(Larson) Introduction To Real Analysis
(Larson) Introduction To Real Analysis
Lee Larson
University of Louisville
May 1, 2015
About This Document
I often teach the MATH 501-502: Introduction to Real Analysis course at the
University of Louisville. These are notes Ive compiled from those experiences. They
cover the basic ideas of analysis on the real line. The course is intended for a mix of
mostly senior mathematics majors and some beginning graduate students.
Some early versions of the notes leaked out onto the Internet and they are being
offered by some of the usual download sites. The early versions were meant for
me and my classes only, and contain many typos and some gasp! outright
mistakes. The notes are updated and corrected quite often. The date of the version
youre reading is at the bottom-left of most pages. The latest version is available
for download at the Web address math.louisville.edu/lee/ira.
Feel free to use these notes for any purpose, as long as you give me blame or
credit. In return, I only ask you to tell me about mistakes. Any suggestions for
improvements and additions are appreciated. I can be contacted using the email
address on the Web page referenced above.
May 1, 2015
i
Contents
Basic Ideas
In the end, all mathematics can be boiled down to logic and set theory. Because
of this, any careful presentation of fundamental mathematical ideas is inevitably
couched in the language of logic and sets. This chapter defines enough of that
language to allow the presentation of basic real analysis. Much of it will be familiar
to you, but look at it anyway to make sure you understand the notation.
1. Sets
Set theory is a large and complicated subject in its own right. There is no time
in this course to touch on any but the simplest parts of it. Instead, well just look
at a few topics from what is often called naive set theory.
We begin with a few definitions.
A set is a collection of objects called elements. Usually, sets are denoted by the
capital letters A, B, , Z. A set can consist of any type and number of elements.
Even other sets can be elements of a set. The sets dealt with here usually have real
numbers as their elements.
If a is an element of the set A, we write a A. If a is not an element of the set
A, we write a / A.
If all the elements of A are also elements of B, then A is a subset of B. In this
case, we write A B or B A. In particular, notice that whenever A is a set, then
A A.
Two sets A and B are equal, if they have the same elements. In this case we
write A = B. It is easy to see that A = B iff A B and B A. Establishing that
both of these containments are true is the most common way to show that two sets
are equal.
If A B and A 6= B, then A is a proper subset of B. In cases when this is
important, it is written A ( B instead of just A B.
There are several ways to describe a set.
A set can be described in words such as P is the set of all presidents of the
United States. This is cumbersome for complicated sets.
All the elements of the set could be listed in curly braces as S = {2, 0, a}. If
the set has many elements, this is impractical, or impossible.
More common in mathematics is set builder notation. Some examples are
and
A = {n : n is a prime number} = {2, 3, 5, 7, 11, }.
1-1
1-2 CHAPTER 1. BASIC IDEAS
Definition 1.1. Given any set A, the power set of A, written P(A), is the set
consisting of all subsets of A; i. e.,
P(A) = {B : B A}.
For example, P({a, b}) = {, {a}, {b}, {a, b}}. Also, for any set A, it is always
true that P(A) and A P(A). If a A, it is never true that a P(A), but it
is true that {a} P(A). Make sure you understand why!
An amusing example is P() = {}. (Dont confuse with {}! The former is
empty and the latter has one element.) Now, consider
P() = {}
P(P()) = {, {}}
P(P(P())) = {, {}, {{}}, {, {}}}
After continuing this n times, for some n N, the resulting set,
P(P( P() )),
is very large. In fact, since a set with k elements has 2k elements in its power set,
22
there are22 = 65, 536 elements after only five iterations of the example. Beyond
this, the numbers are too large to print. Number sequences such as this one are
sometimes called tetrations.
2. Algebra of Sets
Let A and B be sets. There are four common binary operations used on sets.1
The union of A and B is the set containing all the elements in either A or B:
A B = {x : x A x B}.
The intersection of A and B is the set containing the elements contained in
both A and B:
A B = {x : x A x B}.
1In the following, some logical notation is used. The symbol is the logical nonexclusive or.
The symbol is the logical and. Their truth tables are as follows:
T F T F
T T F T T T
F F F F T F
A B A B
A B A B
A\B AB
A B A B
Figure 1. These are Venn diagrams showing the four standard binary
operations on sets. In this figure, the set which results from the operation
is shaded.
A \ B = {x : x A x
/ B}.
The symmetric difference of A and B is the set of elements in one of the sets,
but not the other:
AB = (A B) \ (A B).
(a) A \ (B C) = (A \ B) (A \ C)
(b) A \ (B C) = (A \ B) (A \ C)
3. Indexed Sets
We often have occasion to work with large collections of sets. For example, we
could have a sequence of sets A1 , A2 , A3 , , where there is a set An associated
with each n N. In general, let be a set and suppose for each there is a set
A . The set {A : } is called a collection of sets indexed by . In this case,
is called the indexing set for the collection.
Example 1.1. For each n N, let An = {k Z : k 2 n}. Then
A1 = A2 =A3 = {1, 0, 1}, A4 = {2, 1, 0, 1, 2}, ,
A50 = {7, 6, 5, 4, 3, 2, 1, 0, 1, 2, 3, 4, 5, 6, 7},
is a collection of sets indexed by N.
Two of the basic binary operations can be extended to work with indexed
collections. In particular, using the indexed collection from the previous paragraph,
we define [
A = {x : x A for some }
and \
A = {x : x A for all }.
DeMorgans Laws can be generalized to indexed collections.
2The logical symbol is the same as if, and only if. If A and B are any two statements,
then A B is the same as saying A implies B and B implies A. It is also common to use iff
in this way.
4.2. Relations.
Definition 1.6. If A and B are sets, then any R A B is a relation from A
to B. If (a, b) R, we write aRb.
In this case,
dom (R) = {a : (a, b) R}
is the domain of R and
ran (R) = {b : (a, b) R}
is the range of R.
In the special case when R A A, for some set A, there is some additional
terminology.
R is symmetric, if aRb bRa.
R is reflexive, if aRa whenever a dom (A).
R is transitive, if aRb bRc = aRc.
R is an equivalence relation on A, if it is symmetric, reflexive and transitive.
Example 1.3. Let R be the relation on Z Z defined by aRb a b.
Then R is reflexive and transitive, but not symmetric.
Example 1.4. Let R be the relation on Z Z defined by aRb a < b.
Then R is transitive, but neither reflexive nor symmetric.
Example 1.5. Let R be the relation on Z Z defined by aRb a2 = b2 .
In this case, R is an equivalence relation. It is evident that aRb iff b = a or b = a.
4.3. Functions.
Definition 1.7. A relation R A B is a function if
aRb1 aRb2 = b1 = b2 .
If f A B is a function and dom (f ) = A, then we usually write f : A B
and use the usual notation f (a) = b instead of af b.
If f : A B is a function, the usual intuitive interpretation is to regard f
as a rule that associates each element of A with a unique element of B. Its not
necessarily the case that each element of B is associated with something from A;
i.e., B may not be ran (f ).
Example 1.6. Define f : N Z by f (n) = n2 and g : Z Z by g(n) = n2 .
In this case ran (f ) = {n2 : n N} and ran (g) = ran (f ) {0}. Notice that even
though f and g use the same formula, they are actually different functions.
b
c
a f
A B
b
c
a g
A B
g
f
1
f
f 1
A f B
A1 A2 A3 A4 A5
B1 B2 B3 B4
f (A)
B5
B
Figure 4. Here are the first few steps from the construction used in
the proof of Theorem 1.16.
This implies
h(x) = g 1 (x) = f g f f g (x1 ) = f (y)
| {z }
k 1 f s and k 1 gs
3
This is often called the Cantor-Schroder-Bernstein or Cantor-Bernstein Theorem, despite the
fact that it was apparently first proved by Richard Dedekind.
so that
y = g f g f f g (x1 ) Ak1 A.
| {z }
k 2 f s and k 1 gs
This contradiction shows that h(x) 6= h(y). We conclude h is injective.
To show h is surjective, let y B. If y Bk for some k, then h(Ak ) = g 1 (Ak ) =
Bk shows y h(A). If y / Bk for any k, y f (A) because B1 = B \ f (A), and
g(y)
/ A, so y = h(x) = f (x) for some x A. This shows h is surjective.
The Schroder-Bernstein theorem has many consequences, some of which are at
first a bit unintuitive, such as the following theorem.
Corollary 1.17. There is a bijective function h : N N N
Proof. If f : N N N is f (n) = (n, 1), then f is clearly injective. On the
other hand, suppose g : N N N is defined by g((a, b)) = 2a 3b . The uniqueness
of prime factorizations guarantees g is injective. An application of Theorem 1.16
yields h.
To appreciate the power of the Schroder-Bernstein theorem, try to find an
explicit bijection h : N N N.
5. Cardinality
There is a way to use sets and functions to formalize and generalize how we
count. For example, suppose we want to count how many elements are in the set
{a, b, c}. The natural way to do this is to point at each element in succession and
say one, two, three. What is really happening is that were defining a bijective
function between {a, b, c} and the set {1, 2, 3}. This idea can be generalized.
Definition 1.18. Given n N, the set n = {1, 2, , n} is called an initial
segment of N. The trivial initial segment is 0 = . A set S has cardinality n, if
there is a bijective function f : S n. In this case, we write card (S) = n.
The cardinalities defined in Definition 1.18 are called the finite cardinal numbers.
They correspond to the everyday counting numbers we usually use. The idea can
be generalized still further.
Definition 1.19. Let A and B be two sets. If there is an injective function
f : A B, we say card (A) card (B).
According to Theorem 1.16, the Schroder-Bernstein Theorem, if card (A)
card (B) and card (B) card (A), then there is a bijective function f : A B. As
expected, in this case we write card (A) = card (B). When card (A) card (B), but
no such bijection exists, we write card (A) < card (B). Theorem 1.15 shows that
card (A) = card (B) is an equivalence relation between sets.
The idea here, of course, is that card (A) = card (B) means A and B have the
same number of elements and card (A) < card (B) means A is a smaller set than B.
This simple intuitive understanding has some surprising consequences when the sets
involved do not have finite cardinality.
In particular, the set A is countably infinite, if card (A) = card (N). In this case,
it is common to write card (N) = 0 .4 When card (A) 0 , then A is said to be a
4The symbol is the Hebrew letter aleph and is usually pronounced aleph nought.
0
countable set. In other words, the countable sets are those having finite or countably
infinite cardinality.
Example 1.9. Let f : N Z be defined as
(
n+1
f (n) = 2 , when n is odd
.
n
1 2 , when n is even
Its easy to show f is a bijection, so card (N) = card (Z) = 0 .
Theorem 1.20. Suppose A and B are countable sets.
(a) A B is countable.
(b) A B is countable.
Proof. (a) This is a consequence of Theorem 1.17.
(b) This is Exercise 1.20.
An alert reader will have noticed from previous examples that
0 = card (Z) = card () = card (N) = card (N N) = card (N N N) =
A logical question is whether all sets either have finite cardinality, or are countably
infinite. That this is not so is seen by letting S = N in the following theorem.
Theorem 1.21. If S is a set, card (S) < card (P(S)).
Proof. Noting that 0 = card () < 1 = card (P()), the theorem is true when
S is empty.
Suppose S 6= . Since {a} P(S) for all a S, it follows that card (S)
card (P(S)). Therefore, it suffices to prove there is no surjective function f : S
P(S).
To see this, assume there is such a function f and let T = {x S : x / f (x)}.
Since f is surjective, there is a t S such that f (t) = T . Either t T or t / T.
If t T = f (t), then the definition of T implies t / T , a contradiction. On
the other hand, if t / T = f (t), then the definition of T implies t T , another
contradiction. These contradictions lead to the conclusion that no such function f
can exist.
A set S is said to be uncountably infinite, or just uncountable, if 0 < card (S).
Theorem 1.21 implies 0 < card (P(N)), so P(N) is uncountable. In fact, the same
argument implies
0 = card (N) < card (P(N)) < card (P(P(N))) <
So, there are an infinite number of distinct infinite cardinalities.
In 1874 Georg Cantor [5] proved card (R) = card (P(N)) and card (R) > 0 ,
where R is the set of real numbers. (A version of Cantors theorem appears in
Theorem 2.25 below.) This naturally led to the question whether there are sets S
such that 0 < card (S) < card (R). Cantor spent many years trying answer this
question and never succeeded. His assumption that no such sets exist came to be
called the continuum hypothesis.
The importance of the continuum hypothesis was highlighted by David Hilbert
at the 1900 International Congress of Mathematicians in Paris, when he put it first
on his famous list of the 23 most important open problems in mathematics. Kurt
Godel proved in 1940 that the continuum hypothesis cannot be disproved using
standard set theory, but he did not prove it was true. In 1963 it was proved by
Paul Cohen that the continuum hypothesis is actually unprovable as a theorem in
standard set theory.
So, the continuum hypothesis is a statement with the strange property that it
is neither true nor false within the framework of ordinary set theory. This means
that in the standard axiomatic development of set theory, the continuum hypothesis,
or a suitable negation of it, can be taken as an additional axiom without causing
any contradictions. The technical terminology is that the continuum hypothesis is
independent of the axioms of set theory.
The proofs of these theorems are extremely difficult and entire broad areas of
mathematics were invented just to make their proofs possible. Even today, there
are some deep philosophical questions swirling around them. A more technical
introduction to many of these ideas is contained in the book by Ciesielski [8].
A nontechnical and very readable history of the efforts by mathematicians to
understand the continuum hypothesis is the book by Aczel [1].
6. Exercises
1.1. If a set S has n elements for n , then how many elements are in P(S)?
1.13. Let S be a set. Prove the following two statements are equivalent:
(a) S is infinite; and,
(b) there is a proper subset T of S and a bijection f : S T .
This statement is often used as the definition of when a set is infinite.
1.14. If S is an infinite set, then there is a countably infinite
S collection of nonempty
pairwise disjoint infinite sets Tn , n N such that S = nN Tn .
1.16. If f : [0, ) (0, ) and g : (0, ) [0, ) are given by f (x) = x + 1 and
g(x) = x, then the proof of the Schroder-Bernstein theorem yields what bijection
h : [0, ) (0, )?
1.20. If A and B are sets such that card (A) = card (B) = 0 , then card (A B) =
0 .
1.21. Using the notation from the proof of the Schroder-Bernstein Theorem, let
A = [0, ), B = (0, ), f (x) = x + 1 and g(x) = x. Determine h(x).
1.22. Using the notation from the proof of the Schroder-Bernstein Theorem, let
A = N, B = Z, f (n) = n and
(
1 3n, n 0
g(n) = .
3n 1, n > 0
Calculate h(6) and h(7).
S
1.24. If {An : n N} is a collection of countable sets, then nN An is countable.
1.25. If A and B are sets, the set ofall functions f : A B is often denoted by
S
B A . If S is a set, prove that card 2 = card (P(S)).
This chapter concerns what can be thought of as the rules of the game: the
axioms of the real numbers. These axioms imply all the properties of the real
numbers and, in a sense, any set satisfying them is uniquely determined to be the
real numbers.
The axioms are presented here as rules without very much justification. Other
approaches can be used. For example, a common approach is to begin with the
Peano axioms the axioms of the natural numbers and build up to the real
numbers through several completions of the natural numbers. Its also possible to
begin with the axioms of set theory to build up the Peano axioms as theorems and
then use those to prove our axioms as further theorems. No matter how its done,
there are always some axioms at the base of the structure and the rules for the real
numbers are the same, whether theyre axioms or theorems.
We choose to start at the top because the other approaches quickly turn into
a long and tedious labyrinth of technical exercises without much connection to
analysis.
2-1
2-2 CHAPTER 2. THE REAL NUMBERS
There are many other properties of fields which could be proved here, but they
correspond to the usual properties of the real numbers learned in beginning algebra,
so we omit them. Some of them are in the exercises at the end of this chapter.
From now on, the standard notations for algebra will usually be used; e. g., we
will allow ab instead of a b and a/b instead of a b1 . The reader may also use
the standard facts she learned from elementary algebra.
A metric is a function which defines the distance between any two points of a
set.
Example 2.3. Let S be a set and define d : S S S by
(
1, x 6= y
d(x, y) = .
0, x = y
It can readily be verified that d is a metric on S. This simplest of all metrics is
called the discrete metric and it can be defined on any set. Its not often useful.
Theorem 2.11. If F is an ordered field, then d(x, y) = |x y| is a metric on F.
Proof. Use parts (a), (b) and (e) of Theorem 2.9.
The metric on F derived from the absolute value function is called the standard
metric on F. There are other metrics sometimes defined for specialized purposes,
but we wont have need of them.
2Some people prefer the notation sup A and inf A instead of lub A and glb A, respectively.
They stand for the supremum and infimum of A.
Proof. If the theorem is false, then a is an upper bound for N. Let = lub N.
According to Theorem 2.17 there is an m N such that m > 1. But, this is a
contradiction because = lub N < m + 1 N.
Some other variations on this theme are in the following corollaries.
Corollary 2.21. Let a, b R with a > 0.
(a) There is an n N such that an > b.
(b) There is an n N such that 0 < 1/n < a.
(c) There is an n N such that n 1 a < n.
Proof. (a) Use Theorem 2.20 to find n N where n > b/a.
(b) Let b = 1 in part (a).
(c) Theorem 2.20 guarantees that S = {n N : n > a} 6= . If n is the least
element of this set, then n 1
/ S and n 1 a < n.
Corollary 2.22. If I is any interval from R, then I Q 6= and I Qc 6= .
Proof. Left as an exercise.
A subset of R which intersects every interval is said to be dense in R. Corollary
2.22 shows both the rational and irrational numbers are dense.
4. Comparisons of Q and R
All of the above still does not establish that Q is different from R. In Theorem
2.12, it was shown that the equation x2 = 2 has no solution in Q. The following
theorem shows x2 = 2 does have solutions in R. Since a copy of Q is embedded in
R, it follows, in a sense, that R is bigger than Q.
Theorem 2.23. There is a positive R such that 2 = 2.
Proof. Let S = {x > 0 : x2 < 2}. Then 1 S, so S 6= . If x 2, then
Theorem 2.5(c) implies x2 4 > 2, so S is bounded above. Let = lub S. It will
be shown that 2 = 2.
