You are on page 1of 119

MAT 215: Analysis in a single variable

Course notes, Fall 2012


Michael Damron
Compiled from lectures and exercises designed with Mark McConnell
following Principles of Mathematical Analysis, Rudin
Princeton University
1
Contents
1 Fundamentals 4
1.1 Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2 Relations and functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Cardinality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.4 Natural numbers and induction . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.5 Cardinality and the natural numbers . . . . . . . . . . . . . . . . . . . . . . 14
1.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2 The real numbers 18
2.1 Rationals and suprema . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.2 Existence and properties of real numbers . . . . . . . . . . . . . . . . . . . . 19
2.3 R
n
for n 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3 Metric spaces 25
3.1 Denitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.2 Open and closed sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.3 Limit points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.4 Compactness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.5 Heine-Borel Theorem: compactness in R
n
. . . . . . . . . . . . . . . . . . . . 32
3.6 The Cantor set . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.7 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4 Sequences 40
4.1 Denitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
4.2 Subsequences, Cauchy sequences and completeness . . . . . . . . . . . . . . 45
4.3 Special sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
5 Series 52
5.1 Denitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
5.2 Ratio and root tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.3 Non non-negative series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
6 Function limits and continuity 62
6.1 Function limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
6.2 Continuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
6.3 Relations between continuity and compactness . . . . . . . . . . . . . . . . . 66
6.4 Connectedness and the IVT . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
6.5 Discontinuities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
6.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
2
7 Derivatives 74
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
7.2 Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
7.3 Mean value theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
7.4 LHopitals rule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
7.5 Power series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
7.6 Taylors theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
7.7 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
8 Integration 91
8.1 Denitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
8.2 Properties of integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
8.3 Fundamental theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
8.4 Change of variables, integration by parts . . . . . . . . . . . . . . . . . . . . 101
8.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
A Real powers 105
A.1 Natural roots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
A.2 Rational powers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
A.3 Real powers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
B Logarithm and exponential functions 110
B.1 Logarithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
B.2 Exponential function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
B.3 Sophomores dream . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
C Dimension of the Cantor set 115
C.1 Denitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
C.2 The Cantor set . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
C.3 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
3
1 Fundamentals
1.1 Sets
We begin with the concepts of set, object and set membership. We will leave these as primitive
in a sense; that is, undened. You can think of a set as a collection of objects and if a is
an object and A is a set then a A means a is a member of A. If A and B are sets, we
say that A is a subset of B (written A B) whenever a A we have a B. If A B and
B A we say the sets are equal and we write A = B. A is a proper subset of B if A B
but A ,= B. Note that , the set with no elements, is a subset of every set.
There are many operations we can perform with sets.
If A and B are sets, A B is the union of A and B and is the set
A B = a : a A or a B .
If A and B are sets, A B is the intersection of A and B and is the set
A B = a : a A and a B .
Of course we can generalize these to arbitrary numbers of sets. If ( is a (possibly
innite) collection of sets (that is, a set whose elements are themselves sets), we dene
_
AC
A = a : a A for some A (

AC
A = a : a A for all A ( .
The sets A and B are called disjoint if A B = .
These operations obey the following properties
A (B C) = (A B) (A C)
A (B C) = (A B) (A C) .
Let us give a proof of the rst. To show these sets are equal, we must show each is contained
in the other. So let a A (B C). We would like to show that a (A B) (A C).
We know a A and a (B C). One possibility is that a A and a B, in which case
a AB, giving a (AB) (AC). The only other possibility is that a A and a C,
since a must be in either B or C. Then a AC and the same conclusion holds. The other
direction is an exercise.
If A and B are sets then dene the dierence A B as
A B = a : a A but a / B .
4
One can verify the following as well.
A (B C) = (A B) (A C)
A (B C) = (A B) (A C) .
Finally the symmetric dierence is
AB = (A B) (B A) .
1.2 Relations and functions
Our last important way to build a set from other sets is the product. We write
A B = (a, b) : a A, b B .
Denition 1.2.1. A relation R between sets A and B is any subset of AB. If (a, b) R
we think of a as being related to b.
We will rst mention types of relations on a single set.
A relation R between A and A is reexive if (a, a) R for all a A.
It is symmetric if whenever (a
1
, a
2
) R we have (a
2
, a
1
) R.
It is transitive if whenever (a
1
, a
2
) R and (a
2
, a
3
) R we have (a
1
, a
3
) R.
Denition 1.2.2. A relation R on A which is reexive, symmetric and transitive is called
an equivalence relation. Given a A and an equivalence relation R on A we write
[a]
R
= a

A : (a, a

) R
for the equivalence class of a.
Sometimes the condition (a, a

) R is written a a

(and sometimes R is not even


mentioned). An example is equality of sets; that is, dening the relation A B if A = B
gives an equivalence relation. And here we have not specied R or the larger set on which
R is a relation. You can check set equality is reexive, symmetric and transitive.
Proposition 1.2.3. If R is an equivalence relation on a nonempty set A and a
1
, a
2
R
then
either [a
1
]
R
= [a
2
]
R
or [a
1
]
R
[a
2
]
R
= .
Proof. We will rst show that both conditions cannot simultaneously hold. Then we will
show that at least one must hold. To show the rst, note that a
1
[a
1
]
R
and a
2
[a
2
]
R
since R is reexive. Therefore if [a
1
]
R
= [a
2
]
R
then a
1
[a
1
]
R
[a
2
]
R
, giving nonempty
intersection.
For the second claim, suppose that [a
1
]
R
[a
2
]
R
,= , giving some a in the intersection.
We claim that [a]
R
= [a
1
]
R
. If a

[a]
R
then (a

, a) R. But a [a
1
]
R
, so (a, a
1
) R.
By transitivity, (a

, a
1
) R, so a

[a
1
]
R
. This proves [a]
R
[a
1
]
R
. To show the other
containment, let a

[a
1
]
R
so that (a

, a
1
) R. Again, (a, a
1
) R, giving (a
1
, a) R.
Transitivity then implies (a

, a) R, so a

[a]
R
.
5
The picture is that the equivalence classes of R partition A.
Denition 1.2.4. A partition of A is a collection T of subsets of A such that
1. A =

SP
S and
2. S
1
S
2
= whenever S
1
and S
2
in T are not equal.
Using this denition, we can say that if R is an equivalence relation on a set A then the
collection
(
R
= [a]
R
: a A
of equivalence classes form a partition of A.
Just a note to conclude. If we have an equivalence relation R on a set A, it is standard
notation to write
R/A = [a]
R
: a A
for the set of equivalence classes of A under R. This is known as taking the quotient by an
equivalence relation. At times the relation R is written in an implied manner using a symbol
like . For instance, (a, b) R would be written a b. In this case, the quotient is R/ .
We will spend much of the course talking about functions, which are special kinds of
relations.
Denition 1.2.5. Let A and B be sets and f a relation between A and B. We say that f
is a (well-dened) function from A to B, written f : A B if the following hold.
1. For each a A, there is at least one b B such that (a, b) f.
2. For each a A, there is at most one b B such that (a, b) f. That is, if we ever
have (a, b
1
) f and (a, b
2
) f for b
1
, b
2
B, it follows that b
1
= b
2
.
The set A is called the domain of f and the B is called the codomain of f.
Of course we will not continue to use this notation for a function, but the more familiar
notation: if (a, b) f then because of item 2 above, we can unambiguously write f(a) = b.
We will be interested in certain types of functions.
Denition 1.2.6. The function f : A B is called one-to-one (injective) if whenever
a
1
,= a
2
then f(a
1
) ,= f(a
2
). It is called onto (surjective) if for each b B there exists a A
such that f(a) = b.
Another way to dene onto is to rst dene the range of a function f : A B by
f(A) = f(a) : a A
and say that f is onto if f(A) = B.
Many times we want to compose functions to build other ones. Suppose that f : A B
and g : B C are functions. Then
(g f) : A C is dened as (g f)(a) = g((f(a)) .
6
Formally speaking we dene g f A C by
(a, c) g f if (a, b) f and (b, c) g for some b B .
You can check that this denes a function.
Proposition 1.2.7. Let f : A B and g : B C be functions.
1. If f and g are one-to-one then so is g f.
2. If f and g are onto then so is g f.
Proof. We start with the rst statement. Suppose that f and g are one-to-one; we will show
that g f must be one-to-one. Suppose then that a and a

