on the interval (−1, 1).
To show that the Taylor series of f at 0 converges to f , it is sufficient to show
that f (x) = g(x) for |x| < 1, and for this it is, in turn, sufficient to show that
d g(x)
f (0) = g(0), = 0, x ∈ (−1, 1). (7.7)
dx f (x)
Indeed, the second of these equations implies that f (x)/g(x) is constant, and then
the first equation implies that this constant is 1.
Now, it is obvious from the definitions of f and g that f (0) = 1 = g(0) and the
quotient rule gives
g 0 (x)f (x) − g(x)f 0 (x)
d g(x)
=
dx f (x) f 2 (x)
(1 + x)r g 0 (x) − r(1 + x)r−1 g(x)
=
(1 + x)2r
(1 + x)g 0 (x) − rg(x)
= .
(1 + x)r+1
130 7 Power series and Taylor series
To complete this example, it will be sufficient to show that (1+x)g 0 (x) = rg(x) for
|x| < 1. Since g is a power series, its derivative g 0 can be found by term-by-term
differentiation (by Theorem 7.8), and hence
∞
0
X r n−1
(1 + x)g (x) = (1 + x) n x
n=1
n
∞ ∞
X r n−1 X r n
= n x + n x
n=1
n n=1
n
X ∞ ∞
r r n
X r n
= + (n + 1) x + n x
1 n=1
n+1 n=1
n
∞
X r r
=r+ (n + 1) +n xn .
n=1
n + 1 n
6. Although it is true for all of the common functions, it is not always the case
that the Taylor series of a C∞ -function converges to that function. The simplest
function for which this does not occur is probably the one in the following example.
The function f in this example satisfies f (0) = 0, 0 ≤ f (x) ≤ 1 for all x, is
asymptotic to 1 in both directions, and is incredibly flat at the origin. (Try using
plotting the graph of this function using some software package.)
Consider the function f defined by
( 2
e−1/x , x 6= 0,
f (x) =
0, x = 0.
2
Then f 0 (x) = 2x−3 e−1/x for x 6= 0 by the usual formulae for differentiation,
whereas f 0 (0) must be calculated directly from the definition of the derivative and
l’Hôspital’s rule:
2
0 f (h) − f (0) e−1/h
f (0) = lim = lim
h→0 h h→0 h
1/h k
= lim 1/h2 = lim k2
h→0 e k→±∞ e
1
= lim 2 = 0.
k→±∞ 2kek
Exercises
E7.1 Consider a power series ∞ n an
P
n=0 an (x − c) . Show that, if the sequence (| an+1 |)
converges or diverges to ∞, then the radius of convergence of the power series
is a
n
R = lim .
n→∞ an+1
an xn , where an is as
P
E7.2 Find the radius of convergence of the power series
given below, and determine whether the series converges absolutely, converges
conditionally, or diverges at the endpoints of the interval of convergence.
(a) an = 2−n n4 (n5 + 1)
(n!)2
(b) an =
(2n)!
(c) an = (3n + 4)n 5−n
E7.3 Determine the radius of convergence, R, of the power series ∞ n
P
n=1 an x , where
an is as given below. Also, determine whether the series converges absolutely,
converges conditionally, or diverges at the points x = ±R.
(a) an = 10−n + 10n
(n!)2
(b) an = 1/ 2n n
= (2n)! .
E7.4 Determine the radius of convergence of each of the following power series:
P∞ n10 n
(a) x
Pn=1
∞
n!
n! n
(b) n=1 nn (x − 1)
E7.5 Determine the interval of convergence of each of the following series. (Do not
investigate convergence or divergence at the endpoints.)
∞ n
bn
X a
(a) + 2 (x − 1)2n , where a and b are two positive numbers
n=1
n n
∞
X (2x + 3)n
(b)
n=1
n
∞
X en x n
(c) − 1
n=1
n3 + 1 2
E7.6 Determine the interval of convergence of the power series
∞
X (2x)2n
.
n=1
n log n
E7.7 Determine the interval of convergence of and an explicit formula for the power
series
f (x) = 1 + 2x + x2 + 2x3 + x4 + · · · .
Exercises for Chapter 7 133
is R = ∞, so that the series converges absolutely for all x ∈ R (you don’t need
to show this). Let Jr (x) denote the function represented by this series. This
function is called the r-th order Bessel function of the first kind.
Verify that Jr (x) is a solution y(x) to Bessel’s differential equation
x2 y 00 + xy 0 + (x2 − r2 )y = 0.
You must justify all the steps in your verification by quoting appropriate the-
orems wherever necessary.
(Remark : Bessel functions have many applications; they arise for instance
when studying the propagation of electromagnetic waves through cylindrical
waveguides.)
E7.17 It is shown in Example 7.13–4 of the course reader (read it!) that the Taylor
series about 0 of f (x) = (1 + x)r , where r is a real number, is given by
P ∞ r
n
n=0 n x . Since r can be any real number, not necessarily a positive integer,
the binomial coefficient n is to be interpreted as r(r−1)(r−2)...(r−n+1)
r
n!