Suppose first that 2 < 2. This assumption implies (2 2 )/(2 + 1) > 0.
According to Corollary 2.21, there is an n N large enough so that
1 2 2 2 + 1
0< < = 0 < < 2 2 .
n 2 + 1 n
Therefore,
2
1 2 1 1 1
+ = 2 +
+ 2 = 2 + 2 +
n n n n n
(2 + 1)
< 2 + < 2 + (2 2 ) = 2
n
contradicts the fact that = lub S. Therefore, 2 2.
Next, assume 2 > 2. In this case, choose n N so that
1 2 2 2
0< < = 0 < < 2 2.
n 2 n
Then
2
1 2 1 2
= 2 + 2 > 2 > 2 (2 2) = 2,
n n n n
Proof. It suffices to prove that card ([0, 1]) > 0 . If this is not true, then there
is a bijection : N [0, 1]; i.e.,
(5) [0, 1] = {n : n N}.
P
Each x [0, 1] can be written in the decimal form x = n=1 x(n)/10n where
x(n) {0, 1, 2, 3, 4, 5, 6, 7, 8, 9} for each n N. This decimal representation is not
necessarily unique. For example,
1 5 4 X 9
= = + .
2 10 10 n=2 10n
In such a case, there is a choice of x(n) so it is constantly 9 or constantly 0 from
some N onward. When given a choice, we will always opt to end the number with a
string of nines. With this convention, the decimal representation of x is unique.
Define
Pz [0, 1] by choosing z(n) {d : d 8} such that z(n) 6= n (n).
Let z = n=1 z(n)/10n . Since z [0, 1], there is an n N such that z = (n).
But, this is impossible because z(n) differs from n in the nth decimal place. This
contradiction shows card ([0, 1]) > 0 .
Around the turn of the twentieth century these then-new ideas about infinite
sets were very controversial in mathematics. This is because some of these ideas are
very unintuitive. For example, the rational numbers are a countable set and the
irrational numbers are uncountable, yet between every two rational numbers are an
uncountable number of irrational numbers and between every two irrational numbers
there are a countably infinite number of rational numbers. It would seem there are
either too few or too many gaps in the sets to make this possible. Such a seemingly
paradoxical situation flies in the face of our intuition, which was developed with
finite sets in mind.
This brings us back to the discussion of cardinalities and the Continuum
Hypothesis at the end of Section 5. Most of the time, people working in real
analysis assume the Continuum Hypothesis is true. With this assumption and
Theorem 2.25 it follows that whenever A R, then either card (A) 0 or
card (A) = card (R) = card (P(N)).3 Since P(N) has many more elements than N,
any countable subset of R is considered to be a small set, in the sense of cardinality,
even if it is infinite. This works against the intuition of many beginning students
who are not used to thinking of Q, or any other infinite set as being small. But it
turns out to be quite useful because the fact that the union of a countably infinite
number of countable sets is still countable can be exploited in many ways.4
In later chapters, other useful small versus large dichotomies will be found.
5. Exercises
2.9. If F is an ordered field and a F such that 0 a < for every > 0, then
a = 0.
2.10. Let x R. Prove |x| < for all > 0 iff x = 0.
2.12. If p is a prime number and > 0, then there are x, y Q such that
x2 < p < y 2 < x2 + .
3Since is the smallest infinite cardinal, is used to denote the smallest uncountable
0 1
cardinal. You will also see card (R) = c, where c is the old-style German letter c, standing for the
cardinality of the continuum. Assuming the continuum hypothesis, it follows that 0 < 1 = c.
4See Problem 24 on page 1-13.
2.20. If A R and B = {x : x is an upper bound for A}, then lub (A) = glb (B).
Sequences
1. Basic Properties
Definition 3.1. A sequence is a function a : N R.
Instead of using the standard function notation of a(n) for sequences, it is
usually more convenient to write the argument of the function as a subscript, an .
Example 3.1. Let the sequence an = 1 1/n. The first three elements are
a1 = 0, a2 = 1/2, a3 = 2/3, etc.
Example 3.2. Let the sequence bn = 2n . Then b1 = 2, b2 = 4, b3 = 8, etc.
The converse of this theorem is not true. For example, an = (1)n is bounded,
but does not converge. The main use of Corollary 3.7 is as a quick first check to see
whether a sequence might converge. Its usually pretty easy to determine whether a
sequence is bounded. If it isnt, it must diverge.
The following theorem lets us analyze some complicated sequences by breaking
them down into combinations of simpler sequences.
Theorem 3.8. Let an and bn be sequences such that an A and bn B.
Then
(a) an + bn A + B,
(b) an bn AB, and
(c) an /bn A/B as long as bn 6= 0 for all n N and B 6= 0.
Proof. (a) Let > 0. There are N1 , N2 N such that n N1 im-
plies |an A| < /2 and n N2 implies |bn B| < /2. Define
N = max{N1 , N2 }. If n N , then
|(an + bn ) (A + B)| |an A| + |bn B| < /2 + /2 = .
Therefore an + bn A + B.
(b) Let > 0 and > 0 be an upper bound for |an |. Choose N1 , N2 N such
that n N1 = |an A| < /2(|B|+1) and n N2 = |bn B| < /2.
If n N = max{N1 , N2 }, then
|an bn AB| = |an bn an B + an B AB|
|an bn an B| + |an B AB|
= |an ||bn B| + ||B||an A|
< + |B|
2 2(|B| + 1)
< /2 + /2 = .
(c) First, notice that it suffices to show that 1/bn 1/B, because part (b) of
this theorem can be used to achieve the full result.
Let > 0. Choose N N so that the following two conditions are
satisfied: n N = |bn | > |B|/2 and |bn B| < B 2 /2. Then, when
n N,
2
1 = B bn < B /2 = .
1
bn B bn B (B/2)B
Therefore 1/bn 1/B.
If youre not careful, you can easily read too much into the previous theorem
and try to use its converse. Consider the sequences an = (1)n and bn = an .
Their sum, an + bn = 0, product an bn = 1 and quotient an /bn = 1 all converge,
but the original sequences diverge.
2. Monotone Sequences
One of the problems with using the definition of convergence to prove a given
sequence converges is the limit of the sequence must be known in order to verify the
sequence converges. This gives rise in the best cases to a chicken and egg problem
of somehow determining the limit before you even know the sequence converges. In
the worst case, there is no nice representation of the limit to use, so you dont even
have a target to shoot at. The next few sections are ultimately concerned with
removing this deficiency from Definition 3.2, but some interesting side-issues are
explored along the way.
Not surprisingly, we begin with the simplest case.
Definition 3.10. A sequence an is increasing, if an+1 an for all n N. It is
strictly increasing if an+1 > an for all n N.
A sequence an is decreasing, if an+1 an for all n N. It is strictly decreasing
if an+1 < an for all n N.
If an is any of the four types listed above, then it is said to be a monotone
sequence.
Notice the and in the definitions of increasing and decreasing sequences,
respectively. Many calculus texts use strict inequalities because they seem to better
match the intuitive idea of what an increasing or decreasing sequence should do.
For us, the non-strict inequalities are more convenient.
Theorem 3.11. A bounded monotone sequence converges.
Proof. Suppose an is a bounded increasing sequence, L = lub {an : n N}
and > 0. Clearly, an L for all n N. According to Theorem 2.17, there
exists an N N such that aN > L . Because the sequence is increasing,
L an aN > L for all n N . This shows an L.
If an is decreasing, let bn = an and apply the preceding argument.
The key idea of this proof is the existence of the least upper bound of the sequence
when viewed as a set. This means the Completeness Axiom implies Theorem 3.11.
In fact, it isnt hard to prove Theorem 3.11 also implies the Completeness Axiom,
showing they are equivalent statements. Because of this, Theorem 3.11 is often used
as the Completeness Axiom on R instead of the least upper bound property we used
in Axiom 8.
n
Example 3.9. The sequence en = 1 + n1 converges.
and
n+1
X
n+1 1
(10) en+1 = .
k (n + 1)k
k=0
The main use of Theorem 3.14 is not to show that sequences converge, but, rather
to show they diverge. It gives two strategies for doing this: find two subsequences
converging to different limits, or find a divergent subsequence. In Example 3.11, the
subsequences bn and cn demonstrate the first strategy, while dn demonstrates the
second.
Even if the original sequence diverges, it is possible there are convergent subse-
quences. For example, consider the divergent sequence an = (1)n . In this case, an
diverges, but the two subsequences a2n = 1 and a2n+1 = 1 are constant sequences,
so they converge.
Theorem 3.15. Every sequence has a monotone subsequence.
Proof. Let an be a sequence and T = {n N : m > n = am an }. There
are two cases to consider, depending on whether T is finite.
First, assume T is infinite. Define (1) = min T and assuming (n) is defined,
set (n + 1) = min T \ {(1), (2), . . . , (n)}. This inductively defines a strictly
increasing function : N N. The definition of T guarantees a(n) is an increasing
subsequence of an .
Now, assume T is finite. Let (1) = max T + 1. If (n) has been chosen for
some n > max T , then the definition of T implies there is an m > (n) such that
am a(n) . Set (n + 1) = m. This inductively defines the strictly increasing
function : N N such that a(n) is a decreasing subsequence of an .
If the sequence in Theorem 3.15 is bounded, then the corresponding monotone
subsequence is also bounded. Recalling Theorem 3.11, we see the following.
Theorem 3.16 (Bolzano-Weierstrass). Every bounded sequence has a convergent
subsequence.
6. Cauchy Sequences
Often the biggest problem with showing that a sequence converges using the
techniques we have seen so far is that we must know ahead of time to what it
converges. This is the chicken and egg problem mentioned above. An escape from
this dilemma is provided by Cauchy sequences.
Definition 3.21. A sequence an is a Cauchy sequence if for all > 0 there is
an N N such that n, m N implies |an am | < .
1May 1, 2015
Lee
c Larson ([email protected])
This definition is a bit more subtle than it might at first appear. It sort of says
that all the terms of the sequence are close together from some point onward. The
emphasis is on all the terms from some point onward. To stress this, first consider
a negative example.
Pn
Example 3.16. Suppose an = k=1 1/k for n N. Theres a trick for showing
the sequence an diverges. First, note that an is strictly increasing. For any n N,
consider
n j
2X 1 n1
X 2X1
1 1
a2n 1 = = j +k
k j=0
2
k=1 k=0
j
n1
X 2X 1 n1
1 X 1 n
> = =
j=0 k=0
2j+1 j=0
2 2
Hence, the subsequence a2n 1 is unbounded and the sequence an diverges. (To see
how this works, write out the first few sums of the form a2n 1 .)
On the other hand, |an+1 an | = 1/(n + 1) 0 and indeed, if m is fixed,
|an+m sn | 0. This makes it seem as though the terms are getting close together,
as in the definition of a Cauchy sequence. But, sn is not a Cauchy sequence, as
shown by the following theorem.
Theorem 3.22. A sequence converges iff it is a Cauchy sequence.
Proof. () Suppose an L and > 0. There is an N N such that n N
implies |an L| < /2. If m, n N , then
|am an | = |am L + L an | |am L| + |L am | < /2 + /2 = .
This shows an is a Cauchy sequence.
() Let an be a Cauchy sequence. First, we claim that an is bounded. To see
this, let = 1 and choose N N such that n, m N implies |an am | < 1. In this
case, aN 1 < an < aN + 1 for all n N , so {an : n N } is a bounded set. The
set {an : n < N }, being finite, is also bounded. Since {an : n N} is the union of
these two bounded sets, it too must be bounded.
Because an is a bounded sequence, Theorem 3.16 implies it has a convergent
subsequence bn = a(n) L. Let > 0 and choose N N so that n, m N implies
|an am | < /2 and |bn L| < /2. If n N , then (n) n N and
|an L| = |an bn + bn L|
|an bn | + |bn L|
= |an a(n) | + |bn L|
< /2 + /2 = .
Therefore, an L.
The fact that Cauchy sequences converge is yet another equivalent version
of completeness. In fact, most advanced texts define completeness as Cauchy
sequences converge. This is convenient in general spaces because the definition of a
Cauchy sequence only needs the metric on the space and none of its other structure.
A typical example of the usefulness of Cauchy sequences is given below.
Definition 3.23. A sequence xn is contractive if there is a c (0, 1) such that
|xk+1 xk | c|xk xk1 | for all k > 1. c is called the contraction constant.
This shows the claim is true in the case n + 1. Therefore, by induction, the claim is
true for all n N.
To show xn is a Cauchy sequence, let > 0. Since cn 0, we can choose N N
so that
cN 1
(14) |x1 x2 | < .
(1 c)
Let n > m N . Then
|xn xm | = |xn xn1 + xn1 xn2 + xn2 xm+1 + xm+1 xm |
|xn xn1 | + |xn1 xn2 | + + |xm+1 xm |
Now, use (13) on each of these terms.
cn2 |x1 x2 | + cn3 |x1 x2 | + + cm1 |x1 x2 |
= |x1 x2 |(cn2 + cn3 + + cm1 )
Apply the formula for a geometric sum.
1 cnm
= |x1 x2 |cm1
1c
cm1
(15) < |x1 x2 |
1c
Use (14) to estimate the following.
cN 1
|x1 x2 |
1c
< |x1 x2 |
|x1 x2 |
=
This shows xn is a Cauchy sequenceand must converge by Theorem 3.22.
Pn
Example 3.17. Let 1 < r < 1 and define the sequence sn = k=0 rk . (You
no doubt recognize this as the geometric series from your calculus course.) If r = 0,
the convergence of sn is trivial. So, suppose r 6= 0. In this case,
|sn+1 sn | rn+1
= n = |r| < 1
|sn sn1 | r
and sn is contractive. Theorem 3.24 implies sn converges.
7. Exercises
6n 1
3.1. Let the sequence an = . Use the definition of convergence for a
3n + 2
sequence to show an converges.
3.3. Let an be a sequence such that a2n A and a2n a2n1 0. Then an A.
3.4. If an is a sequence of positive numbers converging to 0, then an 0.
3.10. Find a sequence an such that given x [0, 1], there is a subsequence bn of
an such that bn x.
3.17. Let an and bn be sequences. Prove that both sequences an and bn converge
iff both an + bn and an bn converge.
3.18. Let an be a bounded sequence. Prove that given any > 0, there is an
interval I with length such that {n : an I} is infinite. Is it necessary that an be
bounded?
Pn
3.19. A sequence an converges in the mean if an = n1 k=1 ak converges. Prove
that if an L, then an L, but the converse is not true.
3.21. If an is a Cauchy sequence whose terms are integers, what can you say about
the sequence?
Pn
3.22. Show an = k=0 1/k! is a Cauchy sequence.
3.25. If 0 < < 1 and sn is a sequence satisfying |sn+1 | < |sn |, then sn 0.
3.32. If an is a sequence of positive numbers, then lim inf an = lim sup 1/an .
3.34. The equation x3 4x + 2 = 0 has one real root lying between 0 and 1.
Find a sequence of rational numbers converging to this root. Use this sequence to
approximate the root to five decimal places.
Series
Given a sequence an , in many contexts it is natural to ask about the sum of all
the numbers in the sequence. If only a finite number of the an are nonzero, this
is trivialand not very interesting. If an infinite number of the terms arent zero,
the path becomes less obvious. Indeed, its even somewhat questionable whether it
makes sense at all to add an infinite number of numbers.
There are many approaches to this question. The method given below is the
most common technique. Others are mentioned in the exercises.
1. What is a Series?
The idea behind adding up an infinite collection of numbers is a reduction to
the well-understood idea of a sequence. This is a typical approach in mathematics:
reduce a question to a previously solved problem.
Definition 4.1. Given a sequence an , the series having an as its terms is the
new sequence
Xn
sn = ak = a1 + a2 + + an .
k=1
The numbers sn are called the partial sums of the series. If sn S R, then the
series converges to S. This is normally written as
X
ak = S.
k=1
Example 4.2 (Geometric Series). Recall that a sequence of the form an = c rn1
is called a geometric sequence. It gives rise to a series
X
c rn1 = c + cr + cr2 + cr3 +
n=1
4-1
4-2 Series
called a geometric series. The number r is called the ratio of the series.
Suppose an = rn1 for r 6= 1. Then,
s1 = 1, s2 = 1 + r, s3 = 1 + r + r2 , . . .
In general, it can be shown by induction (or even long division) that
n n
X X 1 rn
(23) sn = ak = rk1 = .
1r
k=1 k=1
steps
2 1 1/2 0
distance from wall
In some cases, the geometric series has an intuitively plausible limit. If you
start two meters away from a wall and keep stepping halfway to the wall, no number
of steps will get you to the wall, but a large number of steps will get you as close to
the wall as you want. (See Figure 1.) So, the total distance stepped has limiting
value 2. The total distance after n steps is the nth partial sum of a geometric series
with ratio r = 1/2 and c = 1.
P
Example 4.3 (Harmonic Series). The series n=1 1/n is called the harmonic
series. It was shown in Example 3.16 that the harmonic series diverges.
Example 4.4. The terms of the sequence
1
an = 2 , n N.
n +n
can be decomposed into partial fractions as
1 1
an = .
n n+1
If sn is the series having an as its terms, then s1 = 1/2 = 1 1/2. We claim that
sn = 1 1/(n + 1) for all n N. To see this, suppose sk = 1 1/(k + 1) for some
k N. Then
1 1 1 1
sk+1 = sk + ak+1 = 1 + =1
k+1 k+1 k+2 k+2
and the claim is established by induction. Now its easy to see that
X 1 1
= lim 1 = 1.
n=1
n2 + n n n+2
This is an example of a telescoping series. The name is apparently based on the
idea that the middle terms of the series cancel, causing the series to collapse like a
hand-held telescope.
The following theorem is an easy consequence of the properties of sequences
shown in Theorem 3.8.
P P
Theorem 4.2. Let an and bn be convergent series.
P P
(a) If
P c R, then P n P an .
c a = c
(b) (an + bn ) = an + bn .
(c) an 0
Pn Pn
Proof. Let An = k=1 ak and Bn = k=1 bk be the sequences of partial
sums for each of the two series. By assumption, there are numbers A and B where
An AP and Bn B.P
n n
(a) k=1 c ak = c k=1 ak = cAn cA.