in A are such that (g f)(a) =


(g f)(a

). Then by denition, g(f(a)) = g(f(a

)). But g is one-to-one, so f(a) = f(a

).
Now since f is one-to-one, we nd a = a

. This shows that if (g f)(a) = (g f)(a

) then
a = a

, proving g f is one-to-one.
Suppose then that f and g are onto. To show that g f is onto we must show that
for each c C there exists a A such that (g f)(a) = c.This is the same statement as
g(f(a)) = c. We know that g is onto, so there exists b B such that g(b) = c. Furthermore,
f is onto, so for this specic b, there exists a A such that f(a) = b. Putting these together,
(g f)(a) = g(f(a)) = g(b) = c .
This completes the proof.
If a function is both one-to-one and onto we can dene an inverse function.
Denition 1.2.8. If f : A B is both one-to-one and onto we call f a bijection.
Theorem 1.2.9. Let f : A B. There exists a function f
1
: B A such that
f
1
f = id
A
and f f
1
= id
B
, (1)
where id
A
: A A and id
B
: B B are the identity functions
id
A
(a) = a and id
B
(b) = b
if and only if f is a bijection. The meaning of the above equations is f
1
(f(a)) = a and
f(f
1
(b)) = b for all a A and b B.
Proof. Suppose that f : A B is a bijection. Then dene f
1
B A by
f
1
= (b, a) : (a, b) f .
This is clearly a relation. We claim it is a function. To show this we must prove that
for all b B there exists a A such that (b, a) f
1
and
7
for all b B there exists at most one a A such that (b, a) f
1
.
Restated, these are
for all b B there exists a A such that f(a) = b and
for all b B there exists at most one a A such that f(a) = b.
These are exactly the conditions that f be a bijection, so f
1
is a function.
Now we must show that f
1
f = id
A
and f f
1
= id
B
. We show only the rst; the
second is an exercise. For each a A, there is a b B such that f(a) = b. By denition of
f
1
, we then have (b, a) f
1
; that is, f
1
(b) = a. Therefore (a, b) f and (b, a) f
1
,
giving (a, a) f
1
f, or
(f
1
f)(a) = a = id
A
(a) .
We have now shown that if f is a bijection then there is a function f
1
that satises (1).
For the other direction, suppose that f : A B is a function and g : B A is a function
such that
g f = id
A
and f g = id
B
.
We must show then that f is a bijection. To show one-to-one, suppose that f(a
1
) = f(a
2
).
Then a
1
= id
A
(a
1
) = g(f(a
1
)) = g(f(a
2
)) = id
A
(a
2
) = a
2
., giving that f is one-to-one. To
show onto, let b B; we claim that f maps the element g(b) to b. To see this, compute
b = id
B
(b) = f(g(b)). This shows that f is onto and completes the proof.
Here are some more facts about inverses and injectivity/surjectivity.
If f : A B is a bijection then so is f
1
: B A.
If f : A B and g : B C are bijections then so is g f.
The identity map id
A
: A A is a bijection.
If a function f : A B is not a bijection then there is no inverse function f
1
: B A.
However we can in all cases consider the inverse image.
Denition 1.2.10. Given f : A B and C B we dene the inverse image of C as
f
1
(C) = a A : f(a) C .
Note that if we let C be a singleton set b for some b B then we retrieve all elements
a A mapped to b:
f
1
(b) = a A : f(a) = b .
In the case that f is invertible, this just gives the singleton set consisting of the point f
1
(b).
We note the following properties of inverse images (proved in the homework). For f : A B
and C
1
, C
2
B,
f
1
(C
1
C
2
) = f
1
(C
1
) f
1
(C
2
).
f
1
(C
1
C
2
) = f
1
(C
1
) f
1
(C
2
).
8
1.3 Cardinality
The results of the previous section allow us to dene an equivalence relation on sets:
Denition 1.3.1. If A and B are sets, we say that A and B are equivalent (A B or A
and B have the same cardinality) if there exists a bijection f : A B. The cardinality of a
set A (written (A)) is dened as the equivalence class of A under this relation. That is
(A) = B : A B .
To compare cardinalities, we introduce a new relation on sets.
Denition 1.3.2. If A and B are sets then we write (A) (B) if there exists a one-to-one
function f : A B. Write (A) < (B) if (A) (B) but (A) ,= (B).
The following properties follow. (Exercise: verify the rst two.)
1. (reexivity) For each set A, (A) (A).
2. (transitivity) For all sets A, B, C, if (A) (B) and (B) (C) then (A) (C).
3. (antisymmetry) For all sets A and B, if (A) (B) and (B) (A) then (A) = (B).
Any relation on a set that satises these properties is called a partial order. For cardi-
nality, establishment of antisymmetry is done by the Cantor-Bernstein theorem, which we
will skip.
Theorem 1.3.3 (Cantors Theorem). For any set A let T(A) be the power set of A; that is,
the set whose elements are the subsets of A. Then (A) < (T(A)).
Proof. We rst show that (A) ,= (T(A)). We proceed by contradiction. Suppose that A is
a set but assume that (A) = (T(A)). Then there exists a bijection f : A T(A). Using
this function, dene the set
S = a A : a / f(a) .
Since this is a subset of A, it is an element of T(A). As f is a bijection, it is onto and
therefore there exists s A such that f(s) = S. There are now two possibilities; either
s S or s / S. In either case we will derive a contradiction, proving that the assumption
we made cannot be true: no such f can exist and (A) ,= (T(A)).
In the rst case, s S. Then as S = f(s), we have s f(s). But then by denition of
S, it must actually be that s / S, a contradiction. In the second case, s / S, giving by the
denition of S that s f(s). However f(s) = S so s S, another contradiction.
Second we must show that (A) (T(A)). To do this we dene the function
f : A T(A) by f(a) = a .
To prove injectivity, suppose that f(a
1
) = f(a
2
). Then a
1
= a
2
and therefore a
1
=
a
2
.
9
Let us now give an example of two sets with the same cardinality. If A and B are sets
we write B
A
for the set of functions f : A B. Let F
2
be a set with two elements, which
we call 0 and 1. We claim that
(T(A)) = (F
A
2
) .
To see this we must display a bijection between the two. Dene f : T(A) F
A
2
by the
following. For any subset S A associate the characteristic function
S
: A F
2
by