, with the
r
understanding that 0 = 1. It is also shown in the example that the Taylor
series, in fact, converges to f (x) for all x ∈ (−1, 1).
(a) Use the above fact (and some algebraic manipulations)
P∞ 2n nto deduce that the
1
Taylor series about 0 of 1−4x is the series n=0 n x .
√
P∞ 2n n 1
Conclude from this that n=0 n x converges to √1−4x for x ∈
(−1/4, 1/4).
Exercises for Chapter 7 135
(b) (This is hard, and need not be attempted, but you may wish P∞to give itn
2n
a shot.) A solution to Exercise E5.13 shows that the series n=0 n x
converges conditionally at x = −1/4. Show that
∞
X 2n 1 1
(−1/4)n = p =√ .
n=0
n 1 − 4(−1/4) 2
1 2n
E7.18 The nth Catalan number is defined as Cn = n+1 n
, for all integers n ≥ 0.
(a) Use the inequality
2n −n 1 1
4 ≤ √ e 24n
n πn
P∞
to deduce that n=0 Cn 4−n converges.
(b) Determine the interval of convergence, I, of the power series ∞ n
P
P∞ n=0 Cn x .
(c) Show that n=0 Cn xn converges uniformly on I.
(d) Deduce from part (c) that the function f (x) = ∞ n
P
n=0 Cn x is continuous
on I.
(e) Use the result of Exercise E7.17(a) to derive an expression for ∞ n
P
n=0 Cn x
that is valid for all x ∈ I. You must carefully justify all the steps in your
derivation.
(Catalan numbers occur in various counting problems; the question above
derives a generating function for the Catalan numbers.)
series ∞ n
P
E7.19 If the radius of convergence of the powerP n=0 an x is R, what is the
∞ 2n
radius of convergence of the power series n=0 an x ?
Suppose that the power series ∞ an xn has radius of convergence R ∈ (0, ∞).
P
E7.20 n=0 P∞ 2
Find the radius of convergence of n=0 an xn .
X∞
E7.21 Suppose that the power series f (x) = an xn has a nonzero radius of con-
n=0
vergence.
(a) If f is an odd function, show that
a0 = a2 = a4 = · · · = 0.
a1 = a3 = a5 = · · · = 0.
(a) Determine the radius of convergence of this series and the interval I con-
sisting of all the numbers x for which the series converges.
(b) Show that it converges uniformly on the set I.
Exercises for Chapter 7 137
(cf. the proof of Lemma 1(ii) from the proof of Theorem 7.12).
E7.36 The binomial theorem states that
p(p − 1) 2 p(p − 1)(p − 2) 3
(1 + x)p = 1 + px + x + x + ···
2! 3!
for |x| ≤ 1. Use the familiar formula
Z x
−1 du
sin x= √
0 1 − u2
to deduce that
π 1 1 1·3
= + 3
+ + ··· .
6 2 2·3·2 2 · 4 · 5 · 25
Justify each of the steps in your work.
E7.37 Use the binomial theorem to obtain simple approximations for the following
expressions valid for arguments indicated:
1
(a) for |u| 1
1−u
−1/2
v2 v2
(b) 1 − 2 for 2 1
c c
2 3/2 2
(c) (1 + x ) for x 1
(d) (1 + 3x)π for |x| 1
E7.38 Compute the first few terms in the power series expansion about x = 0 of the
following functions:
(a) x3 cos x
e−x
(b) (1−x)
(c) ex sin x
(d) tan x
Index
absolute convergence, 80, 116 construction of the reals, 17, 19
alternating harmonic series, 78, 79, 115 continuity, 5, 58
alternating series test, 81 continuous, 58
remainder, 83 contraction, 47
contradiction, 6
balls, 50 contrapositive, 6
closed, 50 convergence
open, 50 absolute, 80, 116
punctured, 50 conditional, 82
binomial coefficient, 129 of power series, 115
Bolzano-Weierstrass Theorem pointwise, 92
for sequences, 29 uniform, 116
for sets, 54 for sequences, 96
boundary, 56 convergence tests
boundary point, 56 absolute convergence, 80
bounded, 103 alternating series test, 81
above, 20, 21 Cauchy criterion, 105
below, 20, 21 comparison test, 72
sequence, 26, 28, 29, 46 integral test, 74, 80
set, 20, 21 p-series, 75
ratio test, 75, 80, 114
Cauchy root test, 76, 81
pointwise, 105 convergent sequence, 26, 28, 45
Cauchy criterion Cauchy, 36, 47
for sequences, 36, 47 uniform, 96
Cauchy sequence, 36, 47 uniform Cauchy, 105
Cauchy–Schwartz inequality, 45 convergent sequences, 30
closed and bounded set, 57 convergent series, 69, 80
closed ball, 50 absolute, 80, 116
closed set, 51, 56 conditional, 82
cluster point, 54–56 uniform, 116
compact set, 59 convergent subsequence, 29, 57
comparison test, 72 converse, 5
complement, 2 cosine function
complement of a set, 51 Taylor series of, 128
completeness, 19 counterexample, 9
conditional convergence, 82
conjunction, 3 de Morgan’s laws, 53
140 Index