Pn Pn Pn
(b) k=1 (ak + bk ) = k=1 ak + k=1 bk = An + Bn A + B.
Pn Pn1
(c) For n > 1, an = k=1 ak k=1 ak = An An1 A A = 0.
Notice that the first two parts of Theorem 4.2 show that the set of all convergent
series is closed under linear combinations.
Theorem 4.2(c) is most useful because its contrapositive provides the most basic
test for divergence.
P
Corollary 4.3 (Going to Zero Test). If an 6 0, then an diverges.
Many have made the mistake of reading too much into Corollary 4.3. It can
only be used to show divergence. When the terms of a series do tend to zero, that
does not guarantee convergence. Example 4.3, shows Theorem 4.2(c) is necessary,
but not sufficient for convergence.
Another useful observation is that the partial sums of a convergent series are a
Cauchy sequence. The Cauchy criterion for sequences can be rephrased for series as
the following theorem, the proof of which is Exercise 4.4.
P
Theorem 4.4 (Cauchy Criterion). Let an be a series. The following state-
ments are equivalent.
P
(a) an converges.
(b) For every > 0 there is an N N such that whenever n m N ,
then n
X
ai < .
i=m
2. Positive Series
Most of the time, it is very hard or impossible to determine the exact limit of
a convergent series. We must satisfy ourselves with determining whether a series
converges, and then approximating its sum. For this reason, the study of series
usually involves learning a collection of theorems that might answer whether a given
series converges, but dont tell us to what it converges. These theorems are usually
called the convergence tests. The reader probably remembers a battery of such tests
from her calculus course. There is a myriad of such tests, and the standard ones are
presented in the next few sections, along with a few of those less widely used.
Since convergence of a series is determined by convergence of the sequence of its
partial sums, the easiest series to study are those with well-behaved partial sums.
Series with monotone sequences of partial sums are certainly the simplest such
series.
P
Definition 4.5. The series an is a positive series, if an 0 for all n.
The advantage of a positive series is that its sequence of partial sums is non-
negative and increasing. Since an increasing sequence converges if and only if it is
bounded above, there is a simple criterion to determine whether a positive series
converges. All of the standard convergence tests for positive series exploit this
criterion.
2.1. The Most Common Convergence Tests. All beginning calculus courses
contain several simple tests to determine whether positive series converge. Most of
them are presented below.
2.1.1. Comparison Tests. The most basic convergence tests are the comparison
tests. In these tests, the behavior of one series is inferred from that of another
series. Although theyre easy to use, there is one often fatal catch: in order to use
a comparison test, you must have a known series to which you can compare the
mystery series. For this reason, a wise mathematician collects example series for her
toolbox. The more samples in the toolbox, the more powerful are the comparison
tests.
P P
Theorem 4.6 (Comparison Test). Suppose an and bn are positive series
with an bn for all n.
P P
(a) If P bn converges, then so doesP an .
(b) If an diverges, then so does bn .
P P
Proof. Let An and Bn be the partial sums of an and bn , respectively. It
follows from the assumptions that An and Bn are increasing and for all n N,
(24) An Bn .
P P
If bn = B, then (24) implies
P B is an upper bound for An , and an converges.
On the other hand, if an diverges, An and the Sandwich Theorem
3.9(b) shows Bn .
a1 + a2 + a3 + a4 + a5 + a6 + a7 + a8 + a9 + + a15 +a16 +
| {z } | {z } | {z }
2a2 4a4 8a8
a1 + a2 + a3 + a4 + a5 + a6 + a7 + a8 + a9 + + a15 +a16 +
|{z} | {z } | {z } | {z }
a2 2a4 4a8 8a16
The following easy corollary is the form this test takes in most calculus books.
Its easier to use than Theorem 4.8 and suffices most of the time.
P P
Corollary 4.9 (Limit Comparison Test). Suppose an and bn are positive
series with
an
(29) = lim .
n bn
P P
If (0, ), then an and bn either both converge or both diverge.
X 1
Example 4.8. To test the series for convergence, let
2n n
1 1
an = n and bn = n .
2 n 2
Then
an 1/(2n n) 2n 1
lim = lim n
= lim n = lim = 1 (0, ).
n bn n 1/2 n 2 n n 1 n/2n
P n
Example 4.9. For any x R, the series |x |/n! converges. To see this, note
that according to Exercise 4.7,
n 1/n
|x | |x|
= 0 < 1.
n! (n!)1/n
Applying the Root Test shows the series converges.
1/n2 . The first diverges
P P
Example 4.10. Consider the p-series 1/n and
1/n 2/n
and the second converges. Since n 1 and n 1, it can be seen that when
= 1, the Root Test in inconclusive.
P
Theorem 4.11 (Ratio Test). Suppose an is a positive series. Let
an+1 an+1
r = lim inf lim sup = R.
n an n an
P P
If R < 1, then an converges. If r > 1, then an diverges.
Proof. First, suppose R < 1 and (R, 1). There exists N N such that
an+1 /an < whenever n N . This implies an+1 < an whenever n N . From
this its easy to prove by induction that aN +m < m aN whenever m N. It follows
that, for n > N ,
n
X N
X n
X
ak = ak + ak
k=1 k=1 k=N +1
N
X nN
X
= ak + aN +k
k=1 k=1
N
X nN
X
< ak + aN k
k=1 k=1
N
X aN
< ak + .
1
k=1
P P
Therefore, the partial sums of an are bounded, and an converges.
If r > 1, then choose N N so that an+1 > an for all n N . Its now apparent
that an 6 0.
In calculus books, the ratio test usually takes the following simpler form.
P
Corollary 4.12 (Ratio Test). Suppose an is a positive series. Let
an+1
r = lim .
n an
P P
If r < 1, then an converges. If r > 1, then an diverges.
From a practical viewpoint, the ratio test is often easier to apply than the root
test. But, the root test is actually the stronger of the two in the sense that there
are series for which the ratio test fails, but the root test succeeds. (See Exercise
4.11, for example.) This happens because
an+1 an+1
(30) lim inf lim inf a1/n
n lim sup a1/n
n lim sup .
an an
To see this, note the middle inequality is always true. To prove the right-hand
1/n
inequality, choose r > lim sup an+1 /an . It suffices to show lim sup an r. As in
the proof of the ratio test, an+k < rk an . This implies
an
an+k < rn+k n ,
r
which leads to
a 1/(n+k)
1/(n+k) n
an+k <r n .
r
Finally,
a 1/(n+k)
1/(n+k) n
lim sup a1/n
n = lim sup an+k lim sup r = r.
k k rn
The left-hand inequality is proved similarly.
2.2. Kummer-Type Tests. Most times the simple tests of the preceding sec-
tion suffice. However, more difficult series require more delicate tests. There dozens
of other, more specialized, convergence tests. Several of them are consequences of
the following theorem.
P
Theorem 4.13 (Kummers Test). Suppose an is a positive series, pn is a
sequence of positive numbers and
an an
(31) = lim inf pn pn+1 lim sup pn pn+1 =
an+1 an+1
P P P
If > 0, then an converges. If 1/pn diverges and < 0, then an diverges.
Pn
Proof. Let sn = k=1 ak , suppose > 0 and choose r (0, ). There must
be an N > 1 such that
an
pn pn+1 > r, n N.
an+1
Rearranging this gives
(32) pn an pn+1 an+1 > ran+1 , n N.
For M > N , (32) implies
M
X M
X
(pn an pn+1 an+1 ) > ran+1
n=N n=N
pN aN pM +1 aM +1 > r(sM sN 1 )
pN aN pM +1 aM +1 + rsN 1 > rsM
pN aN + rsN 1
> sM
r
P
Since N is fixed, the left side is an upper bound for sM , and it follows that an
converges. P
Next suppose 1/pn diverges and < 0. There must be an N N such that
an
pn pn+1 < 0, n N.
an+1
This implies
pn an < pn+1 an+1 , n N.
2See 5.2.
From Raabes test, Theorem 4.14, it follows that the series converges when p > 2
and diverges when p < 2. Raabes test is inconclusive when p = 2.
Now, suppose p = 2. Consider
an (4 + 3 n)
lim ln n n 1 1 = lim ln n 2 =0
n an+1 n 4 (1 + n)
and Bertrands test, Theorem 4.15, shows divergence.
The series (33) converges only when p > 2.
is a positive series that converges. Since |s2n s2n1 | = 1/2n 0, its clear
that
P s2n1 must also converge to the same limit. Therefore, sn converges and
(1)n+1 /n is conditionally convergent. (See Figure 3.)
To summarize: absolute convergence implies convergence, but convergence does
not imply absolute convergence.
There are a few tests that address conditional convergence. Following are the
most well-known.
Theorem 4.18 (Abels Test). Let an and bn be sequences satisfying
Pn
(a) sn = k=1 ak is a bounded sequence.
(b) bn bn+1 , n N
1.0
0.8
0.6
0.4
0.2
0 5 10 15 20 25 30 35
Figure 3. This plot shows the first 35 partial sums of the alternating
harmonic series. It converges to ln 2 0.6931, which is the level of the
dashed line. Notice how the odd partial sums decrease to ln 2 and the
even partial sums increase to ln 2.
(c) bn 0
P
Then an bn converges.
Pn
Proof. To prove the theorem, suppose | k=1 ak | < M for all n N. Let
> 0 and choose N N such that bN < /2M . If N m < n, use Lemma 4.19 to
write
n n m1
X X X
a` b` = a` b` a` b`
`=m `=1 `=1
n n `
X X X
= bn+1 a` (b`+1 b` ) ak
`=1 `=1 k=1
m1 m1 `
!
X X X
bm a` (b`+1 b` ) ak
`=1 `=1 k=1
4. Rearrangements of Series
We want to use our standard intuition about adding lists of numbers when
working with series. But, this intuition has been formed by working with finite sums
and does not always work with series.
The problem with Example 4.14 is that the series is conditionally convergent.
Such examples cannot happen with absolutely convergent series. For the most part,
absolutely convergent series behave as we are intuitively led to expect.
P P
PTheorem P 4.22. If P an is absolutely convergent and a(n) is a rearrangement
of an , then a(n) = an .
P
Pn Proof. Let an = s and > 0. Choose N N so that N m < n implies
k=m |ak | < . Choose M N such that
Proof. (Theorem 4.23) Let bn and cn be as in Lemma 4.24 and define the
subsequence a+n of bn by removing those terms for which bn = 0 and an = 6 0. Define
the
P +subsequence a of
Pn n c by removing those terms for which c n = 0. The series
a
n=1 n and a
n=1 n are still divergent because only terms equal to zero have
been removed from bn and cn .
Now, let c R and m0 = n0 = 0. According to Lemma 4.24, we can define the
natural numbers
Xn m1
X Xn
m1 = min{n : a+k > c} and n 1 = min{n : a+
k + a
` < c}.
k=1 k=1 `=1
and
p mk+1 nk+1
( !
X X X
np+1 = min n : a+
` a
`
k=0 `=mk +1 `=nk +1
np+1 n
X X
+ a+
` a
` < c .
`=mp +1 `=np +1
and
p mk+1 nk
!
X X X
0<c a+
` a
` a
np
k=0 `=mk +1 `=nk +1
Since both a+
0 and
mp anp 0, the result follows
from the Squeeze Theorem.
The argument when c is infinite is left as Exercise 4.32.
5. Exercises
X
4.3. The series (an an+1 ) converges iff the sequence an converges.
n=1
P
4.4. Prove or give a counter example: If |an | converges, then nan 0.
P
4.6.
P If n=1 an converges and bn is a bounded monotonic sequence, then
a b
n=1 n n converges.
4.7.
P Let xn be a sequence with range {0, 1, 2, 3, 4, 5, 6, 7, 8, 9}. Prove that
n
n=1 n 10
x converges.
4.13. Does
1 12 123 1234
+ + + +
3 35 357 3579
converge?
4.16. Let an be a sequence such that a2n A and a2n a2n1 0. Then
an A.
P
4.20. Find all values of x for which k=0 k 2 (x + 3)k converges.
converge?
X (x + 3)n
4.22. For what values of x does converge absolutely, converge
n=1
n4n
conditionally or diverge?
X n+6
4.23. For what values of x does 2 (x 1)n
converge absolutely, converge
n=1
n
conditionally or diverge?
P
4.24. For what positive values of does n=1 n n converge?
X n
4.25. Prove that cos sin converges.
3 n
P
4.26. For a series k=1 an with partial sums sn , define
n
1X
n = sn .
n
k=1
P
P that if k=1 an = s, then n s. Find an example where n converges, but
Prove
k=1 an does not. (If n converges, the sequence is said to be Cesaro summable.)
P
P If an is a sequence with a subsequence
4.27. Pbn , then n=1 bn is a subseries
P
of n=1 an . Prove that if every subseries of n=1 an converges, then n=1 an
converges absolutely.
P P
4.28. If n=1 an is a convergent positive series, then so is n=1 a2n . Give an
example to show the converse is not true.
P P
4.29. Prove or give a counter example: If n=1 an converges, then n=1 a2n
converges.
P
4.30. For what positive values of does n=1 n n converge?
4.31. If an 0 for all n N and there is a p > 1 such that limn np an exists
P
and is finite, then n=1 an converges. Is this true for p = 1?
transformed it to
1 x
+ = 0,
1 1/x 1 x
and then used geometric series to write it as
1 1
(44) + 2 + + 1 + x + x2 + x3 + = 0.
x x
Show how Euler did his calculation and find his mistake.
The Topology of R
Theorem 5.9. A set S R is closed iff it contains all its limit points.
Proof. () Suppose S is closed and x0 is a limit point of S. If x0 / S, then
S c open implies the existence of > 0 such that (x0 , x0 + ) S = . This
contradicts the fact that x0 is a limit point of S. Therefore, x0 S, and S contains
all its limit points.
() Since S contains all its limit points, if x0
/ S, there must exist an > 0
such that (x0 , x0 + ) S = . It follows from this that S c is open. Therefore S
is closed.
3.2. Compact Sets. There is a class of sets for which the conclusion of
Lindelofs theorem can be strengthened.
such a set is any countable set, which is a countable union of singletons. All other
sets are of the second category.4 Theorem 5.29 can be stated as Any open interval
is of the second category. Or, more generally, as Any nonempty open set is of the
second category.
A set is called a G set, if it is the countable intersection of open sets. It is
called an F set, if it is the countable union of closed sets. DeMorgans laws show
that the complement of an F set is a G set and vice versa. Its evident that any
countable subset of R is an F set, so Q is an F set.
On the other hand,
T suppose Q is a G set. Then there is a sequence of open sets
Gn such that Q = nN Gn . Since Q is dense, each Gn must be dense S and Example
5.11 shows Gcn is nowhere dense. From DeMorgans law, R = Q nN Gcn , showing
R is a first category set and violating the Baire category theorem. Therefore, Q is
not a G set.
Essentially the same argument shows any countable subset of R is a first category
set. The following protracted example shows there are uncountable sets of the first
category.
Corollaries 5.4 and 5.11 show C is closed and nonempty. In fact, the latter is
apparent because {0, 1/3, 2/3, 1} Cn for every n. At each step in the construction,
2n open middle thirds, each of length 3(n+1) were removed from the intervals
comprising Cn . The total length of the open intervals removed was
n
X 2n 1X 2
= = 1.
n=0
3n+1 3 n=0 3
Because of this, Example 5.11 implies C is nowhere dense in [0, 1].
4Baire did not define any categories other than these two. Some authors call first category
sets meager sets, so as not to make students fruitlessly wait for definitions of third, fourth and
fifth category sets.
5Cantors original work [6] is reprinted with an English translation in Edgars Classics on
Fractals [9]. Cantor only mentions his eponymous set in passing and it had actually been presented
earlier by others.
Figure 1. Shown here are the first few steps in the construction of the
Cantor Middle-Thirds set.
C is an example of a perfect set; i.e., a closed set all of whose points are limit
points of itself. (See Exercise 5.23.) Any closed set without isolated points is perfect.
The Cantor Middle-Thirds set is interesting because it is an example of a perfect
set without any interior points. Many people call any bounded perfect set without
interior points a Cantor set. Most of the time, when someone refers to the Cantor
set, they mean C.
There is another way to view the Cantor set. Notice that at the nth stage of
the construction, removing the middle thirds of the intervals comprising Cn removes
those points whose base 3 representation contains the digit 1 in position n + 1.
Then,
( )
X cn
(47) C= c= : cn {0, 2} .
n=1
3n
So, C consists of all numbers c [0, 1] that can be written in base 3 without using
the digit 1.6 P
If c C, then (47) shows c = n=1 3cnn for some sequence cn with range in
{0, 2}. Moreover, every such sequence corresponds to a unique element of C. Define
: C [0, 1] by
!
X cn X cn /2
(48) (c) = n
= .
n=1
3 n=1
2n
Since cn is a sequence from {0, 2}, then cn /2 is a sequence from {0, 1} and (c) can
be considered the binary representation of a number in [0, 1]. According to (47), it
follows that is a surjection and
( ) ( )
X cn /2 X bn
(C) = : cn {0, 2} = : bn {0, 1} = [0, 1].
n=1
2n n=1
2n
Therefore, card (C) = card ([0, 1]) > 0 .
The Cantor set is topologically small because it is nowhere dense and large from
the set-theoretic viewpoint because it is uncountable.
The Cantor set is also a set of measure zero. To see this, let Cn be as in the
construction of the Cantor set given above. Then C Cn and Cn consists of 2n
pairwise disjoint closed intervals each of length 3n . Their total length is (2/3)n .
Given > 0, choose n N so (2/3)n < /2. Each of the closed intervals comprising
Cn can be placed inside a slightly longer open interval so the sums of the lengths of
the 2n open intervals is less than .
5. Exercises
5.8. An open set G is called regular if G = (G) . Find an open set that is not
regular.
5.10. If X R and S is the collection of all sets relatively open in X, then (X, S)
is a topological space.
5.15. Prove that the set of accumulation points of any sequence is closed.
5.16. Prove any closed set is the set of accumulation points for some sequence.
5.20. Prove the union of a finite number of compact sets is compact. Give an
example to show this need not be true for the union of an infinite number of compact
sets.
5.21. (a) Give an example of a set S such that S is disconnected, but S {1} is
connected. (b) Prove that 1 must be a limit point of S.