S
(a) =
_
1 if a S
0 if a / S
.
Exercise: show that the function f : T(A) F
A
2
given by f(S) =
S
is a bijection.
1.4 Natural numbers and induction
To introduce the natural numbers in an axiomatic way we will use the Peano axioms.
Assumption. We assume the existence of a set N, an element 1 N and a function
s : N N with the following properties.
1. For each n N, s(n) (the successor of n) is not equal to 1.
2. s is injective.
3. (Inductive axiom) If any subset S N contains 1 and has the property that whenever
n S then s(n) S, it follows that S = N.
The third property seems a bit weird at rst, but actually there are many sets which satisfy
the rst two properties and are not N. For instance, the set n/2 : n N does. So we need
it to really pin down N.
From these axioms many properties follow. Here is one.
for all n N, s(n) ,= n.
Proof. Let S = n N : s(n) ,= n. Clearly 1 S. Now suppose that n S for some
n. Then we claim that s(n) S. To see this, note that by injectivity of s, s(n) ,= n
implies that s(s(n)) ,= s(n). Thus s(n) S. By the inductive axiom, since 1 S and
whenever n S we have s(n) S, we see that S = N. In other words, s(n) ,= n for
all n.
Addition
It is customary to call s(1) = 2, s(2) = 3, and so on. We dene addition on the natural
numbers in a recursive manner:
for any n N, dene n + 1 to be s(n) and
10
for any n, m N, dene n + s(m) to be s(n + m).
That this indeed denes a function + : NN N requires proof, but we will skip this and
assume that addition is dened normally. Of course, addition satises the commutative and
associative laws.
1. For any m, n, r N, m + (n + r) = (m + n) + r.
Proof. First we show the statement for r = 1 and all m, n. We have
m + (n + 1) = m + s(n) = s(m + n) = (m + n) + 1 ,
where we have used the inductive denition of addition. Now suppose that the formula
holds for some r N; we will show it holds for s(r). Indeed,
m + (n + s(r)) = m + (n + (r + 1)) = m + ((n + r) + 1) = m + s(n + r)
= s(m + (n + r)) = s((m + n) + r) = (m + n) + s(r) .
In other words, the set
S = r N : m + (n + r) = (m + n) + r for all m, n N
has 1 S and whenever r S, also s(r) S. By the inductive axiom, S = N.
2. For any m, n N, m + n = n + m.
Proof. Again we use an inductive argument. Dene
S = n N : n + m = m + n for all m N .
The rst step is to show that 1 S; that is, that 1 + m = m + 1 for all m N. For
this we also do an induction. Set
T = m N : 1 + m = m + 1 .
First, 1 T since 1 +1 = 1 +1. Suppose then that m T. We claim that this implies
m + 1 T. To see this, write
1 + (m + 1) = (1 + m) + 1 = (m + 1) + 1 .
By the induction, T = N.
Now that we have shown 1 S, we assume n S and prove n + 1 S. For m N,
(n + 1) + m = n + (1 + m) = n + (m + 1) = (n + m) + 1
= (m + n) + 1 = m + (n + 1) .
By the inductive axiom, S = N and we are done.
11
3. For all n, m N, n + m ,= n.
Proof. Dene the set
S = n N : n + m ,= nfor all m N .
Then since by the Peano axioms,
1 + m = s(m) ,= 1 for all m N ,
so 1 N. Suppose then that n S; that is, n is such that n + m ,= n for all m N.
Then by injectivity, for m N,
(n + 1) + m = (n + m) + 1 = s(n + m) ,= s(n) = n + 1 ,
giving n + 1 S. By the inductive axiom, S = N and we are done.
Last, for proving facts about ordering we show
s is a bijection from N to N 1.
Proof. We know s does not map any element to 1 so s is in fact a function to N 1.
Also it is injective. To show surjective, consider the set
S = 1 s(n) : n N .
Clearly 1 S. Supposing that n S then n N, so s(n) S. Therefore S = N.
Therefore if k ,= 1 then k = s(n) for some n N.
The above lets us dene n 1 for n ,= 1. It is the element such that (n 1) + 1 = n.
Ordering
We also dene an ordering on the natural numbers. We say that m n for m, n N if
either m = n or m + a = n for some a N. This denes a total ordering of N; that is, it is
a partial ordering that also satises
for all m, n N, m n or n m.
In the case that m n but m ,= n we write m < n. Note that by item 3 above, n < n + m
for all n, m N. In particular, n < s(n).
Proposition 1.4.1. is a total ordering of N.
12
Proof. First each n n so it is reexive. Next if n
1
n
2
and n
2
n
3
then if n
1
= n
2
or
n
2
= n
3
, we clearly have n
1
n
3
. Otherwise there exists m
1
, m
2
N such that n
1
+m
1
= n
2
and n
2
+ m
2
= n
3
. In this case,
n
3
= n
2
+ m
2
= (n
1
+ m
1
) + m
2
= n
1
+ (m
1
+ m
2
) ,
giving n
1
n
3
.
For antisymmetry, suppose that m n and n m. For a contradiction, if m ,= n then
there exists a, b N such that m = n+a and n = m+b. Then m = (m+a)+b = m+(a+b),
a contradiction with item 3 above. Therefore m = n.
So far we have proved that is a partial order. We now prove is a total ordering. To
begin with, we claim that for all n N, 1 n. Clearly this is true for n = 1. If we assume
it holds for some n then
n + 1 = 1 + n 1 ,
verifying the claim by induction.
Now for any m > 1 (that is, m N with m ,= 1), dene the set
S = n N : n m n N : m n .
By the above remarks, 1 S. Supposing now that n S for some n N, we claim that
n + 1 S. To show this, we have three cases.
1. Case 1: n = m. In this case, n + 1 = m + 1 m, giving n + 1 S.
2. Case 2: n > m, so there exists a N such that n = m+a. Then n+1 = m+a+1 m,
giving n + 1 S.
3. Case 3: n < m, so there exists a N such that m = n+a. If a = 1 then n+1 = m S.
Otherwise a > 1, implying that a 1 N (that is, a 1 is dened), so
m = n + a = n + a 1 + 1 = (n + 1) + a 1 > n + 1 ,
so that n + 1 S. By the inductive axiom, S = N and therefore for all n, we have
n m or m n.
A consequence of these properties is trichotomy of the natural numbers. For any m, n
N, exactly one of the following holds: m < n, m = n or n < m.
A property that relates addition and ordering is
if m, n, r N such that m < n then m + r < n + r.
Proof. There must be a N such that n = m+a. Then n+r = m+a+r = m+r +a,
giving m + r < n + r.
Clearly then if m n and r N we have m + r n + r.
13
If n < k then n + 1 k.
Proof. If n < k then there exists j N such that n + j = k. Because 1 j we nd
n + 1 n + j = k.
Multiplication.
We dene multiplication inductively by
n 1 = n for all n N
n s(m) = n + (n m) .
One can prove the following properties; (try it!) let m, n, r, s N:
1. for all n, m, r N,
n (m + r) = (n m) + (n r) .
2. n m = m n.
3. (n m) r = n (m r).
4. if n < m and r s then rn < sm.
1.5 Cardinality and the natural numbers
For each n N we write the set
J
n
= m N : m n .
Note that J
1
= 1 and for n 1, we have
J
n+1
= J
n
n + 1 .
To show this let k be in the right side. If k = n + 1 then k J
n+1
. Otherwise k n, giving
by n n + 1 the inequality k n + 1, or k J
n+1
. To prove the inclusion , suppose that
k J
n+1
. If k J
n
we are done, so suppose that k / J
n
. Therefore k > n, so k n +1. On
the other hand, k n + 1, so k = n + 1.
Denition 1.5.1. For an arbitrary set A we say that A has cardinality n if A J
n
. In this
case we say A is nite and we write (A) = n. If A is not equivalent to any J
n
we say A is
innite.
In this denition, (A) is an equivalence class of sets and n is a number, so what we have
written here is purely symbolic: it means A J
n
.
Lemma 1.5.2. If A and B are sets such that A B then (A) (B).
Proof. Dene f : A B by f(a) = a. Then f is an injection.
14
Theorem 1.5.3. For all n N, (J
n
) < (J
n+1
) < N.
Proof. Each set above is a subset of the next, so the proposition holds using instead of
<. We must then prove ,= in each spot above. Assume rst that we have proved that
(J
n
) ,= (J
n+1
) for all n N; we will show that (J
n
) ,= N for all n N. If we had equality,
then we would nd (J
n+1
) N = (J
n
). This contradicts the rst inequality.
To prove the inequality (J
n
) ,= (J
n+1
), we use induction. Clearly it holds for n = 1
since J
1
= 1 and J
2
= 1, 2 and any function from J
1
to J
2
can only have one element
in its range (cannot be onto). Suppose then that (J
n
) ,= (J
n+1
); we will prove that
(J
n+1
) ,= (J
n+2
) by contradiction. Assume that there is a bijection f : J
n+1
J
n+2
. Then
some element must be mapped to n + 2; call this k J
n+1
. Dene h : J
n+1
J
n+1
by
h(m) =
_

_
m m ,= k, n + 1
n + 1 m = k
k m = n + 1
.
This function just swaps k and n + 1. It follows then that

f = f h : J
n+1
J
n+2
is a
bijection that maps n + 1 to n + 2.
Now J
n
is just J
n+1
n + 1 and J
n+1
is just J
n+2
n + 2, so dene g : J
n
J
n+1
to do exactly what

f does: g(m) =

f(m). It follows that g is a bijection from J
n
to J
n+1
,
giving J
n
J
n+1
, a contradiction.
Because of the proposition, if a set A has A N it must be innite. In this case we say
that A is countable. Otherwise, if A is innite and (A) ,= N, we say it is uncountable. From
this point on, we will be more loose about working with the natural numbers. For example,
we will use the terms nite and innite in the same way that we normally do a set is nite
if it has nitely many elements and innite otherwise. Of course every proof we write from
now on could be done using the Peano axioms, but we will be spared that.
Theorem 1.5.4. Let S be an innite subset of N. Then S is countably innite.
Proof. We must construct a bijection from N to S. We can actually do this using the well-
ordering property: that each non-empty subset of N has a least element. Dene f : N S
recursively: f(1) is the least element of S and, assuming we have dened f(1), . . . , f(n),
dene f(n + 1) to be the least element of S f(1), . . . , f(n).
This is a bijection.
Denition 1.5.5. We say a set A is countable if it is either nite or countably innite.
Note that A is countable if and only if there is an injection f : A N; that is, that
(A) N.
Theorem 1.5.6. Let ( be a countable collection of countable sets. Then
AC
A is countable.
15
Proof. To prove this we need to construct a bijection from N. We will do this somewhat
non-rigorously, thinking of a bijection from N as a listing of elements of
AC
A in sequence.
For example, given a countably innite set S we may take a bijection f : N S and list all
of the elements of S as
f(1), f(2), f(3), . . .
If S is nite then this corresponds to a nite list.
Since each A ( is countable, we may list its elements. The collection ( itself is countable
so we can list the elements of
AC
A in an array:
a
1
a
2
b
1
b
2
b
3

c
1
d
1
d
2
d
3
d
4


Note that some rows are nite. We now list the elements according to diagonals. That is,
we write the list as
a
1
, b
1
, a
2
, c
1
, b
2
, d
1
, b
3
, d
2
, . . .
Because we want the list to correspond to a bijection, we need to make sure that no element
is repeated. So, for instance, if b
1
and a
2
are equal we would only include the rst.
1.6 Exercises
1. Let f : A B and g : B C be functions. Show that the relation g f A C,
dened by
(a, c) g f if (a, b) f and (b, c) g for some b B
is a function.
2. Show that the function f : T(A) F
A
2
mentioned at the end of Section 1.3 and given
by f(S) =
S
is a bijection.
3. Prove the properties of multiplication listed at the end of Section 1.4.
4. Prove the following statements by induction.
(a) For all n N,
1 + 2 + + n =
n(n + 1)
2
.
(b) For all n N,
1
2
+ 2
2
+ + n
2
=
n(n + 1)(2n + 1)
6
.
16
5. Strong Induction. In this exercise we introduce strong mathematical induction,
which, although being referred to as strong, is actually equivalent to mathematical
induction. Suppose we are given a collection P(n) : n N of mathematical state-
ments. To show P(n) is true for all n, mathematical induction dictates that we show
two things hold: P(1) is true and if P(n) is true for some n N then P(n+1) is true.
To argue instead using strong induction we prove that
P(1) is true and
if n N is such that P(k) is true for all k n then P(n + 1) is true.
(a) Dene a sequence (a
n
) of real numbers recursively by
a
1
= 1 and a
n
= a
1
+ + a
[n/2]
for n 2 .
(Here [n/2] is the largest integer no bigger than n/2.) Prove by strong induction
that a
n
2
n1
for n 2. Is it possible to nd b < 2 such that a
n
b
n1
for all
n 2?
(b) Why does strong induction follow from mathematical induction? In other words
in the second step of strong induction, why are we allowed to assume that P(k)
is true for all k n to prove that P(n + 1) is true?
6. Prove that any non-empty subset S N has a least element. That is, there is an s S
such that for all t S we have s t. This is a major result about N, expressed by
saying that N is well-ordered.
Hint. Assume there is no least element. Let
M = m N : t S, m t .
Use Peanos induction axiom to prove that M = N. Does this lead to a contradiction?
17
2 The real numbers
2.1 Rationals and suprema
From now on we will proceed through Rudin, using the standard notations
Z = . . . , 1, 0, 1, . . .
Q = m/n : m, n Z and n ,= 0 .
When thinking about the rational numbers, we quickly come to realize that they do not
capture all that we wish to express using numbers. For instance,
Theorem 2.1.1. There is no rational number whose square is 2.
Proof. We argue by contradiction, so assume that 2 = (m/n)
2
for some m, n Z with n ,= 0.
We may assume that m and n are not both even; otherwise, we can reduce the fraction,
removing enough factors of 2 from the numerator and denominator. Then
2n
2
= m
2
,
so m
2
is even. This actually implies that m must be even, for otherwise m
2
would be odd
(since the square of an odd number is odd). Therefore we can write m = 2s for some s Z.
Plugging back in, we nd
2n
2
= 4s
2
or n
2
= 2s
2
,
so n
2
is also even, giving that n is even. This is a contradiction.
From the previous theorem, what we know as

2 is not a rational number. Therefore if


we were to construct a theory from only rationals, we would have a hole where we think

2 should be. What is even stranger is that there are rational numbers arbitrarily close to
this hole.
Theorem 2.1.2. If q Q satises 0 < q
2
< 2 then we can nd another rational q Q such
that
q
2
< q
2
< 2 .
Similarly, for each r Q such that r
2
> 2, there is another rational r such that 2 < r
2
< r
2
.
Proof. Suppose that q > 0 satises q
2
< 2 and dene
q = q +
2 q
2
q + 2
.
Then q > q and
q
2
2 =
2(q
2
2)
(q + 2)
2
,
giving q
2
< 2.
18
We see from above that the set q Q : q
2
< 2 does not have a largest element. This
leads us to study largest elements of sets more carefully.
Denition 2.1.3. If A is a set with a partial ordering we say that a A is an upper
bound for a subset B A if b a for all b B. We say that a is a least upper bound for B
if whenever a

is an upper bound for B, we have a a

. We dene lower bound and greatest


lower bound similarly.
Note that if a is a least upper bound for B then a is unique. Indeed, assume that a and
a

are least upper bounds. Since they are both upper bounds, we have a a

and a

a, so
by antisymmetry of partial orderings, a = a

. Because of this uniqueness, there is no harm


in writing
a = sup B when a is the least upper bound of B
and
a = inf B when a is the greatest lower bound of B .
Proposition 2.1.4. Let A be a totally ordered set and B a subset. Dene C to be the set
of all upper bounds for B. Then sup B = inf C.
Proof. We are trying to show that some element (inf C) is the supremum of B, so we must
show two things: inf C is an upper bound for B and any other upper bound a for B satises
inf C a. The second statement is easy because if a is an upper bound for B then a C.
As inf C is a lower bound for C we then have inf C a.
For the rst, assume that inf C is not an upper bound for B, so there exists b B such
that inf C is not b. By trichotomy, inf C < b. We claim then that b is a lower bound for C
which is larger than the greatest lower bound, a contradiction. Why is this? If c C then
c is an upper bound for B, giving c b, or b c.
Note that the second statement of Theorem 2.1.2 states that the set q Q : q
2
> 2
does not have a supremum in Q. Indeed, if it did have a supremum r, then r would be a
rational upper bound for this set and then we could nd a smaller r that is still an upper
bound, a contradiction. So one way of formulating the fact that there are holes in Q is to
say that it does not have the least upper bound property.
Denition 2.1.5. Let A be a totally ordered set with order . We say that A has the least
upper bound property if each nonempty subset B A with an upper bound in A has a least
upper bound in A.
2.2 Existence and properties of real numbers
Therefore we are led to extend the rational numbers to ll in the holes. This is actually
quite a dicult procedure and there are many routes to its end. We will not discuss these,
however, and will instead state the main theorem about the existence of the real numbers
without proof. The main point of this course will be to understand properties of the real
numbers, and not its existence and uniqueness.
19
For the statement, one needs the denition of an ordered eld, which is a certain type of
totally ordered set with multiplication and addition (like the rationals).
Theorem 2.2.1 (Existence and uniqueness of R). There exists a unique ordered eld with
the least upper bound property.
The sense in which uniqueness holds is somewhat technical; it is not that any two ordered
elds as above must be equal, but they must be isomorphic. Again we defer to Rudin for
these denitions. We will now assume the existence of R, that it contains Q and Z, and its
usual properties.
One extremely useful property of R that follows from the least upper bound property is
Theorem 2.2.2 (Archimedean property of R). Given x, y R with x ,= 0, there exists
n Z such that
nx > y .
Proof. First let x, y R such that x, y > 0 and assume that there is no such n. Then the
set
nx : n N
is bounded above by y. As it is clearly nonempty, it has a supremum s. Then s x < s, so
s x cannot be an upper bound, giving the existence of some m N such that
s x < mx .
However this implies that s < (m+1)x, so s was actually not an upper bound, contradiction.
This proves the statement for the case x < y. The other cases can be obtained from this one
by instead considering x and/or y.
The Archimedean property implies
Corollary 2.2.3 (Density of Q in R). Let x, y R with x < y. There exists q Q such
that x < q < y.
Proof. Apply the Archimedean property to y x and 1 to nd n Z such that n(y x) > 1.
We can also nd m
1
> nx and m
2
> nx, so
m
2
< nx < m
1
.
It follows then that there is an m Z such that m1 nx < m. Finally,
nx < m 1 + nx < ny .
Dividing by n we get x < m/n < y.
Now we return to countability.
Theorem 2.2.4. The set Q is countable, whereas R is uncountable.
20
Proof. We already know that N N is countable: this is from setting up the array
(1, 1) (2, 1) (3, 1)
(1, 2) (2, 2) (3, 2)
(1, 3) (2, 3) (3, 3)

and listing the elements along diagonals. On the other hand, there is an injection
f : Q
+
N N ,
where Q
+
is the set of positive rationals. One such f is given by f(m/n) = (m, n), where
m/n is the reduced fraction for the rational, expressed with m, n N. Therefore Q
+
is
countable. Similarly, Q

, the set of negative rationals, is countable. Last, Q = Q


+
Q

0
is a union of 3 countable sets and is thus countable.
To prove R is uncountable, we will use decimal expansions for real numbers. In other
words, we write
x = .a
1
a
2
a
3
. . .
where a
i
0, . . . , 9 for all i. Since we have not proved anything about decimal expansions,
we are certainly assuming a lot here, but this is how things go. Note that each real number
has at most 2 decimal expansions (for instance, 1/4 = .2500 . . . = .2499 . . .).
Assume that R is countable. Then as there are at most two decimal expansions for each
real number, the set of decimal expansions is countable (check this!) Now write the set of
all expansions in a list:
1 .a
0
a
1
a
2
. . .
2 .b
0
b
1
b
2
. . .
3 .c
0
c
1
c
2
. . .

We will show that no matter what list we are given (as above), there must be a sequence
that is not in the list. This implies that there can be no such list, and thus R is uncountable.
Consider the diagonal element of the list. That is, we take a
0
for the rst digit, b
1
for
the second, c
2
for the third and so on:
.a
0
b
1
c
2
d
3
. . .
We now have a rule to transform this diagonal element into a new one. We can use many,
but here is one: change each digit to a 0 if it is not 0, and replace it with 9 if it is 0. For
example,
.0119020 . . . .9000909 . . .
Note that this procedure changes the diagonal number into a new one that diers from the
diagonal element in every decimal place. Call this new expansion A = . a
0
a
1
. . .
Now our original list contains all expansions, so it must contain A at some point; let us
say that the n-th element of the list is A. Then consider the n-th digit a
n
of A. On the
one hand, by construction, a
n
is not equal to the n-th digit of the diagonal element. On the
other hand, by the position in the list, a
n
equals the n-th digit of the diagonal element. This
is a contradiction.
21
2.3 R
n
for n 2
A very important extension of R is given by n-dimensional Euclidean space.
Denition 2.3.1. For n 2, the set R
n
is dened as
R
n
= a = (a
1
, . . . , a
n
) : a
i
R for all i .
Addition of elements is dened as
a +

b = (a
1
, . . . , a
n
) + (b
1
, . . . , b
n
) = (a
1
+ b
1
, . . . , a
n
+ b
n
)
and multiplication of elements by numbers is
ca = c(a
1
, . . . , a
n
) = (ca
1
, . . . , ca
n
), c R .
Note that this denition gives us R for n = 1.
On R
n
we place a distance, but to do that, we need the existence of square roots. We
will take this for granted now, since we will prove it later using continuity.
Lemma 2.3.2. For each x R with x 0 there exists a unique y R such that y
2
= x.
This element is written y =

x.
Denition 2.3.3. On the set R
n
we dene the norm
[a[ = [(a
1
, . . . , a
n
)[ =
_
a
2
1
+ + a
2
n
and inner product
a