5.26. Let f : [a, b] R be a function such that for every x [a, b] there is a
x > 0 such that f is bounded on (x x , x + x ). Prove f is bounded.
Limits of Functions
1. Basic Definitions
Definition 6.1. Let D R, x0 be a limit point of D and f : D R. The
limit of f (x) at x0 is L, if for each > 0 there is a > 0 such that when x D
with 0 < |x x0 | < , then |f (x) L| < . When this is the case, we write
limxx0 f (x) = L.
It should be noted that the limit of f at x0 is determined by the values of f
near x0 and not at x0 . In fact, f need not even be defined at x0 .
Figure 1. This figure shows a way to think about the limit. The graph
of f must stay inside the box (x0 , x0 + ) (L , L + ), except
possibly the point (x0 , f (x0 )).
A useful way of rewording the definition is to say that limxx0 f (x) = L iff
for every > 0 there is a > 0 such that x (x0 , x0 + ) D \ {x0 } implies
f (x) (L , L + ). This can also be succinctly stated as
> 0 > 0 ( f ( (x0 , x0 + ) D \ {x0 } ) (L , L + ) ) .
Example 6.1. If f (x) = c is a constant function and x0 R, then for any
positive numbers and ,
x (x0 , x0 + ) D \ {x0 } |f (x) c| = |c c| = 0 < .
This shows the limit of every constant function exists at every point, and the limit
is just the value of the function.
1
May 1, 2015
Lee
c Larson ([email protected])
6-1
6-2 CHAPTER 6. LIMITS OF FUNCTIONS
-2 2
Figure 2. The function from Example 6.3. Note that the graph is a
line with one hole in it.
1.0
0.5
-0.5
-1.0
Figure 4. This is the function from Example 6.6. The graph shown
here is on the interval [0.03, 1]. There are an infinite number of oscillations
from 1 to 1 on any open interval containing the origin.
0.10
0.05
-0.05
-0.10
Figure 5. This is the function from Example 6.7. The bounding lines
y = x and y = x are also shown. There are an infinite number of
oscillations between x and x on any open interval containing the origin.
In the same manner as Example 6.8, it can be shown for every rational function
f (x), that limxx0 f (x) = f (x0 ) whenever f (x0 ) exists.
2. Unilateral Limits
Definition 6.5. Let f : D R and x0 be a limit point of (, x0 ) D.
f has L as its left-hand limit at x0 if for all > 0 there is a > 0 such that
f ((x0 , x0 ) D) (L , L + ). In this case, we write limxx0 f (x) = L.
3. Continuity
Definition 6.9. Let f : D R and x0 D. f is continuous at x0 if for
every > 0 there exists a > 0 such that when x D with |x x0 | < , then
|f (x) f (x0 )| < . The set of all points at which f is continuous is denoted C(f ).
Several useful ways of rephrasing this are contained in the following theorem.
They are analogous to the similar statements made about limits. Proofs are left to
the reader.
Theorem 6.10. Let f : D R and x0 D. The following statements are
equivalent.
(a) x0 C(f ),
2 Calculus books often use the notation lim
xx0 f (x) = limxx0 f (x) and limxx0 f (x) =
limxx0 + f (x).
On the other hand, suppose limxx0 f (x) = f (x0 ) and > 0. Choose
according to the definition of limit. When x (x0 , x0 + ) D \ {x0 }, then
f (x) (f (x0 ) , f (x0 ) + ). It follows from this that when x = x0 , then f (x)
f (x0 ) = f (x0 ) f (x0 ) = 0 < . Therefore, when x (x0 , x0 + ) D, then
f (x) (f (x0 ) , f (x0 ) + ), and x0 C(f ), as desired.
Example 6.12. If f (x) = c, for some c R, then Example 6.1 and Theorem
6.11 show that f is continuous at every point.
Example 6.13. If f (x) = x, then Example 6.2 and Theorem 6.11 show that f
is continuous at every point.
Corollary 6.12. Let f : D R and x0 D. x0 C(f ) iff whenever xn is a
sequence from D with xn x0 , then f (xn ) f (x0 ).
Proof. Combining Theorem 6.11 with Theorem 6.2 shows this to be true.
Example 6.15 (Salt and Pepper Function). Since Q is a countable set, it can
be written as a sequence, Q = {qn : n N}. Define
(
1/n, x = qn ,
f (x) =
0, x Qc .
If x Q, then x = qn , for some n and f (x) = 1/n > 0. There is a sequence xn
from Qc such that xn x and f (xn ) = 0 6 f (x) = 1/n. Therefore C(f ) Q = .
On the other hand, let x Qc and > 0. Choose N N large enough so
that 1/N < and let = min{|x qn | : 1 n N }. If |x y| < , there
are two cases to consider. If y Qc , then |f (y) f (x)| = |0 0| = 0 < . If
y Q, then the choice of guarantees y = qn for some n > N . In this case,
|f (y) f (x)| = f (y) = f (qn ) = 1/n < 1/N < . Therefore, x C(f ).
This shows that C(f ) = Qc .
It is a consequence of the Baire category theorem that there is no function f
such that C(f ) = Q. Proving this would take us too far afield.
The following theorem is an almost immediate consequence of Theorem 6.4.
Theorem 6.13. Let f : D R and g : D R. If x0 C(f ) C(g), then
(a) x0 C(f + g),
(b) x0 C(f ), R,
(c) x0 C(f g), and
(d) x0 C(f /g) when g(x0 ) 6= 0.
4. Unilateral Continuity
Definition 6.16. Let f : D R and x0 D. f is left-continuous at x0 if for
every > 0 there is a > 0 such that f ((x0 , x0 ] D) (f (x0 ) , f (x0 ) + ).
Let f : D R and x0 D. f is right-continuous at x0 if for every > 0 there
is a > 0 such that f ([x0 , x0 + ) D) (f (x0 ) , f (x0 ) + ).
7
6
5
4
3
x2 4
2 f (x) = x2
1 2 3 4
Figure 7. The function from Example 6.19. Note that the graph is a
line with one hole in it. Plugging up the hole removes the discontinuity.
In the case, when limxx0 f (x) 6= limxx0 f (x), it is said that a jump discontinuity
occurs at x0 .
Example 6.18. The function
(
|x|/x, x 6= 0
f (x) = .
0, x=0
has a jump discontinuity at x = 0.
In the case when limxx0 f (x) = limxx0 f (x) 6= f (x0 ), it is said that f has a
removable discontinuity at x0 . The discontinuity is called removable because in
this case, the function can be made continuous at x0 by merely redefining its value
at the single point, x0 , to be the value of the two one-sided limits.
2
4
Example 6.19. The function f (x) = xx2 is not continuous at x = 2 because 2
is not in the domain of f . Since limx2 f (x) = 4, if the domain of f is extended to
include 2 by setting f (2) = 4, then this extended f is continuous everywhere. (See
Figure 7.)
Theorem 6.18. If f : (a, b) R is monotone, then (a, b) \ C(f ) is countable.
Proof. In light of the discussion above and Theorem 6.7, it is apparent that
the only types of discontinuities f can have are jump discontinuities.
To be specific, suppose f is increasing and x0 , y0 (a, b) \ C(f ) with x0 < y0 .
In this case, the fact that f is increasing implies
lim f (x) < lim f (x) lim f (x) < lim f (x).
xx0 xx0 xy0 xy0
This implies that for any two x0 , y0 (a, b) \ C(f ), there are disjoint open intervals,
Ix0 = (limxx0 f (x), limxx0 f (x)) and Iy0 = (limxy0 f (x), limxy0 f (x)). For each
x (a, b) \ C(f ), choose qx Ix Q. Because of the pairwise disjointness of the
intervals {Ix : x (a, b) \ C(f )}, this defines an bijection between (a, b) \ C(f ) and
a subset of Q. Therefore, (a, b) \ C(f ) must be countable.
A similar argument holds for a decreasing function.
Theorem 6.18 implies that a monotone function is continuous at nearly every
point in its domain. Characterizing the points of discontinuity as countable is the
best that can be hoped for, as seen in the following example.
5. Continuous Functions
Up until now, continuity has been considered as a property of a function at a
point. There is much that can be said about functions continuous everywhere.
Definition 6.19. Let f : D R and A D. We say f is continuous on A if
A C(f ). If D = C(f ), then f is continuous.
Continuity at a point is, in a sense, a metric property of a function because it
measures relative distances between points in the domain and image sets. Continuity
on a set becomes more of a topological property, as shown by the next few theorems.
Theorem 6.20. f : D R is continuous iff whenever G is open in R, then
f 1 (G) is relatively open in D.
Proof. () Assume f is continuous on D and let G be open in R. Let
x f 1 (G) and choose > 0 such that (f (x) , f (x) + ) G. Using the
continuity of f at x, we can find a > 0 such that f ((x , x + ) D) G. This
implies that (x , x + ) D f 1 (G). Because x was an arbitrary element of
f 1 (G), it follows that f 1 (G) is open.
() Choose x D and let > 0. By assumption, the set f 1 ((f (x)
, f (x) + ) is relatively open in D. This implies the existence of a > 0 such
that (x , x + ) D f 1 ((f (x) , f (x) + ). It follows from this that
f ((x , x + ) D) (f (x) , f (x) + ), and x C(f ).
A function as simple as any constant function demonstrates that f (G) need not
be open when G is open. Defining f : [0, ) R by f (x) = sin x tan1 x shows
that the image of a closed set need not be closed because f ([0, )) = (/2, /2).
Theorem 6.21. If f is continuous on a compact set K, then f (K) is compact.
Proof. Let O be an open cover of f (K) and I = {f 1 (G) : G O}. By
Theorem 6.20, I is a collection of sets which are relatively open in K. Since I covers
K, I is an open cover of K. Using the fact that K is compact, we can choose a finite
subcover of K from I, say {G1 , G2 , . . . , Gn }. There are {H1 , H2 , . . . , Hn } O such
that f 1 (Hk ) = Gk for 1 k n. Then
[ [
f (K) f Gk = Hk .
1kn 1kn
On the other hand, if 0 [a, b], then there must exist an n Z such that both
2 2 2 2
(4n+1) , (4n+3) [a, b]. Since f ( (4n+1) ) = 1, f ( (4n+3) ) = 1 and f is continuous
on the interval between them, we see f ([a, b]) = [1, 1], which is the entire range of
f . The claim now follows.
6. Uniform Continuity
Most of the ideas contained in this section will not be needed until we begin
developing the properties of the integral in Chapter 8.
Definition 6.28. A function f : D R is uniformly continuous if for all > 0
there is a > 0 such that when x, y D with |x y| < , then |f (x) f (y)| < .
The idea here is that in the ordinary definition of continuity, the in the
definition depends on both the and the x at which continuity is being tested. With
uniform continuity, only depends on ; i. e., the same works uniformly across
the whole domain.
Theorem 6.29. If f : D R is uniformly continuous, then it is continuous.
Proof. This proof is left as Exercise 6.31.
7. Exercises
6.2. Give examples of functions f and g such that neither function has a limit at
a, but f + g does. Do the same for f g.
6.5. If lim f (x) = L > 0, then there is a > 0 such that f (x) > 0 when
xa
0 < |x a| < .
6.12. If f : R R and there is an > 0 such that |f (x) f (y)| |x y| for all
x, y R, then show that f is continuous.
6.16. Let f and g be two functions which are continuous on a set D R. Prove
or give a counter example: {x D : f (x) > g(x)} is relatively open in D.
6.17. If f, g : R R are functions such that f (x) = g(x) for all x Q and
C(f ) = C(g) = R, then f = g.
6.19. Find an example to show the conclusion of Problem 18 fails if I = (a, b).
6.20. If f and g are both continuous on [a, b], then {x : f (x) g(x)} is compact.
6.22. Suppose f : R R is a function such that every interval has points at which
f is negative and points at which f is positive. Prove that every interval has points
where f is not continuous.
6.23. If f : [a, b] R has a limit at every point, then f is bounded. Is this true
for f : (a, b) R?
Differentiation
1
May 1, 2015
Lee
c Larson ([email protected])
7-1
7-2 CHAPTER 7. DIFFERENTIATION
Figure 1. These graphs illustrate that the two standard ways of writing
the difference quotient are equivalent.
2. Differentiation Rules
Following are the standard rules for differentiation learned in every beginning
calculus course.
Theorem 7.3. Suppose f and g are functions such that x0 D(f ) D(g).
(a) x0 D(f + g) and (f + g)0 (x0 ) = f 0 (x0 ) + g 0 (x0 ).
(b) If a R, then x0 D(af ) and (af )0 (x0 ) = af 0 (x0 ).
(c) x0 D(f g) and (f g)0 (x0 ) = f 0 (x0 )g(x0 ) + f (x0 )g 0 (x0 ).
(d) If g(x0 ) 6= 0, then x0 D(f /g) and
0
f f 0 (x0 )g(x0 ) f (x0 )g 0 (x0 )
(x0 ) = .
g (g(x0 ))2
2Bolzano presented his example in 1834, but it was little noticed. The 1872 example of
Weierstrass is more well-known [2]. A translation of Weierstrass original paper [19] is presented
by Edgar [9]. Weierstrass example is not very transparent because it depends on trigonometric
series. Many more elementary constructions have since been made. One such will be presented in
Example 9.9.
Proof. (a)
(f + g)(x0 + h) (f + g)(x0 )
lim
h0 h
f (x0 + h) + g(x0 + h) f (x0 ) g(x0 )
= lim
h0 h
f (x0 + h) f (x0 ) g(x0 + h) g(x0 )
= lim + = f 0 (x0 ) + g 0 (x0 )
h0 h h
(b)
(af )(x0 + h) (af )(x0 ) f (x0 + h) f (x0 )
lim = a lim = af 0 (x0 )
h0 h h0 h
(c)
(f g)(x0 + h) (f g)(x0 ) f (x0 + h)g(x0 + h) f (x0 )g(x0 )
lim = lim
h0 h h0 h
Now, slip a 0 into the numerator and factor the fraction.
f (x0 + h)g(x0 + h) f (x0 )g(x0 + h) + f (x0 )g(x0 + h) f (x0 )g(x0 )
= lim
h0 h
f (x0 + h) f (x0 ) g(x0 + h) g(x0 )
= lim g(x0 + h) + f (x0 )
h0 h h
Finally, use the definition of the derivative and the continuity of f and g at x0 .
= f 0 (x0 )g(x0 ) + f (x0 )g 0 (x0 )
(d) It will be proved that if g(x0 ) 6= 0, then (1/g)0 (x0 ) = g 0 (x0 )/(g(x0 ))2 . This
statement, combined with (c), yields (d).
1 1
(1/g)(x0 + h) (1/g)(x0 ) g(x0 + h) g(x0 )
lim = lim
h0 h h0 h
g(x0 ) g(x0 + h) 1
= lim
h0 h g(x0 + h)g(x0 )
g 0 (x0 )
=
(g(x0 )2
Plug this into (c) to see
0 0
f 1
(x0 ) = f (x0 )
g g
1 g 0 (x0 )
= f 0 (x0 ) + f (x0 )
g(x0 ) (g(x0 ))2
f 0 (x0 )g(x0 ) f (x0 )g 0 (x0 )
= .
(g(x0 ))2
Combining Examples 7.1 and 7.2 with Theorem 7.3, the following theorem is
easy to prove.
Corollary 7.4. A rational function is differentiable at every point of its
domain.
Theorem 7.5 (Chain Rule). If f and g are functions such that x0 D(f ) and
f (x0 ) D(g), then x0 D(g f ) and (g f )0 (x0 ) = g 0 f (x0 )f 0 (x0 ).
Proof. Let y0 = f (x0 ). By assumption, there is an open interval J containing
f (x0 ) such that g is defined on J. Since J is open and x0 C(f ), there is an open
interval I containing x0 such that f (I) J.
Define h : J R by
g(y) g(y0 ) g 0 (y ), y 6= y
0 0
h(y) = y y0 .
0, y = y0
From the definition of h f for x I with f (x) 6= f (x0 ), we can solve for
(54) g f (x) g f (x0 ) = (h f (x) + g 0 f (x0 ))(f (x) f (x0 )).
Notice that (54) is also true when f (x) = f (x0 ). Divide both sides of (54) by x x0 ,
and use (53) to obtain
g f (x) g f (x0 ) f (x) f (x0 )
lim = lim (h f (x) + g 0 f (x0 ))
xx0 x x0 xx0 x x0
= (0 + g 0 f (x0 ))f 0 (x0 )
= g 0 f (x0 )f 0 (x0 ).
Theorem 7.6. Suppose f : [a, b] [c, d] is continuous and invertible. If
x0 D(f ) and f 0 (x0 ) 6= 0 for some x0 (a, b), then f (x0 ) D(f 1 ) and
0
f 1 (f (x0 )) = 1/f 0 (x0 ).
Proof. Let y0 = f (x0 ) and suppose yn is any sequence in f ([a, b]) \ {y0 }
converging to y0 and xn = f 1 (yn ). By Theorem 6.24, f 1 is continuous, so
x0 = f 1 (y0 ) = lim f 1 (yn ) = lim xn .
n n
Therefore,
f 1 (yn ) f 1 (y0 ) xn x0 1
lim = lim = 0 .
n yn y0 n f (xn ) f (x0 ) f (x0 )
Example 7.4. It follows easily from Theorem 7.3 that f (x) = x3 is differentiable
everywhere with f (x) = 3x . Define g(x) = x. Then g(x) = f 1 (x). Suppose
0 2 3
In the same manner as Example 7.4, the following corollary can be proved.
Corollary 7.7. Suppose q Q, f (x) = xq and D is the domain of f . Then
0
f (x) = qxq1 on the set
(
D, when q 1
.
D \ {0}, when q < 1
Theorem 7.9 is, of course, the basis for much of a beginning calculus course. If
f : [a, b] R, then the extreme values of f occur at points of the set
C = {x (a, b) : f 0 (x) = 0} {x [a, b] : f 0 (x) does not exist}.
The elements of C are often called the critical points or critical numbers of f on
[a, b]. To find the maximum and minimum values of f on [a, b], it suffices to find its
maximum and minimum on the smaller set C, which is finite in elementary calculus
courses.