b = (a
1
, . . . , a
n
) (b
1
, . . . , b
n
) = a
1
b
1
+ + a
n
b
n
.
Theorem 2.3.4. Suppose a,

b, c R
n
and c R. Then
1. [a[ 0 with [a[ = 0 if and only if a = 0.
2. [ca[ = [c[[a[.
3. (Cauchy-Schwarz inequality) [a

b[ [a[[

b[.
4. (Triangle inequality) [a +

b[ [a[ +[

b[.
5. [a

b[ [a c[ +[c

b[.
Proof. The rst two follow easily; for instance since a
2
0 for all a R (this is actually
part of the denition of ordered eld), we get a
2
1
+ + a
2
n
0 and therefore [a[ 0. If
[a[ = 0 then by uniqueness of square roots, a
2
1
+ + a
2
n
= 0 and so 0 a
2
i
for all i, giving
a
i
= 0 for all i.
For the third item, we rst give a lemma.
22
Lemma 2.3.5. If ax
2
+ bx + c 0 for all x R then b
2
4ac.
Proof. If a = 0 then bx c for all x. Then we claim b must be zero. If not, then plugging
in either 2c/b or 2c/b will give bx < c, a contradiction. Therefore is a = 0 we must have
b = 0 and therefore b
2
4ac as claimed.
Otherwise a ,= 0. First assume that a > 0. Plug in x = b/(2a) to get
b
2
/(4a) + c 0
giving b
2
4ac. Last, if a < 0 then we have (a)x
2
+ (b)x + (c) 0 and applying what
we have proved already to this polynomial, we nd (b)
2
4(a)(c), or b
2
4ac.
To prove Cauchy-Schwarz, note that for all x R,
0 (a
1
x b
1
)
2
+ + (a
n
x b
n
)
2
= (a
2
1
+ + a
2
n
)x
2
2(a
1
b
1
+ + a
n
b
n
)x + (b
2
1
+ + b
2
n
)
= [a[
2
x
2
2(a

b)x +[

b[
2
.
So using the lemma, (a

b)
2
[a[
2
[

b[
2
.
The last two items follow directly from the Cauchy-Schwarz inequality. Indeed,
[a +

b[
2
= (a

b) (a

b)
=a a + 2a

b +

b
[a[
2
+ 2[a[[

b[ +[

b[
2
= ([a[ +[

b[)
2
.
The last inequality follows by taking a c and c

b in the previous.
2.4 Exercises
1. For each of the following examples, nd the supremum and the inmum of the set S.
Also state whether or not they are elements of S.
(a) S = x [0, 5] : cos x = 0.
(b) S = x : x
2
2x 3 < 0.
(c) S = s
n
: s
n
=

n
i=1
2
i
.
2. Prove by induction that for all n N and real numbers x
1
, . . . , x
n
,
[x
1
+ + x
n
[ [x
1
[ + +[x
n
[ .
3. Let A, B R be nonempty and bounded above.
23
(a) Dene the sum set
A + B = a + b : a A, b B .
Prove that sup(A + B) = sup A + sup B.
(b) Dene the product set
A B = a b : a A, b B .
Is it true that sup(AB) = sup Asup B? If so, provide a proof; otherwise, provide
a counterexample.
4. Let ( be a collection of open intervals (sets I = (a, b) for a < b) such that
for all I (, I ,= and
if I, J ( satisfy I ,= J then I J = .
Prove that ( is countable.
Hint. Dene a function f : ( S for some countable set S R by setting f(I) equal
to some carefully chosen number.
24
3 Metric spaces
3.1 Denitions
Denition 3.1.1. A set X with a function d : X X R is a metric space if for all
x, y, z X,
1. d(x, y) 0 and equals 0 if and only if x = y and
2. d(x, y) d(x, z) + d(z, y).
Then we call d a metric.
Examples.
1. A useful example of a metric space is R
n
with metric d(a,

b) = [a

b[.
2. If X is any nonempty set we can dene the discrete metric by
d(x, y) =
_
1 if x ,= y
0 if x = y
.
3. The set F[0, 1] of bounded functions f : [0, 1] R is a metric space with metric
d(f, g) = sup[f(x) g(x)[ : x [0, 1] .
3.2 Open and closed sets
Let (X, d) be a metric space. We are interested in the possible subsets of X and in what
ways we can describe these using the metric d. Lets start with the simplest.
Denition 3.2.1. Let r > 0. The neighborhood of radius r centered at x X is the set
B
r
(x) = y X : d(x, y) < r
For example,
1. in R using the metric d(x, y) = [x y[ we have the open interval
B
r
(x) = (x r, x + r) = y R : x r < y < x + r .
2. In R
n
using the metric d(x, y) = [x y[ we have the open ball
B
r
(x) = (y
1
, . . . , y
n
) : (x
1
y
1
)
2
+ + (x
n
y
n
)
2
< r
2
.
To describe that these sets appear to be open (that is, no point is on the boundary), we
introduce a formal denition of open.
Denition 3.2.2. Let (X, d) be a metric space. A set Y X is open if for each y Y
there exists r > 0 such that B
r
(y) Y .
For each point y we must be able to t a (possibly tiny) neighborhood around y so that
it still stays in the set Y . Thinking of Y as, for example, an open ball in R
n
, as our point
y approaches the boundary of this set, the radius we take for the neighborhood around this
point will have to decrease.
25
Proposition 3.2.3. Any neighborhood is open.
Proof. Let x X and r > 0. To show that B
r
(x) is open we must choose y B
r
(x) and
show that there exists some s > 0 such that B
s
(y) B
r
(x). The radius s will depend on
how close y is to the boundary. Therefore, choose
s = r d(x, y) .
To show that for this s, we have B
s
(y) B
r
(x) we take z B
s
(y). Then
d(x, z) d(x, y) + d(y, z) < d(x, y) + s = r .
Some more examples:
1. In R, the only intervals that are open are the (surprise!) open intervals. For instance,
lets consider the half-open interval (0, 1] = x R : 0 < x 1. If it were open, we
would be able to, given any x (0, 1], nd r > 0 such that B
r
(x) (0, 1]. But clearly
this is false because B
r
(1) contains 1 + r/2.
2. In R
2
, the set (x, y) : y > 0 (x, y) : y < 1 is open.
3. In R
3
, the set (x, y, z) : y > 0 (0, 0, 0) is not open.
Proposition 3.2.4. Let ( be a collection of open sets.
1. The union

OC
O is open.
2. If ( is nite then

OC
O is open. This need not be true if the collection is innite.
Proof. Let x
OC
O. Then there exists O ( such that x O. Since O is open, there
exists r > 0 such that B
r
(x) O. This is also a subset of
OC
O so this set is open.
To show that we cannot allow innite intersections, consider the sets (1/n, 1 + 1/n) in
R. We have

n=1
(1/n, 1 + 1/n) = [0, 1] ,
which is not open (under the usual metric of R).
For nite intersections, let O
1
, . . . , O
n
be the open sets from ( and x
n
i=1
O
i
. Then
for each i, we have x O
i
and therefore there exists r
i
> 0 such that B
r
i
(x) O
i
. Letting
r = minr
1
, . . . , r
n
, we have B
r
(x) B
r
i
(x) for all i and therefore B
r
(x) O
i
for all i.
This implies B
r
(x) is a subset of the intersection and we are done.
Denition 3.2.5. An interior point of Y X is a point y Y such that there exists r > 0
with B
r
(y) Y . Write Y

for the set of interior points of Y .


Directly by denition, Y is open if and only if Y = Y

.
26
Examples:
1. The set of interior points of [0, 1] (under the usual metric) is (0, 1).
2. The set of interior points of
(x, y) : y > 0 (x, y) : x = 1, y 0
is just (x, y) : y > 0.
3. What is the set of interior points of Q?
4. Dene a metric on R
2
by d(x, y) = 1 if x ,= y and 0 otherwise. This can be shown to
be a metric. Given a set Y R
2
, what is Y

?
Denition 3.2.6. A set Y X is closed if its complement X Y is open.
Sets can be both open and closed. Consider , whose complement is clearly open, making
closed. It is also open.
Proposition 3.2.7. Let ( be a collection of closed sets.
1.

CC
C is closed.
2. If ( is nite then

CC
C is closed.
Proof. Just use X [
CC
C] =
CC
(X C).
3.3 Limit points
There is an alternative characterization of closed sets in terms of limit points
Denition 3.3.1. Let Y X. A point x X is a limit point of Y if for each r > 0 there
exists y Y such that y ,= x and y B
r
(x). Write Y

for the set of limit points of Y .


Examples:
1. 0 is a limit point of 1, 1/2, 1/3, . . ..
2. 1, 2, 3 has no limit points.
3. In R
2
, B
1
(0) (0, y) : y R (10, 10) has limit points
(x, y) : x
2
+ y
2
1 (0, y) : y R .
Actually we could have given a dierent denition of limit point.
Proposition 3.3.2. x X is a limit point of Y if and only if for each r > 0 there are
innitely many points of Y in B
r
(x)
27
Proof. We need only show that if x is a limit point of Y and r > 0 then there are in-
nitely many points of Y in B
r
(x). We argue by contradiction; assume there are only
nitely many and label the ones that are not equal to x as y
1
, . . . , y
n
. Choosing r =
mind(x, y
1
), . . . , d(x, y
n
), we then have that B
r
(x) contains no points of Y except pos-
sibly x. This contradicts the fact that x is a limit point of Y .
Here is yet another denition of closed.
Theorem 3.3.3. Y is closed if and only if Y

Y .
Proof. Suppose Y is closed and let y be limit point of Y . If y / Y then because X Y is
open, we can nd r > 0 such that B
r
(y) (X Y ). But for this r, there is no x B
r
(y)
that is also in Y , so that y is not a limit point of Y , a contradiction.
Suppose conversely that Y

Y ; we will show that Y is closed by showing that X Y is


open. To do this, let z X Y . Since z / Y and Y

Y , z cannot be a limit point of Y .