4. Differentiable Functions
Differentiation becomes most useful when a function has a derivative at each
point of an interval.
Definition 7.10. The function f is differentiable on an open interval I if
I D(f ). If f is differentiable on its domain, then it is said to be differentiable. In
this case, the function f 0 is called the derivative of f .
The fundamental theorem about differentiable functions is the Mean Value
Theorem. Following is its simplest form.
Lemma 7.11 (Rolles Theorem). If f : [a, b] R is continuous on [a, b],
differentiable on (a, b) and f (a) = 0 = f (b), then there is a c (a, b) such that
f 0 (c) = 0.
Proof. Since [a, b] is compact, Corollary 6.23 implies the existence of xm , xM
[a, b] such that f (xm ) f (x) f (xM ) for all x [a, b]. If f (xm ) = f (xM ), then f is
constant on [a, b] and any c (a, b) satisfies the lemma. Otherwise, either f (xm ) < 0
or f (xM ) > 0. If f (xm ) < 0, then xm (a, b) and Theorem 7.9 implies f 0 (xm ) = 0.
If f (xM ) > 0, then xM (a, b) and Theorem 7.9 implies f 0 (xM ) = 0.
Rolles Theorem is just a stepping-stone on the path to the Mean Value Theorem.
Two versions of the Mean Value Theorem follow. The first is a version more general
than the one given in most calculus courses. The second is the usual version.4
Theorem 7.12 (Cauchy Mean Value Theorem). If f : [a, b] R and g : [a, b]
R are both continuous on [a, b] and differentiable on (a, b), then there is a c (a, b)
such that
g 0 (c)(f (b) f (a)) = f 0 (c)(g(b) g(a)).
Proof. Let
h(x) = (g(b) g(a))(f (a) f (x)) + (g(x) g(a))(f (b) f (a)).
Because of the assumptions on f and g, h is continuous on [a, b] and differentiable
on (a, b) with h(a) = h(b) = 0. Theorem 7.11 yields a c (a, b) such that h0 (c) = 0.
Then
0 = h0 (c) = (g(b) g(a))f 0 (c) + g 0 (c)(f (b) f (a))
= g 0 (c)(f (b) f (a)) = f 0 (c)(g(b) g(a)).
But, derivatives do share one useful property with continuous functions; they sat-
isfy an intermediate value property. Compare the following theorem with Corollary
6.26.
Theorem 7.16 (Darbouxs Theorem). If f is differentiable on an open set
containing [a, b] and is between f 0 (a) and f 0 (b), then there is a c [a, b] such that
f 0 (c) = .
Proof. If f 0 (a) = f 0 (b), then c = a satisfies the theorem. So, we may as well
assume f 0 (a) 6= f 0 (b). There is no generality lost in assuming f 0 (a) < f 0 (b), for,
otherwise, we just replace f with g = f .
Let h(x) = f (x) x so that D(f ) = D(h) and h0 (x) = f 0 (x) . In particular,
this implies h0 (a) < 0 < h0 (b). Because of this, there must be an > 0 small enough
so that
h(a + ) h(a)
< 0 = h(a + ) < h(a)
and
h(b) h(b )
> 0 = h(b ) < h(b).
(See Figure 3.) In light of these two inequalities and Theorem 6.23, there must
be a c (a, b) such that h(c) = glb {h(x) : x [a, b]}. Now Theorem 7.9 gives
0 = h0 (c) = f 0 (c) , and the theorem follows.
Heres an example showing a possible use of Theorem 7.16.
Example 7.6. Let (
0, x 6= 0
f (x) = .
1, x=0
Theorem 7.16 implies f is not a derivative.
A more striking example is the following
Example 7.7. Define
( (
sin x1 , x 6= 0 sin x1 , x 6= 0
f (x) = and g(x) = .
1, x=0 1, x=0
Since (
0, x 6= 0
f (x) g(x) =
2, x=0
does not have the intermediate value property, at least one of f or g is not a
derivative. (Actually, neither is a derivative because f (x) = g(x).)
5.1. Taylors Theorem. The motivation behind Taylors theorem is the at-
tempt to approximate a function f near a number a by a polynomial. The polynomial
of degree 0 which does the best job is clearly p0 (x) = f (a). The best polynomial of
degree 1 is the tangent line to the graph of the function p1 (x) = f (a) + f 0 (a)(x a).
Continuing in this way, we approximate f near a by the polynomial pn of degree n
(k)
such that f (k) (a) = pn (a) for k = 0, 1, . . . , n. A simple induction argument shows
that
n
X f (k) (a)
(57) pn (x) = (x a)k .
k!
k=0
and define
n
!
X f (k) (x) k
F (x) = f (b) (b x) + (b x)n+1 .
k! (n + 1)!
k=0
From (58) we see that F (a) = 0. Direct substitution in the definition of F shows
that F (b) = 0. From the assumptions in the statement of the theorem, it is easy
to see that F is continuous on [a, b] and differentiable on (a, b). An application of
Rolles Theorem yields a c (a, b) such that
(n+1)
0 f (c) n n
0 = F (c) = (b c) (b c) = = f (n+1) (c),
n! n!
as desired.
4
n=4 n=8
n = 20
2 4 6 8
y = cos(x)
-2
-4 n=2 n=6 n = 10
Figure 4. Here are several of the Taylor polynomials for the function
cos(x) graphed along with cos(x).
6There are several different formulas for the error. The one given here is sometimes called
the Lagrange form of the remainder. In Example 8.4 a form of the remainder using integration
instead of differentiation is derived.
Figure 5 below has a graph of this function. In Example 7.11 below it is shown that
f is differentiable to all orders everywhere and f (n) (0) = 0 for all n 0. With this
function the Taylor polynomial centered at 0 gives a useless approximation.
5.2. LHospitals Rules and Indeterminate Forms. According to Theo-
rem 6.4,
f (x) limxa f (x)
lim =
xa g(x) limxa g(x)
whenever limxa f (x) and limxa g(x) both exist and limxa g(x) 6= 0. But, it
is easy to find examples where both limxa f (x) = 0 and limxa g(x) = 0 and
limxa f (x)/g(x) exists, as well as similar examples where limxa f (x)/g(x) fails
to exist. Because of this, such a limit problem is said to be in the indeterminate
form 0/0. The following theorem allows us to determine many such limits.
Theorem 7.18 (Easy LHospitals Rule). Suppose f and g are each continuous
on [a, b], differentiable on (a, b) and f (b) = g(b) = 0. If g 0 (x) 6= 0 on (a, b) and
limxb f 0 (x)/g 0 (x) = L, where L could be infinite, then limxb f (x)/g(x) = L.
Proof. Let x [a, b), so f and g are continuous on [x, b] and differentiable on
(x, b). Cauchys Mean Value Theorem, Theorem 7.12, implies there is a c(x) (x, b)
such that
f (x) f 0 (c(x))
f 0 (c(x))g(x) = g 0 (c(x))f (x) = = 0 .
g(x) g (c(x))
Since x < c(x) < b, it follows that limxb c(x) = b. This shows that
f 0 (x) f 0 (c(x)) f (x)
L = lim 0
= lim 0 = lim .
xb g (x) xb g (c(x)) xb g(x)
Several things should be noted about this proof. First, there is nothing special
about the left-hand limit used in the statement of the theorem. It could just as
easily be written in terms of the right-hand limit. Second, if limxa f (x)/g(x) is
not of the indeterminate form 0/0, then applying LHospitals rule will usually give
a wrong answer. To see this, consider
x 1
lim = 0 6= 1 = lim .
x0 x + 1 x0 1
Another case where the indeterminate form 0/0 occurs is in the limit at infinity.
That LHopitals rule works in this case can easily be deduced from Theorem 7.18.
Corollary 7.19. Suppose f and g are differentiable on (a, ) and
lim f (x) = lim g(x) = 0.
x x
If g 0 (x) 6= 0 on (a, ) and limx f 0 (x)/g 0 (x) = L, where L could be infinite, then
limx f (x)/g(x) = L.
Proof. There is no generality lost by assuming a > 0. Let
( (
f (1/x), x (0, 1/a] g(1/x), x (0, 1/a]
F (x) = and G(x) = .
0, x=0 0, x=0
Then
lim F (x) = lim f (x) = 0 = lim g(x) = lim G(x),
x0 x x x0
so both F and G are continuous at 0. It follows that both F and G are continuous
on [0, 1/a] and differentiable on (0, 1/a) with G0 (x) = g 0 (x)/x2 = 6 0 on (0, 1/a)
and limx0 F 0 (x)/G0 (x) = limx f 0 (x)/g 0 (x) = L. The rest follows from Theorem
7.18.
The other standard indeterminate form arises when
lim f (x) = = lim g(x).
x x
then
f (x)
lim = L.
x g(x)
Proof. First, suppose L R and let > 0. Choose a1 > a large enough so
that 0
f (x)
g 0 (x) L < , x > a1 .
Since limx f (x) = = limx g(x), we can assume there is an a2 > a1 such
that both f (x) > 0 and g(x) > 0 when x > a2 . Finally, choose a3 > a2 such that
whenever x > a3 , then f (x) > f (a2 ) and g(x) > g(a2 ).
Let x > a3 and apply Cauchys Mean Value Theorem, Theorem 7.12, to f and
g on [a2 , x] to find a c(x) (a2 , x) such that
f 0 (c(x)) f (x) f (a2 ) f (x) 1 ff(a 2)
(x)
(60) = = .
g 0 (c(x))
g(x) g(a2 ) g(x) 1 g(a2 )
g(x)
If
g(a2 )
1 g(x)
h(x) = f (a2 )
,
1 f (x)
then (60) implies
f (x) f 0 (c(x))
= 0 h(x).
g(x) g (c(x))
Since limx h(x) = 1, there is an a4 > a3 such that whenever x > a4 , then
|h(x) 1| < . If x > a4 , then
0
f (x) f (c(x))
g(x) L = g 0 (c(x)) h(x) L
0
f (c(x))
= 0
h(x) Lh(x) + Lh(x) L
g (c(x))
0
f (c(x))
0
L |h(x)| + |L||h(x) 1|
g (c(x))
< (1 + ) + |L| = (1 + |L| + )
3 2 1 1 2 3
The case when L = is done similarly by first choosing a B > 0 and adjusting
(60) so that f 0 (x)/g 0 (x) > B when x > a1 . A similar adjustment is necessary when
L = .
There is a companion corollary to Theorem 7.20 which is proved in the same
way as Corollary 7.19.
Corollary 7.21. Suppose that f and g are continuous on [a, b] and differen-
tiable on (a, b) with g 0 (x) 6= 0 on (a, b). If
f 0 (x)
lim f (x) = lim g(x) = and lim = L R {, },
xa xa xa g 0 (x)
then
f (x)
lim = L.
xa g(x)
Example 7.10. If > 0, then limx ln x/x is of the indeterminate form
/. Taking derivatives of the numerator and denominator yields
1/x 1
lim1
= lim = 0.
x x x x
Theorem 7.20 now implies limx ln x/x = 0, and therefore ln x increases more
slowly than any positive power of x.
Example 7.11. Let f be as in Example 7.9. (See Figure 5.) It is clear f (n) (x)
exists whenever n and x 6= 0. We claim f (n) (0) = 0. To see this, we first prove
that
2
e1/x
(61) lim = 0, n Z.
x0 xn
When n 0, (61) is obvious. So, suppose (61) is true whenever m n for some
n . Making the substitution u = 1/x, we see
2
e1/x un+1
(62) lim n+1 = lim u2 .
x0 x u e
Since
2
(n + 1)un (n + 1)un1 n+1 e1/x
lim 2 = lim 2 = lim =0
u 2ueu u 2eu 2 x0 xn1
by the inductive hypothesis, Theorem 7.20 gives (62) in the case of the right-hand
limit. The left-hand limit is handled similarly. Finally, (61) follows by induction.
When x 6= 0, a bit of experimentation can convince the reader that f (n) (x)
2
is of the form pn (1/x)e1/x , where pn is a polynomial. Induction and repeated
applications of (61) establish that f (n) (0) = 0 for n .
6. Exercises
7.1. If (
x2 , x Q
f (x) = ,
0, otherwise
then show D(f ) = {0} and find f 0 (0).
7.7. If (
1/2, x=0
f1 (x) =
6 0
sin(1/x), x =
and (
1/2, x=0
f2 (x) = ,
sin(1/x), x 6= 0
then at least one of f1 and f2 is not in .
7.9. Suppose f is differentiable everywhere and f (x+y) = f (x)f (y) for all x, y R.
Show that f 0 (x) = f 0 (0)f (x) and determine the value of f 0 (0).
d
7.11. Use the definition of the derivative to find x.
dx
7.12. Let f be continuous on [0, ) and differentiable on (0, ). If f (0) = 0 and
|f 0 (x)| < |f (x)| for all x > 0, then f (x) = 0 for all x 0.
7.15. Let f be continuous on [a, b] and differentiable on (a, b). If f (a) = and
|f 0 (x)| < for all x (a, b), then calculate a bound for f (b).
7.18. Prove or give a counter example: If f D((a, b)) such that f 0 is bounded,
then there is an F C([a, b]) such that f = F on (a, b).
7.20. Suppose that I is an open interval and that f 00 (x) 0 for all x I. If a I,
then show that the part of the graph of f on I is never below the tangent line to
the graph at (a, f (a)).
7A function g is even if g(x) = g(x) for every x and it is odd if g(x) = g(x) for every
x. The terms are even and odd because this is how g(x) = xn behaves when n is an even or odd
integer, respectively.
Integration
Contrary to the impression given by most calculus courses, there are many ways
to define integration. The one given here is called the Riemann integral or the
Riemann-Darboux integral, and it is the one most commonly presented to calculus
students.
1. Partitions
A partition of the interval [a, b] is a finite set P [a, b] such that {a, b} P .
The set of all partitions of [a, b] is denoted part ([a, b]). Basically, a partition should
be thought of as a way to divide an interval into a finite number of subintervals by
choosing some points where it is divided.
If P part ([a, b]), then the elements of P can be ordered in a list as a = x0 <
x1 < < xn = b. The adjacent points of this partition determine n compact
intervals of the form IkP = [xk1 , xk ], 1 k n. If the partition is clear from the
context, we write Ik instead of IkP . Its clear that these intervals only intersect at
their common endpoints and there is no requirement they have the same length.
Since its inconvenient to always list each part of a partition, well use the
partition of the previous paragraph as the generic partition. Unless its necessary
within the context to specify some other form for a partition, assume any partition
is the generic partition. (See Figure 1.)
If I is any interval, its length is written |I|. Using the notation of the previous
paragraph, it follows that
n
X n
X
|Ik | = (xk xk1 ) = xn x0 = b a.
k=1 k=1
I1 I2 I3 I4 I5
x0 x1 x2 x3 x4 x5
a b
kP k kQk. It also follows at once from the definitions that P Q part ([a, b])
with P P Q and Q P Q. The partition P Q is called the common
refinement of P and Q.
2. Riemann Sums
Let f : [a, b] R and P part ([a, b]). Choose xk Ik for each k. The set
{xk : 1 k n} is called a selection from P . The expression
n
X
R (f, P, xk ) = f (xk )|Ik |
k=1
is the Riemann sum for f with respect to the partition P and selection xk . The
Riemann sum is the usual first step toward integration in a calculus course and can
be visualized as the sum of the areas of rectangles with height f (xk ) and width |Ik |
as long as the rectangles are allowed to have negative area when f (xk ) < 0. (See
Figure 2.)
Notice that given a particular function f and partition P , there are an uncount-
ably infinite number of different possible Riemann sums, depending on the selection
xk . This sometimes makes working with Riemann sums quite complicated.
Example 8.1. Suppose f : [a, b] R is the constant function f (x) = c. If
P part ([a, b]) and {xk : 1 k n} is any selection from P , then
n
X n
X
R (f, P, xk ) = f (xk )|Ik | = c |Ik | = c(b a).
k=1 k=1
Example 8.2. Suppose f (x) = x on [a, b]. Choose any P part ([a, b]) where
kP k < 2(b a)/n. (Convince yourself this is always possible.1) Make two specific
y = f (x)
a x1 x1 x2 x2 x3 x3 x4 b
x0 x4
Figure 2. The Riemann sum R (f, P, xk ) is the sum of the areas of the
rectangles in this figure. Notice the right-most rectangle has negative
area because f (x4 ) < 0.
Definition 8.1. The function f is Riemann integrable on [a, b], if there exists a
number R (f ) such that for all > 0 there is a > 0 so that whenever P part ([a, b])
with kP k < , then
|R (f ) R (f, P, xk ) | <
for any selection xk from P .
Theorem 8.2. If f : [a, b] R and R (f ) exists, then R (f ) is unique.
Proof. Suppose R1 (f ) and R2 (f ) both satisfy the definition and > 0. For
i = 1, 2 choose i > 0 so that whenever kP k < i , then
|Ri (f ) R (f, P, xk ) | < /2,
as in the definition above. If P part ([a, b]) so that kP k < 1 2 , then
|R1 (f ) R2 (f )| |R1 (f ) R (f, P, xk ) | + |R2 (f ) R (f, P, xk ) | <
and it follows R1 (f ) = R2 (f ).
3. Darboux Integration
As mentioned above, a difficulty with handling Riemann sums is there are an
uncountably so many different ways to choose partitions and selections that working
with them is unwieldy. One way to resolve this problem was shown in Example 8.2,
where it was easy to find largest and smallest Riemann sums associated with each
partition. However, thats not always a straightforward calculation, so to use that
idea, a little more care must be taken.
Definition 8.4. Let f : [a, b] R be bounded and P part ([a, b]). For each
Ik determined by P , let
Mk = lub {f (x) : x Ik } and mk = glb {f (x) : x Ik }.
The upper and lower Darboux sums for f on [a, b] are
n
X n
X
D (f, P ) = Mk |Ik | and D (f, P ) = mk |Ik |.
k=1 k=1
Then
mk0 ml Ml Mk0 and mk0 mr Mr Mk0
so that
mk0 |Ik0 | = mk0 (|[xk0 1 , x]| + |[x, xk0 ]|)
ml |[xk0 1 , x]| + mr |[x, xk0 ]|
Ml |[xk0 1 , x]| + Mr |[x, xk0 ]|
Mk0 |[xk0 1 , x]| + Mk0 |[x, xk0 ]|
= Mk0 |Ik0 |.