Therefore there is an r > 0 such that B
r
(z) contains no points p ,= z such that p Y . Since
z is also not in Y , we must have B
r
(z) (X Y ), implying that X Y is open.
Examples:
1. Again the set 1, 2, 3 has no limit points (because from the above proposition, a nite
set cannot have limit points). However it is closed by the above theorem.
2. Is Q closed in R? How about Q
2
in R
2
?
3. The set Z has no limit points in R, so it is closed.
Denition 3.3.4. The closure of Y in X is the set Y = Y Y

.
Theorem 3.3.5. Let ( be the collection of all sets C X such that C is closed and Y C.
Then
Y =

CC
C .
Proof. We rst show the inclusion . To do this we need to show that each y Y and
each y Y

must be in the intersection on the right (call it J). First if y Y then because
each C ( contains Y , we have y J. Second, if y Y

and C ( we also claim that


y C. This is because y, being a limit point of Y , is also a limit point of C (directly from
the denition). However C is closed, so it contains its limit points, and y C.
For the inclusion , we will show that Y (. This implies that Y is one of the sets we
are intersecting to form J, and so J Y . Clearly Y Y , so we need to show that Y is
closed. If x / Y then x is not in Y and x is not a limit point of Y , so there exists r > 0 such
that B
r
(x) does not intersect Y . Since B
r
(x) is open, each point in it has a neighborhood is
contained in B
r
(x) and therefore does not intersect Y . This means that each point in B
r
(x)
is not in Y and is not a limit point of Y , giving B
r
(x)
_
Y
_
c
, so the complement of Y is
open. Thus Y is closed.
From the theorem above, we have a couple of consequences:
28
1. For all Y X, Y is closed. This is because the intersection of closed sets is closed.
2. Y = Y if and only if Y is closed. One direction is clear: that if Y = Y then Y is closed.
For the other direction, if Y is closed then Y

Y and therefore Y = Y Y

Y .
Examples:
1. Q = R.
2. R Q = R.
3. 1, 1/2, 1/3, . . . = 1, 1/2, 1/3, . . . 0.
For some practice, we give Theorem 2.28 from Rudin:
Theorem 3.3.6. Let Y R be nonempty and bounded above. Then sup Y Y and therefore
sup Y Y if Y is closed.
Proof. By the least upper bound property, s = sup Y exists. To show s Y we need to
show that s Y or s Y

. If s Y we are done, so we assume s / Y and prove that s Y

.
Since s is the least upper bound, given r > 0 there must exist y Y such that
s r < y s .
If this were not true, then s r would be an upper bound for Y . But now we have found
y Y such that y ,= s and y B
r
(s), proving that s is a limit point for Y .
Note that sup Y is not always a limit point of Y . Indeed, consider the set
Y = 0 .
This set has sup Y = 0 but has no limit points. The set Y can even have limit points but
just not with sup Y a limit point. Consider Y = 0 [2, 1].
3.4 Compactness
It will be very important for us, during the study of continuity for instance, to understand
exactly which sets Y R have the following property: for each innite subset E Y , E
has a limit point in Y . We will soon see that the interval [0, 1] has this property, whereas
(0, 1) does not (take for example the subset 1, 1/2, 1/3, . . .). The reason is that we will
many times nd ourselves exactly in this situation: with an innite subset E of some set Y
and we will want to nd a limit point for E (and hope that it is also in E). This property
is what we will call on the problem set limit point compactness.
Limit point compactness was apparently one of the original notions of compactness (see
the discussion in Munkres topology book at the beginning of the compactness section
thanks Prof. McConnell). However over time it became apparent that there was a stronger
and more general version of compactness (equivalent in metric spaces, but not in all topo-
logical spaces) which could be formulated only in terms of open sets. We give this denition,
now taken to be the standard one, below.
29
Denition 3.4.1. A subset K of a metric space X is compact if for every collection (
of open sets such that K
CC
C, there are nitely many sets C
1
, . . . , C
n
( such that
K
n
i=1
C
i
.
The collection ( is called an open cover for K and C
1
, . . . , C
n
is a nite subcover.
The process of choosing this nite number of sets from ( is referred to as extracting a nite
subcover. The denition, in this language, states that K is compact if from every open cover
of K we can extract a nite subcover of K.
It is quite dicult to gain intuition about the above denition, but it will develop as we
go on and use compactness in various circumstances. The main point is that nite collections
are much more useful than innite collections. This is true for example with numbers: we
already know that a set of nitely many numbers has a min and a max, whereas an innite set
does not necessarily. As we go through the course, to develop a clearer view of compactness,
you should revisit the following phrase: often times, compactness allows us to pass from
local information (valid in each open set from the cover) to global information (valid on
the whole space), by patching together the sets in the nite subcover.
Let us now give some properties of compact sets and try to emphasize where the ability
to extract nite subcovers comes into the proofs.
Theorem 3.4.2. Any compact set is limit point compact.
Proof. Let K X be compact and let E K be an innite set. Assume for a contradiction
that E has no limit point in K, so for each x K we can nd r
x
> 0 such that B
rx
(x)
intersects E only possibly at x. The collection ( = B
rx
(x) : x K is an open cover of
K, so by compactness it can be reduced to a nite subcover of K (and thus of E). But this
means that E must have been nite, a contradiction.
Denition 3.4.3. A set E X is bounded if there exists x X and R > 0 such that
E B
R
(x).
Theorem 3.4.4. Any compact K X is bounded.
Proof. Pick x X and dene a collection ( of open sets by
( = B
R
(x) : R N .
We claim that ( is an open cover of K. We need just to show that each point of X is in at
least one of the sets of (. So let y X and choose R > d(y, x). Then y B
R
(x).
Since K is compact, there exist C
1
, . . . , C
n
( such that K
n
i=1
C
i
. By denition
of the sets in ( we can then nd R
1
, . . . , R
n
such that K
n
i=1
B
R
i
(x). Taking R =
maxR
1
, . . . , R
n
, we then have K B
R
(x), completing the proof.
In the proof it was essential to extract a nite subcover because we wanted to take R to
be the maximum of radii of sets in (. This is clearly innity if we have an innite subcover,
and so in this case the proof would break down (that is, if K we were not able to extract a
nite subcover).
Examples.
30
1. The set 1/2, 1/3, . . . is not compact. This is because we can nd an open cover that
admits no nite subcover. Indeed, consider
( =
__
1
n

1
2n
,
1
n
+
1
2n
_
: n 2
_
.
Each one of the sets in the above collection covers only nitely many elements from
1/2, 1/3, . . ., and so any nite sub collection cannot cover the whole set.
2. However if we add 0, by considering the set 1/2, 1/3, . . . 0, it becomes compact.
To prove this, let ( be any open cover; we will show that there are nitely many sets
from ( that still cover our set.
To do this, note rst that there must be some C ( such that 0 C. Since C is
open, it contains some interval (r, r) for r > 0. Then for n > 1/r, all points 1/n are
in this interval, and thus C contains all but nitely many of the points from our set.
Now we just need to cover the other points, of which there are nitely many. Writing
1/2, . . . , 1/N for these points, choose for each i a set C
i
from ( such that 1/i C
i
.
Then
C, C
2
, . . . , C
N

is a nite subcover.
The main problem in example 1 was actually that the set was not closed. It is not
immediately apparent how that was manifested in our inability to produce a nite subcover,
but it is a general fact:
Theorem 3.4.5. Any compact K X is closed.
Proof. We will show that K
c
= X K is open. Therefore pick x K
c
; we will produce an
r > 0 such that B
r
(x) K
c
.
We rst produce an open cover of K. For each y K, the distance d(x, y) must be
positive, since x ,= y (as x / K). Therefore dene the ball
B
y
= B(y, d(x, y)/2) .
We now dene the collection
( = B
y
: y K .
Since each y B
y
this is an open cover of K.
By compactness, we can extract a nite subcover B
y
1
, . . . , B
yn
. Choosing
r = mind(x, y
i
)/2 : i = 1, . . . , n ,
we claim then that B
r
(x) K
c
. To show this, let z B
r
(x). Then d(z, x) < r and by the
triangle inequality,
d(y
i
, x) < d(z, y
i
) + d(z, x) < d(z, y
i
) + r ,
31
giving
d(x, y
i
)/2 d(y
i
, x) r d(z, y
i
) for all i = 1, . . . , n .
In other words, z / B
y
i
for all i. But the B
y
i
s cover K and therefore z / K. This means
K
c
is open, or K is closed.
We now mention a useful way to produce new compact sets from old ones.
Theorem 3.4.6. If K X is compact and L K is closed, then L is compact.
Proof. Let ( be an open cover of L. Dene
T = ( L
c

and note that T is actually an open cover of K. Therefore, as K is compact, we can extract
from T a nite subcover D
1
, . . . , D
n
. If D
i
( for all i, then we are done; otherwise L
c
it in this set (say it is D
n
) and we consider the collection D
1
, . . . , D
n1
. This is a nite
subcollection of (. We claim that it is an open cover of L as well. Indeed, if x L then
there exists i = 1, . . . , n such that x D
i
. Since x / L
c
, D
i
cannot equal :
c
, meaning that
i ,= n. This completes the proof.
3.5 Heine-Borel Theorem: compactness in R
n
In the above theorems, we see that a compact set is always closed and bounded. The converse
is true, but not in all metric spaces. The fact that it is true in R
n
is called the Heine-Borel
theorem.
Theorem 3.5.1 (Heine-Borel). A set K R
n
is compact if and only if it is closed and
bounded.
To prove this theorem, we will need some preliminary results. Recall that Rudin denes
an n-cell to be a subset of R
n
of the form
[a
1
, b
1
] [a
n
, b
n
] for a
i
b
i
, i = 1, . . . , n .
Lemma 3.5.2. Suppose that C
1
, C
2
, . . . are n-cells that are nested; that is, if
C
i
=
n

k=1
[a
(k)
i
, b
(k)
i
] ,
then
[a
(k)
i
, b
(k)
i
] [a
(k)
i+1
, b
(k)
i+1
] for all i and k .
Then
i
C
i
is nonempty.
32
Proof. We rst consider the case n = 1. That is, take C
i
= [a
i
, b
i
] for i 1 and a
i
b
i
.
Dene A = a
1
, a
2
, . . . and a = sup A. We claim that
a
i
C
i
.
To see this, note that a
i
b
j
for all i, j. Indeed,
a
i
a
j
b
j
if i j
and
a
i
b
i
b
j
if i j .
Therefore b
j
is an upper bound for A. But a is the least upper bound of A so a b
j
for all
m. This gives
a
i
a b
i
for all m ,
or a
i
C
i
.
For the case n 2 we just do the same argument on each of the coordinates to nd
(a(1), . . . , a(n)) such that
a
(k)
i
a(k) b
(k)
i
for all i, k ,
or (a(1), . . . , a(n))
i
C
i
.
Lemma 3.5.3. Any n-cell is compact in R
n
.
Proof. For simplicity, take K = [0, 1] [0, 1] = [0, 1]
n
. Since R
n
is a metric space (with
the usual metric), it suces to prove that K is limit point compact; that is, that each innite
subset of K has a limit point in K. This is from exercise 11 at the end of the Chapter. It
states that compactness and limit point compactness are equivalent in metric spaces.
Suppose that E K is innite. We will produce a limit point of E inside K. We begin
by dividing K into 2
n
sub-cells by cutting each interval [0, 1] into two equal pieces. For
instance, in R
2
we would consider the 4 sub-cells
[0, 1/2] [0, 1/2], [0, 1/2] [1/2, 1], [1/2, 1] [0, 1/2], [1/2, 1] [1/2, 1] .
At least one of these 2
n
sub-cells must contain innitely many points of E. Call this sub-cell
K
1
. Repeat, by dividing K
1
into 2
n
equal sub-cells to nd a sub-sub-cell K
2
which contains
innitely many points of E.
We continue this procedure ad innitum, at stage i 1 nding a sub-cell K
i
of K of the
form
K
i
= [r
1,i
2
i
, (r
1,i
+ 1)2
i
] [r
n,i
2
i
, (r
n,i
+ 1)2
i
]
which contains innitely many points of K. Note that the K
i
s satisfy the conditions of the
previous lemma: they are nested n-cells. Therefore there exists z
i
K
i
. Because each K
i
is a subset of K, we have z K.
We claim that z is a limit point of E. To show this, let r > 0. Note that for all points
x, y K
i
we have
[x y[
2
= (x
1
y
1
)
2
+ + (x
n
y
n
)
2
n(2
i
)
2
=
n
4
i
.
33
Therefore
diam(K
i
) = sup[x y[ : x, y K
i

n
2
i

n
i
.
(You can prove this inequality i 2
i
for all i by induction.) So x any i >

n
r
; then for all
x K
i
we have (because z K
i
)
[x z[ diam(K
i
)

n
i
< r ,
so that K
i
B
r
(z). However K
i
contains innitely many points of E, so we can nd one
not equal to z in B
r
(z). This means z is a limit point of E.
Proof of Heine-Borel. Suppose that K is closed and bounded in R
n
. Then there exists an
n-cell C such that K C. By the previous lemma, C is compact. But K is a closed subset
of C so K is compact.
Suppose conversely that K is compact. Then we have already shown K is closed and
bounded.
Therefore we nd that
(closed and bounded) (compact) in R
n
,
(closed and bounded) (compact) in metric spaces ,
and
(compact) (limit point compact) in metric spaces .
3.6 The Cantor set
We now give a famous example of a compact set. This is the Cantor set, and it has many
interesting properties. We construct it iteratively. At stage 0, we start with
C
0
= [0, 1] ,
the entire unit interval in R. At stage 1, we remove the middle third of C
0
to produce
C
1
=
_
0,
1
3
_

_
2
3
, 1
_
,
a set which is a union of 2 disjoint closed intervals, each of length 1/3. We continue, at stage
n, producing C
n
from C
n1
by removing the middle third of each interval which comprises
C
n1
. For example,
C
2
=
_
0,
1
9
_

_
2
9
,
1
3
_

_
2
3
,
7
9
_

_
8
9
, 1
_
.
It follows that at stage n, the set C
n
is a union of 2
n
disjoint closed intervals, each of length
3
n
. We then dene
C =

n=0
C
n
to be the Cantor set.
Properties.
34
1. C is closed because it is an intersection of closed sets.
2. C is compact because it is closed and bounded (in R).
3. C has total length 0. Although we have not dened this, we can compute the length
of C
n
: it is composed of 2
n
intervals of length 3
n
. Thus its length is (2/3)
n
. Because
this number tends to 0 as n goes to innity (dont worry we will dene these things
rigorously later),
length(C) length(C
n
) = (2/3)
n
for all n, giving length(C) = 0.
4. Although it looks like all that will remain in the end is the endpoints of the intervals
used to construct C, in fact there is much more. The set of such endpoints is countable,
whereas C is uncountable. To see why this is true, note that each x C can be given
an address. The point x is in C
1
, so it is in exactly one of the two intervals of C
1
;
assign the value 0 to x if it is in the rst and 1 if it is in the second. Similarly, the
set C
2
splits each interval of C
1
into two: give x the value 0 if it is in the left such
interval and 1 if it is in the right. Continuing in this way, we can assign to x an innite
sequence of 0s and 1s:
x 0111000110101 . . .
(In fact, this is nothing but the ternary expansion of x, replacing 2s by 1s.) The map
sending x to a sequence is actually a bijection from C to the set of sequences of 0s
and 1s, which we know is uncountable.
One example of an element of C that is not an endpoint is 1/4: its address is
1/4 010101 . . .
(Endpoints have 1-term repeating addresses, like 1/3 011111 . . .)
5. Every point of C is a limit point of C. To show this, we will prove more. We will show
that for each x C and each r > 0, there are points y and z in (x r, x + r) such
that y ,= x ,= z and y C, z / C. To do this, choose N such that N > 1/r and note
that since 2
N
> N we certainly have 3
N
> N, giving 3
N
< r. Since x C it follows
that x C
N
so there is some subinterval I of C
N
of length 3
N
such that x I. This
interval is necessarily contained in (x r, x + r) and so both endpoints of I (which
survive through the construction of C) are in this neighborhood. At least one of these
points is not equal to x, so we have found a point of C in (x r, x +r), giving that x
is a limit point of C.
6. C contains no open intervals. This proof proceeds like in the previous case. If x C
and r > 0 we can nd some N such that C
N
contains an interval I entirely contained
in (x r, x +r). In the next stage of the construction, we remove part of this interval
and so we can nd some y C
c
that is in (x r, x + r). So, each point of C is also
a limit point of C
c
. This implies that no point of C is an interior point, which in R
means that C contains no open intervals.
35
Lets nish with an observation. In exercise 6, you are asked to show that if (X, d) is
a metric space and (O
n
) is a countable collection of open dense subsets of X then

n=1
O
n
is nonempty. From this we can actually derive uncountability of the real numbers. Indeed,
assume for a contradiction that R is countable and list its elements as x
1
, x
2
, . . .. Dene
O
n
= R x
n
. Each O
n
is open and dense in R, so the intersection of all O
n
s is nonempty.
This is a contradiction, since

n
O
n
=

n
[R x
n
] = .
3.7 Exercises
1. For X = R
2
dene the function d : X X R by
d ((x
1
, y
1
), (x
2
, y
2
)) = [x
1
x
2
[ +[y
1
y
2
[ .
Prove that d is a metric on X. Describe the unit ball centered at the origin geometri-
cally. Repeat this question using
d ((x
1
, y
1
), (x
2
, y
2
)) = max[x
1
x
2
[, [y
1
y
2
[ .
2. Let F[0, 1] be the set of all bounded functions from [0, 1] to R. Show that d is a metric,
where
d(f, g) = sup[f(x) g(x)[ : x [0, 1] .
3. Let X be the set of real valued sequences with only nitely many nonzero terms:
X = x = (x
1
, x
2
, . . .) : x
i
R and x
i
,= 0 for only nitely many i .
For an element x X write n(x) for the largest i N such that x
i
,= 0. Dene the
function d : X X R by
d(x, y) =
_
_
max{n(x),n(y)}

i=1
(x
i
y
i
)
2
_
_
1/2
.
(a) Show that (X, d) is a metric space.
(b) For each n N dene e
n
as the element of X that has n-th coordinate equal to
1 and all others zero. Show that the set e
n
: n N is closed and bounded but
does not have a limit point.
4. For each of the following examples, verify that the collection ( is an open cover of E
and determine if it can be reduced to a nite subcover. If it can, give a nite subcover;
otherwise, show why it cannot be reduced.
(a) E = 1, 1/2, 1/4, . . . = 2
n
: n 0, ( = (2
n1
, 3 2
n1
) : n 0.
36
(b) E = [0, 1], ( = (x 10
4
, x + 10
4
) : x Q [0, 1].
5. Prove that an uncountable set E R cannot have countably many limit points.
Hint. Argue by contradiction and assume that there is a set E R such that E