This implies
n
X
D (f, P ) = mk |Ik |
k=1
X
= mk |Ik | + mk0 |Ik0 |
k6=k0
X
mk |Ik | + ml |[xk0 1 , x]| + mr |[x, xk0 ]|
k6=k0
= D (f, Q)
D (f, Q)
X
= Mk |Ik | + Ml |[xk0 1 , x]| + Mr |[x, xk0 ]|
k6=k0
X n
Mk |Ik |
k=1
= D (f, P )
The argument given above shows that the theorem holds if Q has one more point
than P . Using induction, this same technique also shows the theorem holds when Q
has an arbitrarily larger number of points than P .
The main lesson to be learned from Theorem 8.5 is that refining a partition
causes the lower Darboux sum to increase and the upper Darboux sum to decrease.
Moreover, if P, Q part ([a, b]) and f : [a, b] [B, B], then,
B(b a) D (f, P ) D (f, P Q) D (f, P Q) D (f, Q) B(b a).
Therefore every Darboux lower sum is less than or equal to every Darboux upper
sum. Consider the following definition with this in mind.
Definition 8.6. The upper and lower Darboux integrals of a bounded function
f : [a, b] R are
D (f ) = glb {D (f, P ) : P part ([a, b])}
and
D (f ) = lub {D (f, P ) : P part ([a, b])},
respectively.
As a consequence of the observations preceding the definition, it follows that
D (f ) D (f ) always. In the case D (f ) = D (f ), the function is said to be Darboux
integrable on [a, b], and the common value is written D (f ).
The following is obvious.
1
(b a) + mkP k
N
< (b a) + m
2(b a) 2m
= .
Since f (x) = 0 whenever x Qc , it follows that D (f, P ) = 0. Therefore, D (f ) =
D (f ) = 0 and D (f ) = 0.
4. The Integral
There are now two different definitions for the integral. It would be embarassing,
if they gave different answers. The following theorem shows theyre really different
sides of the same coin.2
Theorem 8.9. Let f : [a, b] R.
(a) R (f ) exists iff D (f ) exists.
(b) If R (f ) exists, then R (f ) = D (f ).
Proof. (a) (=) Suppose R (f ) exists and > 0. By Theorem 8.3, f is
bounded. Choose P part ([a, b]) such that
|R (f ) R (f, P, xk ) | < /4
for all selections xk from P . From each Ik , choose xk and xk so that
Mk f (xk ) < and f (xk ) mk < .
4(b a) 4(b a)
Then
n
X n
X
D (f, P ) R (f, P, xk ) = Mk |Ik | f (xk )|Ik |
k=1 k=1
Xn
= (Mk f (xk ))|Ik |
k=1
< (b a) = .
4(b a) 4
In the same way,
R (f, P, xk ) D (f, P ) < /4.
Therefore,
D (f ) D (f )
= glb {D (f, Q) : Q part ([a, b])} lub {D (f, Q) : Q part ([a, b])}
D (f, P ) D (f, P )
< R (f, P, xk ) + R (f, P, xk )
4 4
|R (f, P, xk ) R (f, P, xk )| +
2
< |R (f, P, xk ) R (f ) | + |R (f ) R (f, P, xk ) | +
2
<
Since is an arbitrary positive number, this shows D (f ) exists and equals R (f ),
which is part (b) of the theorem.
2Theorem 8.9 shows that the two integrals presented here are the same. But, there are many
other integrals, and not all of them are equivalent. For example, the well-known Lebesgue integral
includes all Riemann integrable functions, but not all Lebesgue integrable functions are Riemann
integrable. The Denjoy integral is another extension of the Riemann integral which is not the same
as the Lebesgue integral. For more discussion of this, see [10].
(=) Suppose f : [a, b] [B, B], D (f ) exists and > 0. Since D (f ) exists,
there is a P1 part ([a, b]), with points a = p0 < < pm = b, such that
D (f, P1 ) D (f, P1 ) < .
2
Set = /8mB. Choose P part ([a, b]) with kP k < and let P2 = P P1 . Since
P1 P2 , according to Theorem 8.5,
D (f, P2 ) D (f, P2 ) < .
2
Thinking of P as the generic partition, the interiors of its intervals (xi1 , xi )
may or may not contain points of P1 . For 1 i n, let
Qi = {xi1 , xi } (P1 (xi1 , xi )) part (Ii ) .
If P1 (xi1 , xi ) = , then D (f, P ) and D (f, P2 ) have the term Mi |Ii | in
common because Qi = {xi1 , xi }.
Otherwise, P1 (xi1 , xi ) 6= and
D (f, Qi ) BkP2 k BkP k > B.
Since P1 has m 1 points in (a, b), there are at most m 1 of the Qi not contained
in P .
This leads to the estimate
n
X
D (f, P ) D (f, P2 ) = D (f, P ) D (f, Qi ) < (m 1)2B < .
i=1
4
In the same way,
D (f, P2 ) D (f, P ) < (m 1)2B < .
4
Putting these estimates together yields
D (f, P ) D (f, P ) =
D (f, P ) D (f, P2 ) + D (f, P2 ) D (f, P2 ) + (D (f, P2 ) D (f, P ))
< + + =
4 2 4
This shows that, given > 0, there is a > 0 so that kP k < implies
D (f, P ) D (f, P ) < .
Since
D (f, P ) D (f ) D (f, P ) and D (f, P ) R (f, P, xi ) D (f, P )
for every selection xi from P , it follows that |R (f, P, xi )D (f ) | < when kP k < .
We conclude f is Riemann integrable and R (f ) = D (f ).
From Theorem 8.9, we are justified in using a single notation for both R (f )
Rb Rb
and D (f ). The obvious choice is the familiar a f (x) dx, or, more simply, a f .
When proving statements about the integral, its convenient to switch back and
forth between the Riemann and Darboux formulations. Given f : [a, b] R the
following three facts summarize much of what we know.
Rb
(1) a f exists iff for all > 0 there is a > 0 and an R such that whenever
P part ([a, b]) and xi is a selection from P , then |R (f, P, xi ) | < .
Rb
In this case a f = .
|R (f, Q1 , xk ) R (f, Q2 , yk )|
Z b Z b
R (f, Q1 , xk ) f + f R (f, Q2 , yk ) < .
a a
(=) Let > 0 and choose P part ([a, b]) satisfying (b) with /2 in place of
.
We first claim that f is bounded. To see this, suppose it is not. Then it must be
unbounded on an interval Ik0 determined by P . Fix a selection {xk Ik : 1 k n}
and let yk = xk for k 6= k0 with yk0 any element of Ik0 . Then
> |R (f, P, xk ) R (f, P, yk )| = f (xk0 ) f (yk0 ) |Ik0 | .
2
But, the right-hand side can be made bigger than /2 with an appropriate choice
of yk0 because of the assumption that f is unbounded on Ik0 . This contradiction
forces the conclusion that f is bounded.
Thinking of P as the generic partition and using mk and Mk as usual with
Darboux sums, for each k, choose xk , yk Ik such that
Mk f (xk ) < and f (yk ) mk < .
4n|Ik | 4n|Ik |
D (f, P ) D (f, P )
= D (f, P ) R (f, P, xk ) + R (f, P, xk ) R (f, P, yk ) + R (f, P, yk ) D (f, P )
n
X n
X
= (Mk f (xk ))|Ik | + R (f, P, xk ) R (f, P, yk ) + (f (yk ) mk )|Ik |
k=1
k=1
n
X X n
|Mk f (xk )| |Ik | + |R (f, P, xk ) R (f, P, yk )| + (f (yk ) mk )|Ik |
k=1 k=1
n n
X X
< |Ik | + |R (f, P, xk ) R (f, P, yk )| + |Ik |
4n|Ik | 4n|Ik |
k=1 k=1
< + + <
4 2 4
Corollary 8.7 implies D (f ) exists and Theorem 8.9 finishes the proof.
Rb Rd
Corollary 8.11. If a
f exists and [c, d] [a, b], then c
f exists.
Proof. Let P0 = {a, b, c, d} part ([a, b]) and > 0. Using Theorem 8.10,
choose a partition P such that P0 P and whenever P P and P P 0 , then
|R (f, Q1 , xk ) R (f, Q2 , yk ) | =
|R f, P1 Q1 P3 , xk R f, P1 Q2 P3 , yk | <
Rb
for any selections. An application of Theorem 8.10 shows a
f exists.
Then
Z b X n Z b
R (f, P, xk ) f = f (xk )|Ik | f
a
k=1 a
n Z b
X
= || f (xk )|Ik | f
a
k=1
Z b
= || R (f, P, xk ) f
a
< ||
2||
= .
2
Rb Rb
This shows f is integrable and a f = a f .
Assuming 6= 0, in the same way, we can choose a Pg part ([a, b]) such that
when Pg P , then
Z b
R (g, P, xk ) g < .
a 2||
Rb Rb Rb
for any selection. This shows f +g is integrable and a (f +g) = a f + a g.
Rb Rb 2
(b) Claim: If a h exists, then so does a h
To see this, suppose first that 0 h(x) M on [a, b]. If M = 0, the claim is
trivially true, so suppose M > 0. Let > 0 and choose P part ([a, b]) such that
D (h, P ) D (h, P ) .
2M
Since h 0,
k=1
Xn
= (Mk + mk )(Mk mk )|Ik |
k=1
n
!
X
2M (Mk mk )|Ik |
k=1
= 2M D (h, P ) D (h, P )
< .
2
Therefore, h is integrable when h 0.
If h is not nonnegative, let m = glb {h(x) : a x b}. Then h m 0, and
h m is integrable by (a). From the claim, (h m)2 is integrable. Since
h2 = (h m)2 + 2mh m2 ,
it follows from (a) that h2 is integrable.
Finally, f g = 14 ((f + g)2 (f g)2 ) is integrable by the claim and (a).
(c) Claim: If h 0 is integrable, then so is h.
To see this, let > 0 and choose P part ([a, b]) such that
D (h, P ) D (h, P ) < 2 .
For each 1 k n, let
p p
mk = glb { h(x) : x Ik } lub { h(x) : x Ik } = Mk .
and define
A = {k : Mk mk < } and B = {k : Mk mk }.
Then
X
(70) (Mk mk )|Ik | < (b a).
kA
Proof. Let > 0. According to (a) and Definition 8.1, P part ([a, b]) can
be chosen such that
Z b
R (f, P, xk ) f < .
a
for every selection from P . On each interval [xk1 , xk ] determined by P , the
function F satisfies the conditions of the Mean Value Theorem. (See Corollary 7.13.)
Therefore, for each k, there is an ck (xk1 , xk ) such that
F (xk ) F (xk1 ) = F 0 (ck )(xk xk1 ) = f (ck )|Ik |.
So,
Z Z n
b b X
f (F (b) F (a)) = f (F (xk ) F (xk1 )
a a
k=1
Z
b Xn
= f f (ck )|Ik |
a
k=1
Z
b
= f R (f, P, ck )
a
<
and the theorem follows.
x x+h
Figure
Z 3. This figure illustrates a box argument showing
x+h
1
lim f = f (x).
h0 h x
Rb
Proof. To show F C([a, b]), let x0 [a, b] and > 0. Since a f exists,
there is an M > lub {|f (x)| : a x b}. Choose 0 < < /M and x
(x0 , x0 + ) [a, b]. Then
Z x
|F (x) F (x0 )| = f M |x x0 | < M <
x0
and x0 C(F ).
Let x0 C(f )(a, b) and > 0. There is a > 0 such that x (x0 , x0 +)
(a, b) implies |f (x) f (x0 )| < . If 0 < h < , then
Z x0 +h
F (x0 + h) F (x0 ) 1
f (x ) = f f (x )
0 0
h h
x0
Z
1 x0 +h
= (f (t) f (x0 )) dt
h x0
Z x0 +h
1
|f (t) f (x0 )| dt
h x0
1 x0 +h
Z
< dt
h x0
= .
This shows F+0 (x0 ) = f (x0 ). It can be shown in the same way that F0 (x0 ) = f (x0 ).
Therefore F 0 (x0 ) = f (x0 ).
The right picture makes Theorem 8.16 almost obvious. Consider Figure 3.
Suppose x C(f ) and > 0. There is a > 0 such that
f ((x d, x + d) [a, b]) (f (x) /2, f (x) + /2).
Let
m = glb {f y : |x y| < } lub {f y : |x y| < } = M.
8. Change of Variables
Integration by substitution works side-by-side with the Fundamental Theorem of
Calculus in the integration section of any calculus course. Most of the time calculus
books require all functions in sight to be continuous. In that case, a substitution
theorem is an easy consequence of the Fundamental Theorem and the Chain Rule.
(See Exercise 8.13.) More general statements are true, but they are harder to prove.
Theorem 8.17. If f and g are functions such that
(a) g is strictly monotone on [a, b],
(b) g is continuous on [a, b],
(c) g is differentiable on (a, b), and
R g(b) Rb
(d) both g(a) f and a (f g)g 0 exist,
then
Z g(b) Z b
(73) f= (f g)g 0 .
g(a) a
Proof. Suppositions (a) and (b) show g is a bijection from [a, b] to an interval
[c, d]. The correspondence between the endpoints depends on whether g is increasing
or decreasing.
Let > 0.
From (d) and Definition 8.1, there is a 1 > 0 such that whenever P part ([a, b])
with kP k < 1 , then
Z b
(74) R ((f g)g 0 , P, xi ) (f g)g 0 <
a 2
for any selection from P . Choose P1 part ([a, b]) such that kP1 k < 1 .
Using the same argument, there is a 2 > 0 such that whenever Q part ([c, d])
with kQk < 2 , then
Z d
(75) R (f, Q, xi ) f <
c 2
for any selection from Q. As above, choose Q1 part ([c, d]) such that kQ1 k < 2 .
Setting P2 = P1 {g 1 (x) : x Q1 } and Q2 = P1 {g(x) : x P1 }, it is
apparent that P1 P2 , Q1 Q2 , kP2 k kP1 k < 1 , kQ2 k kQ1 k < 2 and
Q2 = {g(x) : x P2 }. From (74) and (75), it follows that
Z
b
(f g)g 0 R ((f g)g 0 , P2 , xi ) <
a 2
(76) and
Z
d
f R (f, Q2 , yi ) <
2
c
for any selections from P2 and Q2 .
Label the points of P2 as a = x1 < x2 < < xn = b and those of Q2 as
c = y0 < y1 < < yn = d. From (b), (c) and the Mean Value Theorem, for each i,
choose ci (xi1 , xi ) such that
(77) g(xi ) g(xi1 ) = g 0 (ci )(xi xi1 ).
Apply the Mean Value Theorem, as in (77), and then use (76).
n Z b
X
= + f (g(ci ))g 0 (ci ) (xi xi1 ) (f g)g 0
2 i=1 a
Z b
= + R ((f g)g 0 , P2 , ci ) (f g)g 0
2 a
< +
2 2
=
and (73) follows.
Now assume g is strictly decreasing on [a, b]. The proof is much the same as above,
except the bookkeeping is trickier because order is reversed instead of preserved
by g. This means g(xk ) = ynk when 0 k n and g(cnk+1 ) (yk1 , yk ) for
0 < k n. Therefore, g(cnk+1 ) is a selection from Q2 . From the Mean Value
Theorem,
yk yk1 = g(xnk ) g(xnk+1 )
(78) = (g(xnk+1 ) g(xnk ))
= g 0 (cnk+1 )(xnk+1 xnk ),
where ck (xk1 , xk ) is as above. The rest of the proof is much like the case when
g is increasing.
Z
g(b) Z b
0
f (f g)g
g(a) a
Z
g(a) Z b
= f + R (f, Q2 , g(cnk+1 )) R (f, Q2 , g(cnk+1 )) (f g)g 0
g(b) a
Z
g(b) Z b
0
f + R (f, Q2 , g(cnk+1 )) + R (f, Q2 , g(cnk+1 )) (f g)g
g(a) a
Use (76), expand the second Riemann sum and apply (78).
n Z b
X
< + f (g(cnk+1 ))(yk yk1 ) (f g)g 0
2 a
k=1
n Z b
X 0
0
= + f (g(cnk+1 ))g (cnk+1 )(xnk+1 xnk ) (f g)|g |
2 a
k=1
Proof. Obviously,
Z b Z b Z b
(79) m g fg M g.
a a a
Rb
If a
g = 0, were done. Otherwise, let
Rb
fg
c = Ra b .
a
g
Rb Rb
Then a
fg = c a
g and from (79), it follows that m c M .
Proof. This follows from Theorem 8.18 and Corollaries 6.23 and 6.26.
10. Exercises
8.2. Let (
1, xQ
f (x) = .
0, x
/Q
(a) Use Definition 8.1 to show f is not integrable on any interval.
(b) Use Definition 8.6 to show f is not integrable on any interval.
R5
8.3. Calculate 2
x2 using the definition of integration.
8.6. If f : [a, b] [0, ) is continuous and D(f ) = 0, then f (x) = 0 for all
x [a, b].
Rb Rb Rb
8.7. If a
f exists, then limxa x
f= a
f.
Rb
8.8. If f is monotone on [a, b], then a
f exists.
x
dt
Z
8.11. If f (x) = for x > 0, then f (xy) = f (x) + f (y) for x, y > 0.
1 t
3This is variously called the Cauchy inequality, Cauchy-Schwarz inequality, or the Cauchy-
Schwarz-Bunyakowsky inequality. Rearranging the last one, some people now call it the CBS
inequality.
8.13. In the statement of Theorem 8.17, make the additional assumptions that f
and g 0 are both continuous. Use the Fundamental Theorem of Calculus to give an
easier proof.
Sequences of Functions
1. Pointwise Convergence
We have accumulated much experience working with sequences of numbers. The
next level of complexity is sequences of functions. This chapter explores several ways
that sequences of functions can converge to another function. The basic starting
point is contained in the following definitions.
Definition 9.1. Suppose S R and for each n N there is a function
fn : S R. The collection {fn : n N} is a sequence of functions defined on S.
For each fixed x S, fn (x) is a sequence of numbers, and it makes sense to ask
whether this sequence converges. If fn (x) converges for each x S, a new function
f : S R is defined by
f (x) = lim fn (x).
n
The function f is called the pointwise limit of the sequence fn , or, equivalently, it is
S
said fn converges pointwise to f . This is abbreviated fn f , or simply fn f , if
the domain is clear from the context.
Example 9.1. Let
0,
x<0
fn (x) = xn , 0 x < 1 .
1, x1
Then fn f where (
0, x<1
f (x) = .
1, x1
(See Figure 1.) This example shows that a pointwise limit of continuous functions
need not be continuous.
9-1
9-2 CHAPTER 9. SEQUENCES OF FUNCTIONS
1.0
0.8
0.6
0.4
0.2
Figure 1. The first ten functions from the sequence of Example 9.1.
0.4
0.2
-3 -2 -1 1 2 3
-0.2
-0.4
Figure 2. The first four functions from the sequence of Example 9.2.
To figure out what this looks like, it might help to look at Figure 3.
The graph of fn is a piecewise linear function supported on [1/2n+1 , 1/2n ] and
the
R 1 area under the isosceles triangle of the graph over this interval is 1. Therefore,
f = 1 for all n.
0 n
If x > 0, then whenever x > 1/2n , we have fn (x) = 0. From this it follows that
fn 0.
The Rlesson to
R 1be learned from this example is that it may not be true that
1
limn 0 fn = 0 limn fn .
64
32
16
8
1 1 1 1
1
16 8 4 2
Figure 3. The first four functions fn 0 from the sequence of Example 9.3.
1/2
1/4
1/8
1 1
Figure 4. The first ten functions of the sequence fn |x| from
Example 9.4.
(See Figure 4.) The parabolic section in the center was chosen so fn (1/n) = 1/n
and fn0 (1/n) = 1. This splices the sections together at (1/n, 1/n) so fn is
differentiable everywhere. Its clear fn |x|, which is not differentiable at 0.
This example shows that the limit of differentiable functions need not be
differentiable.
The examples given above show that continuity, integrability and differentiability
are not preserved in the pointwise limit of a sequence of functions. To have any
hope of preserving these properties, a stronger form of convergence is needed.
2. Uniform Convergence
Definition 9.2. The sequence fn : S R converges uniformly to f : S R
on S, if for each > 0 there is an N N so that whenever n N and x S, then
|fn (x) f (x)| < .
S
In this case, we write fn f , or simply fn f , if the set S is clear from the
context.
f(x) +
f(x)
fn(x)
f(x
a b
The difference between pointwise and uniform convergence is that with pointwise
convergence, the convergence of fn to f can vary in speed at each point of S. With
uniform convergence, the speed of convergence is roughly the same all across S.
Uniform convergence is a stronger condition to place on the sequence fn than
pointwise convergence in the sense of the following theorem.
S S
Theorem 9.3. If fn f , then fn f .
Proof. Let x0 S and > 0. There is an N N such that when n N , then
|f (x) fn (x)| < for all x S. In particular, |f (x0 ) fn (x0 )| < when n N .
This shows fn (x0 ) f (x0 ). Since x0 S is arbitrary, it follows that fn f .
The first three examples given above show the converse to Theorem 9.3 is false.
There is, however, one interesting and useful case in which a partial converse is true.
S
Definition 9.4. If fn f and fn (x) f (x) for all x S, then fn increases to
S
f on S. If fn f and fn (x) f (x) for all x S, then fn decreases to f on S. In
either case, fn is said to converge to f monotonically.
The functions of Example 9.4 decrease to |x|. Notice that in this case, the
convergence is also happens to be uniform. The following theorem shows Example
9.4 to be an instance of a more general phenomenon.
Theorem 9.5 (Dinis Theorem). If
(a) S is compact,
S
(b) fn f monotonically,
(c) fn C(S) for all n N, and
(d) f C(S),
then fn f .
fn (x) = xn
1/2
2 1/n 1
Figure 6. This shows a typical function from the sequence of Example 9.5.
This shows that when n N , then kfn f k . We conclude that f B(S) and
fn f .
A collection of functions S is said to be complete under uniform convergence,
if every Cauchy sequence in S converges to a function in S. Theorem 9.7 shows
B(S) is complete under uniform convergence. Well see several other collections of
functions that are complete under uniform convergence.
Example 9.6. For S R let L(S) be all the functions f : S R such that
f (x) = mx + b for some constants m and b. In particular, let fn be a Cauchy
sequence in L([0, 1]). Theorem 9.7 shows there is an f : [0, 1] R such that fn f .
In order to show L([0, 1]) is complete, it suffices to show f L([0, 1]).
To do this, let fn (x) = mn x + bn for each n. Then fn (0) = bn f (0) and
mn = fn (1) bn f (1) f (0).
Given any x [0, 1],
fn (x) ((f (1) f (0))x + f (0)) = mn x + bn ((f (1) f (0))x + f (0))
= (mn (f (1) f (0)))x + bn f (0) 0.
This shows f (x) = (f (1) f (0))x + f (0) L([0, 1]) and therefore L([0, 1]) is
complete.
1Definition 2.10
4. Series of Functions
The definitions of pointwiseP
and uniform convergence are extended in the natural
way to series of functions. If k=1 fk is a series of functions defined on a set S,
then the series converges
Pnpointwise or uniformly, depending on whether the sequence
of partial sums, sn = k=1 fk converges pointwise or uniformly,
P respectively. It is
absolutely convergent or absolutely uniformly convergent, if n=1 |fn | is convergent
or uniformly convergent on S, respectively.
The following theorem is obvious and its proof is left to the reader.
P P
Theorem 9.8. Let n=1 fn be a series of functions
P defined on S. If n=1 fn
is absolutely convergent, then it is convergent. If n=1 fn is absolutely uniformly
convergent, then it is uniformly convergent.
This shows sn is a Cauchy sequence and must converge according to Theorem 9.7.
Therefore, f is continuous at x0 .
Example 9.1 shows that continuity is not preserved under pointwise convergence.
Corollary 9.12 establishes that if S is compact, then C(S) is complete under the
uniform metric.
The fact that C([a, b]) is closed under uniform convergence is often useful
because, given a bad function f C([a, b]), its often possible to find a sequence
fn of good functions in C([a, b]) converging uniformly to f . Following is the most
widely used theorem of this type.
Proof. Parts (a) and (b) follow easily from the definition of kn .
To prove (c) first note that
Z 1 Z 1/n n
2 n 2 1
1= kn
cn (1 t ) dt cn 1 .
1 1/ n n n
1.2
1.0
0.8
0.6
0.4
0.2
Corollary 9.15. If f C([a, b]) and > 0, then there is a polynomial p such
that kf pk[a,b] < .
The theorems of this section can also be used to construct some striking examples
of functions with unwelcome behavior. Following is perhaps the most famous.
Example 9.9. There is a continuous f : R R that is differentiable nowhere.
Proof. Thinking of the canonical example of a continuous function that fails to
be differentiable at a pointthe absolute value functionwe start with a sawtooth
function. (See Figure 8.)
(
x 2n, 2n x < 2n + 1, n Z
s0 (x) =
2n + 2 x, 2n + 1 x < 2n + 2, n Z
Notice that s0 is continuous and periodic with period 2 and maximum value 1.
Compress it both vertically and horizontally:
n
3
sn (x) = s0 (4n x) , n N.
4
Each sn is continuous and periodic with period pn = 2/4n and ksn k = (3/4)n .
1.0
0.8
0.6
0.4
0.2
Since ksn k = (3/4)n , the Weierstrass M -test implies the series defining f is
uniformly convergent and Corollary 9.12 shows f is continuous on R. We will show
f is differentiable nowhere.
Let x R, m N and hm = 1/(2 4m ).
If n > m, then hm /pn = 4nm1 N, so sn (x hm ) sn (x) = 0 and
m
f (x hm ) f (x) X sk (x hm ) sk (x)
(85) = .
hm hm
k=0
On the other hand, if n < m, then a worst-case estimate is that
n
sn (x hm ) sn (x)
3 /
1
= 3n .
hm 4 4n
3.0
2.5
2.0
1.5
1.0
0.5
This gives
m1 m1
X sk (x hm ) sk (x) X sk (x hm ) sk (x)
hm hm
k=0 k=0
3m 1 3m
(86) < .
31 2
Since sm is linear on intervals of length 4m = 2 hm with slope 3m on those
linear segments, at least one of the following is true:
sm (x + hm ) s(x)
= 3m or sm (x hm ) s(x) = 3m .
(87)
hm hm
Suppose the first of these is true. The argument is essentially the same in the second
case.
Using (85), (86) and (87), the following estimate ensues
f (x + hm ) f (x) X sk (x + hm ) sk (x)
=
hm hm
k=0
m
X sk (x + hm ) sk (x)
=
hm
k=0
m1
sm (x + hm ) sm (x) X sk (x hm ) sk (x)
hm hm
k=0
m 3m 3m
>3 = .
2 2
Since 3m /2 , it is apparent f 0 (x) does not exist.
There are many other constructions of nowhere differentiable continuous func-
tions. The first was published by Weierstrass [19] in 1872, although it was known
in the folklore sense among mathematicians earlier than this. (There is an English
translation of Weierstrass paper in [9].) In fact, it is now known in a technical
sense that the typical continuous function is nowhere differentiable [4].
This is often referred to as passing the limit through the integral. At some point
in her career, any student of advanced analysis or probability theory will be tempted
to just blithely pass the limit through. But functions such as those of Example
9.3 show that some care is needed. A common criterion for doing so is uniform
convergence.
Rb
Theorem 9.16. If fn : [a, b] R such that a fn exists for each n and fn f
on [a, b], then
Z b Z b
f = lim fn
a n a
.
Proof. Some care must be taken in this proof, because there are actually two
things to prove. Before the equality can be shown, it must be proved that f is
integrable.
To show that f is integrable, let > 0 and N N such that kf fN k < /3(ba).
If P part ([a, b]), then
n
X n
X
(88) |R (f, P, xk ) R (fN , P, xk ) | = | f (xk )|Ik | fN (xk )|Ik ||
k=1 k=1
N
X
=| (f (xk ) fN (xk ))|Ik ||
k=1
N
X
|f (xk ) fN (xk )||Ik |
k=1
n
X
< |Ik |
3(b a))
k=1
=
3
According to Theorem 8.10, there is a P part ([a, b]) such that whenever
P Q1 and P Q2 , then
(89) |R (fN , Q1 , xk ) R (fN , Q2 , yk ) | < .
3
Combining (88) and (89) yields
|R (f, Q1 , xk ) R (f, Q2 , yk )|
= |R (f, Q1 , xk ) R (fN , Q1 , xk ) + R (fN , Q1 , xk )
R (fN , Q1 , xk ) + R (fN , Q2 , yk ) R (f, Q2 , yk )|
|R (f, Q1 , xk ) R (fN , Q1 , xk )| + |R (fN , Q1 , xk ) R (fN , Q1 , xk )|
+ |R (fN , Q2 , yk ) R (f, Q2 , yk )|
< + + =
3 3 3
Another application of Theorem 8.10 shows that f is integrable.
Finally, when n N ,
Z Z b Z b Z
b b
f fn = (f fn ) < = <
3(b a) 3
a a a a
Rb Rb
shows that a fn a f .
P
Corollary 9.17. If n=1 fn is a series of integrable functions converging
uniformly on [a, b], then
Z bX X Z b
fn = fn
a n=1 n=1 a
Example 9.10. It was shown in Example 4.2 that the geometric series
X 1
tn = , 1 < t < 1.
n=0
1t
In Exercise 9.3, you are asked to prove this convergence is uniform on any compact
subset of (1, 1). Substituting t for t in the above formula, it follows that
X 1
(t)n
n=0
1+t
Let M N so that
m, n M = kfm fn k < .
3
Notice this implies
(91) kf fn k , n M.
3
For such m, n M and x, y [a, b] with x = 6 y, another application of the
Mean Value Theorem gives
Fn (x) Fn (y) Fm (x) Fm (y)
xy xy
1
= |(Fn (x) Fm (x)) (Fn (y) Fm (y))|
|x y|
1
= |fn (c) fm (c)| |x y| kfn fm k < .
|x y| 3
Letting m , it follows that
Fn (x) Fn (y) F (x) F (y)
(92) , n M.
xy xy 3
Fix n M and x [a, b]. Since Fn0 (x) = fn (x), there is a > 0 so that
Fn (x) Fn (y)
(93) fn (x) < , y (x , x + ) \ {x}.
xy 3
Finally, using (92), (93) and (91), we see
F (x) F (y)
f (x)
xy
F (x) F (y) Fn (x) Fn (y)
=
xy xy
Fn (x) Fn (y)
+ fn (x) + fn (x) f (x)
xy
F (x) F (y) Fn (x) Fn (y)
xy xy
Fn (x) Fn (y)
+ fn (x) + |fn (x) f (x)|
xy
< + + = .
3 3 3
This establishes that
F (x) F (y)
lim = f (x),
yx xy
as desired.
Corollary 9.20. If Gn C([a, b]) is a sequence such that G0n g and Gn (x0 )
converges for some x0 [a, b], then Gn G where G0 = g.
Proof. Suppose Gn (x0 ) . For each n choose an antiderivative Fn of gn
such that Fn (a) = 0. Theorem 9.19 shows g is a derivative and an argument similar
to that in the proof of Theorem 9.19 shows Fn F on [a, b], where F 0 = g. Since
Fn0 G0n = 0, Corollary (7.15) shows Gn (x) = Fn (x) + (Gn (x0 ) Fn (x0 )). Define
G(x) = F (x) + ( F (x0 )).
Let > 0 and x [a, b]. There is an N N such that
n N = kFn F k < and |Gn (x0 ) | < .
3 3
If n N ,
|Gn (x) G(x)| = |Fn (x) + (Gn (x0 ) Fn (x0 )) (F (x) + ( F (x0 )))|
|Fn (x) F (x)| + |Gn (x0 ) | + |Fn (x0 ) F (x0 )|
< + +
3 3 3
=
This shows Gn G on [a, b] where G0 = F 0 = g.
8. Power Series
8.1. The Radius and Interval of Convergence. One place where uniform
convergence plays a key role is with power series. Recall the definition.
Definition 9.22. A power series is a function of the form
X
(95) f (x) = an (x c)n .
n=0
Members of the sequence an are the coefficients of the series. The domain of f is
the set of all x at which the series converges. The constant c is called the center of
the series.
To determine the domain of (95), let x R \ {c} and use the root test to see
the series converges when
lim sup |an (x c)n |1/n = |x c| lim sup |an |1/n < 1
and diverges when
|x c| lim sup |an |1/n > 1.
If r lim sup |an |1/n 1 for some r 0, then these inequalities imply (95) is absolutely
convergent when |x c| < r. In other words, if
(96) R = lub {r : r lim sup |an |1/n < 1},
then the domain of (95) is an interval of radius R centered at c. The root test gives
no information about convergence when |x c| = R. This R is called the radius
of convergence of the power series. Assuming R > 0, the open interval centered
at c with radius R is called the interval of convergence. It may be different from
the domain of the series because the series may converge at one endpoint or both
endpoints of the interval of convergence.
The ratio test can also be used to determine the radius of convergence, but, as
shown in (30), it will not work as often as the root test. When it does,
an+1
(97) R = lub {r : r lim
< 1}.
n an
This is usually easier to compute than (96), and both will give the same value for R.
Example 9.12. Calling to mind Example 4.2, it is apparent the geometric
P
power series n=0 xn has center 0, radius of convergence 1 and domain (1, 1).
P
Example 9.13. For the power series n=1 2n (x + 2)n /n, we compute
n 1/n
2 1
lim sup = 2 = R = .
n 2
Since the series diverges when x = 2 12 , it follows that the interval of convergence
is (5/2, 3/2).
P
Example 9.14. The power series n=1 xn /n has interval of convergence (1, 1)
and domain [1, 1). Notice it is not absolutely convergent when x = 1.
P n 2
Example 9.15. The power series n=1 x /n has interval of convergence
(1, 1), domain [1, 1] and is absolutely convergent on its whole domain.
The preceding is summarized in the following theorem.
Theorem 9.23. Let the power series be as in (95) and R be given by either
(96) or (97).
(a) If R = 0, then the domain of the series is {c}.
(b) If R > 0 the series converges absolutely at x when |c x| < R and
diverges at x when |c x| > R. In the case when R = , the series
converges everywhere.
(c) If R (0, ), then the series may converge at none, one or both of
c R and c + R.
The latter series satisfies the Alternating Series Test. Since 1 5/(15 15!)
1.5 106 , Corollary 4.20 shows
Z 6
X (1)n
f (x) dx 2n+1 1.85194
0 n=0
(2n + 1)(2n + 1)!
The series (101) is called the Taylor series 5 for f centered at c. The Taylor
series can be formally defined for any function that has derivatives of all orders at c,
but, as Example 7.9 shows, there is no guarantee it will converge to the function
anywhere except at c. Taylors Theorem 7.17 can be used to examine the question
5When c = 0, it is often called the Maclaurin series for f .
Let > 0. Choose N N such that whenever n N , then |sn s| < /2. Choose
(0, 1) so
N
X
|sk s| < /2.
k=0
Suppose x is such that 1 < x < 1. With these choices, (103) becomes
XN X
|f (x) s| (1 x) (sk s)xk + (1 x) (sk s)xk
k=0 k=N +1
N
X X
< |sk s| + (1 x) xk < + =
2 2 2
k=0 k=N +1
n 0 and
(a) naP
(b) A n=0 an = A,
P
then n=0 an =A.
Pn
Proof. Let sn = k=0 ak . For x (0, 1) and n N,
X X n Xn X
sn ak xk = ak ak xk ak xk
k=0 k=0 k=0 k=n+1
n
X X
k
= ak (1 x ) ak xk
k=0 k=n+1
n
X X
k1 k
= ak (1 x)(1 + x + + x ) ak x
k=0 k=n+1
n
X
X
(105) (1 x) k|ak | + |ak |xk .
k=0 k=n+1
Let > 0. According to (a) and Exercise 3.19, there is an N N such that
n
1X
(106) n N = n|an | < and k|ak | < .
2 n 2
k=0
Let n N and 1 1/n < x < 1. Using the right term in (106),
n n
X 1X
(107) (1 x) k|ak | < k|ak | < .
n 2
k=0 k=0
9. Exercises
9.1. If fn (x) = nx(1 x)n for 0 x 1, then show fn converges pointwise, but
not uniformly on [0, 1].
9.2. Show sinn x converges uniformly on [0, a] for all a (0, /2). Does sinn x
converge uniformly on [0, /2)?
X 1
9.16. Is Abel convergent?
n=2
n ln n
Fourier Series
to approximate his solutions. Series of this form are called trigonometric series,
and the ones derived from Fouriers methods are called Fourier series. Much of
the mathematical research done in the nineteenth and early twentieth century was
devoted to understanding the convergence of Fourier series. This chapter presents
nothing more than the tip of that huge iceberg.
1. Trigonometric Polynomials
Definition 10.1. A function of the form
n
X
(111) p(x) = k cos kx + k sin kx
k=0
1
Fouriers methods can be seen in most books on partial differential equations, such as [3]. For
example, see solutions of the heat and wave equations using the method of separation of variables.
10-1
10-2 CHAPTER 10. FOURIER SERIES
Z 0,
m 6= n
(113) sin mx sin nx dx = 0, m = 0 or n = 0
m = n 6= 0
and
Z 0,
m 6= n
(114) cos mx cos nx dx = 2 m=n=0.
m = n 6= 0
(Remember that all but a finite number of the an and bn are 0!)
At this point, the logical question is whether this same method can be used to
represent a more general 2-periodic function. For any function f , integrable on
[, ], the coefficients can be defined as above; i.e., for n ,
1 1
Z Z
(116) an = f (x) cos nx dx and bn = f (x) sin nx dx.
The numbers an and bn are called the Fourier coefficients of f . The problem is
whether and in what sense an equation such as (115) might be true. This turns out
to be a very deep and difficult question with no short answer.2 Because we dont
know whether equality in the sense of (115) is true, the usual practice is to write
a0 X
(117) f (x) + an cos nx + bn sin nx,
2 n=1
indicating that the series on the right is calculated from the function on the left
using (116). The series is called the Fourier series for f .
2Many people, including me, would argue that the study of Fourier series has been the
most important area of mathematical research over the past two centuries. Huge mathematical
disciplines, including set theory, measure theory and harmonic analysis trace their lineage back to
basic questions about Fourier series. Even after centuries of study, research in this area continues
unabated.
Figure 1. This shows f (x) = |x|, s1 (x) and s3 (x), where sn (x) is
the nth partial sum of the Fourier series for f .
Example 10.1. Let f (x) = |x|. Since f is an even functions and sin nx is odd,
1
Z
bn = |x| cos nx dx = 0
for all n N. On the other hand,
1
Z , n=0
an = |x| cos nx dx = 2(cos n 1)
, nN
n2
for n Z. Therefore,
4 cos x 4 cos 3x 4 cos 5x 4 cos 7x 4 cos 9x
|x| +
2 9 25 49 81
(See Figure 10.1.)
There are at least two fundamental questions arising from (117): Does the
Fourier series of f converge to f ? Can f be recovered from its Fourier series, even
if the Fourier series does not converge to f ? These are often called the convergence
and representation questions, respectively. The next few sections will give some
partial answers.
Proof. Since the two limits have similar proofs, only the first will be proved.
Let > 0 and P be a generic partition of [a, b] satisfying
Z b
0< f D (f, P ) < .
a 2
For mi = glb {f (x) : xi1 < x < xi }, define a function g on [a, b] by g(x) = mi
Rb
when xi1 x < xi and g(b) = mn . Note that a g = D (f, P ), so
Z b
(118) 0< (f g) < .
a 2
Choose
n
4X
(119) > |mi |.
i=1
Since f g,
Z Z
b b Z b
f (t) cos t dt = (f (t) g(t)) cos t dt + g(t) cos t dt
a a a
Z Z
b b
(f (t) g(t)) cos t dt + g(t) cos t dt
a a
Z b 1 Xn
(f g) + mi (sin(xi ) sin(xi1 ))
a
i=1
Z b n
2X
(f g) + |mi |
a i=1
15
12
- -
2 2
-3
n
a0 X
sn (x) = + (ak cos kx + bk sin kx)
2
k=1
Z n
1 1
X Z
= f (t) dt + (f (t) cos kt cos kx + f (t) sin kt sin kx) dt
2
k=1
n
Z !
1 X
= f (t) 1 + 2(cos kt cos kx + sin kt sin kx) dt
2
k=1
n
Z !
1 X
= f (t) 1 + 2 cos k(x t) dt
2
k=1
sin(n + 1/2)s
(e) Dn (s) = for each n N and s/2 not an integer multiple
sin s/2
of .
Proof. Properties (a)(d) follow from the definition of the kernel.
The proof of property (e) uses some trigonometric manipulation. Suppose n N
and s 6= m for any m Z.
n
X
Dn (s) = 1 + 2 cos ks
k=1
Use the facts that the cosine is even and the sine is odd.
n n
X cos 2s X
= cos ks + sin ks
sin 2s
k=n k=n
n
1 X s s
= sin cos ks + cos sin ks
sin 2s 2 2
k=n
n
1 X 1
= sin(k + )s
sin 2s 2
k=n
Figure 3. This graph shows D50 (t) and the envelope y = 1/ sin(t/2).
As n gets larger, the Dn (t) fills the envelope more completely.
This is similar to a situation weve seen before within the proof of the Weierstrass
approximation theorem, Theorem 9.13. The integral given above is a convolution
integral similar to that used in the proof of Theorem 9.13, although the Dirichlet
kernel isnt a convolution kernel in the sense of Lemma 9.14 because it doesnt
satisfy conditions (a) and (c) of that lemma. (See Figure 10.3.)
- - 2
2 2
-
2
Figure 4. This plot shows the function of Example 10.2 and s8 (x)
for that function.
In particular, when x = /2, (122) gives another way to derive (104). When x = ,
the series converges to 0, which is the middle of the jump for f .
This behavior of converging to the middle of a jump discontinuity is typical. To
see this, denote the one-sided limits of f at x by
f (x) = lim f (t) and f (x+) = lim f (t),
tx tx
The Dini test given above provides a powerful condition sufficient to ensure the
pointwise convergence of a Fourier series. There is a plethora of ever more abstruse
conditions that can be proved in a similar fashion to show pointwise convergence.
The problem is complicated by the fact that there are continuous functions
with Fourier series divergent at a point and integrable functions with Fourier series
diverging everywhere [12]. In Section 6, a continuous function whose Fourier series
diverges is constructed.
5. Gibbs Phenomenon
For x [, ) define
(
|x|
(123) f (x) = x , 0<x<
0, x = 0,
and extend f 2-periodically to all of R. This function is often called a square wave.
A straightforward calculation gives
4 X sin(2k 1)x
f .
2k 1
k=1
Corollary 10.7 shows sn (x) f (x) everywhere. This convergence cannot be uniform
because all the partial sums are continuous and f is discontinuous at every integer
multiple of . A plot of s19 (x) is shown in Figure 5. Notice the higher peaks in
the oscillation of sn (x) just before and after the jump discontinuities of f . This
behavior is not unique to f , as it can also be seen in Figure 4. If a function is
discontinuous at some point, the partial sums of its Fourier series will always have
such higher peaks near that point. This behavior is called Gibbs phenomenon.3
Instead of doing a general analysis of Gibbs phenomenon, well only analyze
the simple case shown in the square wave f . Its basically a calculus exercise.
3It is named after the American mathematical physicist, J. W. Gibbs, who pointed it out
in 1899. He was not the first to notice the phenomenon, as the British mathematician Henry
Wilbraham had published a little-noticed paper on it it 1848. Gibbs interest in the phenomenon
was sparked by investigations of the American experimental physicist A. A. Michelson who wrote
a letter to Nature disputing the possibility that a Fourier series could converge to a discontinuous
function. The ensuing imbroglio is recounted in a marvelous book by Paul J. Nahin [15].
The last sum is a midpoint Riemann sum for the function sinx x on the interval [0, ]
using a regular partition of n subintervals. Example 9.16 shows
2 sin x
Z
dx 1.17898.
0 x
Since f (0+) f (0) = 2, this is an overshoot of a bit less than 9%. There is a
similar undershoot on the other side. It turns out this is typical behavior at points
of discontinuity [20].
Proof.
n n
X 1 X 1
kt = t sin sin kt
sin 2 k=m 2
k=m
n
1 X 1 1
= cos k t cos k + t
sin 2t k=m 2 2
cos m 12 t cos n + 12 t
=
2 sin 2t
Well not have much use for the conjugate Dirichlet kernel, except as a convenient
way to refer to sums of the form (124).
Lemma 10.8 immediately gives the following bound.
Proof.
n n
X sin kt X 1
= Dk (t) Dk1 (t)
k k
k=m k=m
4In this case, the word conjugate does not refer to the complex conjugate, but to the
harmonic conjugate. They are related by Dn (t) + iDn (t) = 1 + 2 n kit .
P
k=1 e
Proof.
n
X sin kt X sin kt X sin kt
= +
k k k
k=1 1k 1t 1
t <kn
X X
sin kt
sin kt
+
1k 1 k 1 <kn k
t t
X kt 1
+ 1 t
1
k t sin 2
1k t
X 1
t+ 1 2 t
1k 1t t 2
1+
This closed form for fn combined with Proposition 10.12 implies the sequence of
functions fn is uniformly bounded:
n
X sin kt
(126) |fn (t)| = 2 sin nt 2 + 2.
k
k=1
The Weierstrass M-Test along with (126) implies F is uniformly convergent and
therefore continuous on R. Using (125), consider
Z
1
s2m (F, 0) =
3 F (t)D2m3 (t) dt
2
Z X
1 f2n3 (t)
= D2m3 (t) dt
2 n=1 n2
Use (125).
X 1
= s 3 f2n3 , 0
2 2m
n=m
n
1
> 2 s2m3 f2m3 , 0
m
m3
2
1 X1
= 2
m k
k=1
1 3
> 2 ln 2m
m
= m ln 2
This implies, lim sup sn (F, 0) limm m ln 2 = , so sn (F, 0) does not converge.
The trigonometric polynomials n (f, x) are called the Cesaro means of the partial
sums. If limn n (f, x) exists, then the Fourier series for f is said to be (C, 1)
summable at x. The idea is that this averaging will smooth out the partial sums,
making them more nicely behaved. It is not hard to show that if sn (f, x) converges
at some x, then n (f, x) will converge to the same thing. But there are sequences
for which n (f, x) converges and sn (f, x) does not. (See Exercises 3.19 and 4.26.)
As with sn (x), well simply write n (x) instead of n (f, x), when it is clear
which function is being considered.
We start with a calculation.
n
1 X
n (x) = sk (x)
n+1
k=0
n Z
1 X 1
= f (x t)Dk (t) dt
n+1 2
k=0
Z n
1 1 X
(*) = f (x t) Dk (t) dt
2 n+1
k=0
Z n
1 1 X sin(k + 1/2)t
= f (x t) dt
2 n+1 sin t/2
k=0
Z n
1 1 X
= f (x t) sin t/2 sin(k + 1/2)t dt
2 (n + 1) sin2 t/2 k=0
10
- -
2 2
Let be as in (e). In light of (a), it suffices to prove (e) for the interval [, ].
Noting that sin t/2 is decreasing on [, ], it follows that for t ,
2
sin n+1
2 t
1
Kn (t) =
(n + 1) sin 2t
2
1 1
(n + 1) sin 2t
1 1
0
(n + 1) sin2 2
It follows that Kn 0 on [, ] and (e) has been proved.
Theorem 9.16 and (e) imply (f).
2
2
- - 2 - - 2
2 2 2 2
- -
2 2
- -
Proof. By assumption, n (f, t) = n (g, t) for all n and Theorem 10.14 implies
0 = n (f, t) n (g, t) f g.
A perspicacious reader will have noticed the similarity between Theorem 10.16
and the Weierstrass approximation theorem, Theorem 9.13. In fact, the Weierstrass
approximation theorem can be proved from Theorem 10.16 using power series and
Theorem 9.24.
Example 10.4. As in Example 10.2, let f (x) = x for < x and extend
f to be periodic on R with period 2. Figure 9 shows the difference between the
Fejer and classical methods of summation. Notice that the Fejer sums remain much
more smoothly affixed to the function and do not show Gibbs phenomenon.
8. Exercises
X
10.5. Is sin nx the Fourier series of some function?
n=1
10.10. Prove
n
X sin 2nt
cos(2k 1)t =
2 sin t
k=1
Pn
for t 6= k, k Z and n N. (Hint: 2 k=1 cos(2k 1)t = D2n (t) Dn (2t).)
[1] A.D. Aczel. The mystery of the Aleph: mathematics, the Kabbalah, and the search for infinity.
Washington Square Press, 2001.
[2] Carl B. Boyer. The History of the Calculus and Its Conceptual Development. Dover Publica-
tions, Inc., 1959.
[3] James Ward Brown and Ruel V Churchill. Fourier series and boundary value problems.
McGraw-Hill, New York, 8th ed edition, 2012.
[4] Andrew Bruckner. Differentiation of Real Functions, volume 5 of CRM Monograph Series.
American Mathematical Society, 1994.
[5] Georg Cantor. Uber eine Eigenschaft des Inbegriffes aller reelen algebraishen Zahlen. Journal
fur die Reine und Angewandte Mathematik, 77:258262, 1874.
[6] Georg Cantor. De la puissance des ensembles parfait de points. Acta Math., 4:381392, 1884.
[7] Lennart Carleson. On convergence and growth of partial sums of Fourier series. Acta Math.,
116:135157, 1966.
[8] Krzysztof Ciesielski. Set Theory for the Working Mathematician. Number 39 in London
Mathematical Society Student Texts. Cambridge University Press, 1998.
[9] Gerald A. Edgar, editor. Classics on Fractals. Addison-Wesley, 1993.
[10] James Foran. Fundamentals of Real Analysis. Marcel-Dekker, 1991.
[11] Nathan Jacobson. Basic Algebra I. W. H. Freeman and Company, 1974.
[12] Yitzhak Katznelson. An Introduction to Harmonic Analysis. Cambridge University Press,
Cambridge, UK, 3rd ed edition, 2004.
[13] Charles C. Mumma. n! and The Root Test. Amer. Math. Monthly, 93(7):561, August-
September 1986.
[14] James R. Munkres. Topology; a first course. Prentice-Hall, Englewood Cliffs, N.J., 1975.
[15] Paul J Nahin. Dr. Eulers Fabulous Formula: Cures Many Mathematical Ills. Princeton
University Press, Princeton, NJ, 2006.
[16] Paul du Bois Reymond. Ueber die Fourierschen Reihen. Nachr. Kon. Ges. Wiss. Gttingen,
21:571582, 1873.
[17] Hans Samelson. More on Kummers test. The American Mathematical Monthly, 102(9):817
818, 1995.
[18] Jingcheng Tong. Kummers test gives characterizations for convergence or divergence of all
positive series. The American Mathematical Monthly, 101(5):450452, 1994.
[19] Karl Weierstrass. Uber continuirliche Functionen eines reelen Arguments, di fur keinen Werth
des Letzteren einen bestimmten Differentialquotienten besitzen, volume 2 of Mathematische
werke von Karl Weierstrass, pages 7174. Mayer & Muller, Berlin, 1895.
[20] Antoni Zygmund. Trigonometric Series. Cambridge University Press, Cambridge, UK, 3rd ed.
edition, 2002.
A-1
Index
A
P
n=1 , Abel sum, 9-21 , symmetric difference, 1-3
0 , cardinality of N, 1-10 , product (Cartesian or real), 1-5, 2-1
S c , complement of S, 1-3 T , trigonometric polynomials, 10-1
c, cardinality of R, 2-10 uniform convergence, 9-4
D (.) Darboux integral, 8-5 , union, 1-2
D (.) . lower Darboux integral, 8-5 Z, integers, 1-2
D (., .) lower Darboux sum, 8-4 Abels test, 4-10
D (.) . upper Darboux integral, 8-5 absolute value, 2-4
D (., .) upper Darboux sum, 8-4 almost every, 5-9
\, set difference, 1-3 alternating harmonic series, 4-10
Dn (t) Dirichlet kernel, 10-5 Alternating Series Test, 4-12
Dn (t) conjugate Dirchlet kernel, 10-11 and , 1-2
, element, 1-1 Archimedean Principle, 2-7
6, not an element, 1-1 axioms of R
, empty set, 1-2 additive inverse, 2-1
, logically equivalent, 1-4 associative laws, 2-1
Kn (t) Fejer kernel, 10-16 commutative laws, 2-1
F , F sigma set, 5-11 completeness, 2-7
B A , all functions f : A B, 1-13 distributive law, 2-1
G , G delta set, 5-11 identities, 2-1
glb , greatest lower bound, 2-6 multiplicative inverse, 2-1
order, 2-3
iff, if and only if, 1-4
= , implies, 1-4
Baire category theorem, 5-10
<, , >, , 2-3
Baire, Rene-Louis, 5-10
, infinity, 2-6
Bertrands test, 4-9
, intersection, 1-3 Bolzano-Weierstrass Theorem, 3-7, 5-3
, logical and, 1-2 bound
, logical or, 1-2 lower, 2-5
lub , least upper bound, 2-6 upper, 2-5
n, initial segment, 1-10 bounded, 2-5
N, natural numbers, 1-2 above, 2-5
, nonnegative integers, 1-2 below, 2-5
part ([a, b]) partitions of [a, b], 8-1
pointwise convergence, 9-1 Cantor, Georg, 1-11
P(A), power set, 1-2 diagonal argument, 2-9
, indexed product, 1-5 middle-thirds set, 5-11
R, real numbers, 2-7 cardinality, 1-10
R (.) Riemann integral, 8-3 countably infinite, 1-10
R (., ., .) Riemann sum, 8-2 finite, 1-10
, subset, 1-1 uncountably infinite, 1-11
(, proper subset, 1-1 Cartesian product, 1-5
, superset, 1-1 Cauchy
), proper superset, 1-1 condensation test, 4-4
A-2
Index A-3
unbounded, 2-6
uniform continuity, 6-12
uniform metric, 9-6
Cauchy sequence, 9-6
complete, 9-6