, the
set of limit points of E, is countable. What can you say about E E

?
6. An open interval I R is a set of the form
(a, b) = x R : a < x < b for a b or
(a, ) = x R : a < x or
(, b) = x R : x < b or
R = (, ).
Let O R be a nonempty open set and let x O. Dene O
x
as the union of all open
intervals I such that x I and I O. Prove that O
x
is a nonempty open interval.
7. Let O R be a nonempty open set. By completing the following two steps, show that
there exists a countable collection ( of open intervals such that
for all I, J ( with I ,= J, we have I J = and (2)
_
IC
I = O . (3)
(a) For x O, let O
x
be dened as in exercise 3. Show that if O
x
O
y
,= for some
x, y O then O
x
= O
y
.
(b) Dene ( = O
x
: x O and complete the proof by showing that ( is countable
and has properties (2) and (3).
8. The Kuratowski closure and complement problem. For subsets A of a metric space X,
consider two operations, the closure A, and the complement A
c
= X A. We can
perform these multiple times, as in A
c
, (A)
c
, etc.
(a) Prove that, starting with a given A, one can form no more than 14 distinct sets by
applying the two operations successively.
(b) Letting X = R, nd a subset A R for which the maximum of 14 is attained.
Hint to get started. Clearly (A
c
)
c
= A, so two complements in a row get you nothing
new. What about two closures in a row? See Rudin, Thm. 2.27.
9. A subset E of a metric space X is dense if each point of X is in E or is a limit point
of E (or both). Let A
1
, A
2
, . . . be open dense sets in R. Show that
n
A
n
,= .
Hint. Dene a sequence of sets as follows. Choose x
1
A
1
and r
1
> 0 such that
B
r
1
(x
1
) A
1
. Then argue that there exists x
2
A
1
A
2
and r
2
> 0 such that
37
B
r
2
(x
2
) B
r
1
/2
(x
1
). Continuing, nd innite sequences r
1
, r
2
, . . . and x
1
, x
2
, . . . such
that x
n

n1
k=1
A
k
and B
rn
(x
n
) B
r
n1
/2
(x
n1
). Then, for each n, dene B
n
=
B
rn/2
(x
n
). What can you say about
n
B
n
?
10. Show that both Q and R Q are dense in R with the usual metric.
11. We now extend the denition of dense. If E
1
, E
2
are subsets of a metric space X then
E
1
is dense in E
2
if each point of E
2
is in E
1
or is a limit point of E
1
. Show that if
E
1
, E
2
, E
3
are subsets of X such that E
1
is dense in E
2
and E
2
is dense in E
3
then E
1
is dense in E
3
.
12. We say a metric space (X, d) has the nite intersection property if whenever ( is a col-
lection of closed sets in X such that each nite subcollection has nonempty intersection,
the full collection has nonempty intersection:

CC
C ,= .
Show that X has the nite intersection property if and only if X is compact. (You
may use Rudin, Theorem 2.36.)
13. Let / be a collection of compact subsets of a metric space X.
(a) Show that

KK
K is compact.
(b) Show that if / is nite then

KK
K is compact. Is this still true if / is innite?
14. Let (X, d) be a metric space. We say that a subset E of X is limit point compact if
every innite subset of E has a limit point in E. We have seen in class that if E is
compact then E is limit point compact. This exercise will serve to show the converse:
that if E is limit point compact then it is compact. For the following questions, x a
subset E that is limit point compact.
(a) Show that E is closed.
(b) Show that if > 0 then there exist nitely many points x
1
, . . . , x
n
in E such that
E
n
_
i=1
B

(x
i
) .
(c) Show that if A E is closed, then A is also limit point compact.
(d) Show that if A
1
, A
2
, . . . are closed subsets of E such that A
n
A
n+1
for all n 1
then
n
A
n
,= .
Hint. Dene a set x
n
: n 1 by choosing x
1
A
1
, x
2
A
2
, and so on.
38
(e) Use the previous parts to argue that E is compact.
Hint. Argue by contradiction and assume that there is an open cover ( of E that
cannot be reduced to a nite subcover. Begin with
1
= 1/2 and apply part (b)
to get points x
1
1
, . . . , x
1
n
1
E such that E
n
1
k=1
B

1
(x
1
k
). Clearly
E
n
1
k=1
_
B

1
(x
1
k
) E
_
.
At least one of these sets, say B

1
(x
1
j
1
) E cannot be covered by a nite number
of sets from (, or else we would have a contradiction. By parts (a) and (c), it has
the limit point property and ( is a cover of it, so repeat the construction using
this set instead of E and
2
= 1/4. Continue, at step n 3 using
n
= 2
n
, to
create a decreasing sequence of closed subsets of E. Use part (d).
15. In this exercise we will consider a construction similar to that of the Cantor set. We
will dene a countable collection of subsets E
n
: n 0 of the interval [0, 1] and we
will set E =

n=0
E
n
.
We dene E
0
= [0, 1], the entire interval. To dene E
1
, we remove a subinterval of E
0
of length 1/4 from the middle of E
0
. Precisely, we set
E
1
=
_
0,
3
8
_

_
5
8
, 1
_
.
Next, let E
2
be the set obtained by removing two subintervals, each of length 1/16,
from the middle of each piece of E
1
. Thus
E
2
=
_
0,
5
32
_

_
7
32
,
3
8
_

_
5
8
,
25
32
_

_
27
32
, 1
_
.
Continuing, at each step n 3, we create E
n
by removing 2
n1
subintervals, each of
length 4
n
, from the middle of each piece of E
n1
. Dene
E =

n=0
E
n
.
(a) Show that each point of E is a limit point of E.
(b) Show that E does not contain any open interval.
(c) What is the total length of E?
39
4 Sequences
4.1 Denitions
Denition 4.1.1. Let (X, d) be a metric space. A sequence is a function f : N X.
We think of a sequence as a list of its elements. We typically write x
1
= f(1), x
2
= f(2)
and forget about f, denoting the sequence as (x
n
) and the elements x
1
, x
2
, . . ..
The most fundamental notion related to sequences is that of convergence.
Denition 4.1.2. A sequence (x
n
) converges to a point x X if for every > 0 there exists
N such that if n N then
d(x
n
, x) < .
In this case we write x
n
x.
We can think of proving convergence of a sequence as follows. We have a sequence (x
n
)
and you tell me it has a limit x. I ask Oh yeah? Well can you show that the terms of the
sequence get very close to x? You say yes and I ask Can you show that all but nitely
many terms are within distance = 1 of x? You say yes and provide an N equal to 600.
Then you proceed to show me that all x
n
for n 600 have d(x
n
, x) < 1. Temporarily
satised, I ask, Well you did it for 1, what about for = .00001? You then dream up of
an N equal to 40 billion such that for n N, d(x
n
, x) < .00001. This game can continue
indenitely, and as long as you can come up with an N for each of my values of , then we
say x
n
converges to x.
Example.
We all believe that the sequence (x
n
) given by x
n
=
1
n
2
+n
(in R) converges to 0, How do
we prove it? Let > 0. We want [x
n
0[ < , so we solve:
1
n
2
+ n
< , which is equivalent to n
2
+ n >
1

.
This will certainly be true if n >
1

, so set
N =
_
1

_
.
Now if n N then
1
n
2
+ n

1
N
2
+ N
<
1
N
2
< .
In the previous example, to show convergence to something we could have noticed that
the sequence is monotonic and bounded.
Denition 4.1.3. A sequence (x
n
) in R is
1. monotone increasing if x
n
< x
n+1
for all n (monotone non-decreasing if x
n
x
n+1
)
and
40
2. monotone decreasing if x
n
> x
n+1
for all n (monotone non-increasing if x
n
x
n+1
).
Theorem 4.1.4. If (x
n
) is monotone (any of the types above) and bounded (that is, x
n
:
n N is bounded) then it converges.
Proof. Suppose (x
n
) is monotone increasing. The other cases are similar. Then
X := x
n
: n N
is nonempty and bounded above so it has a supremum x. We claim that x
n
x. To prove
this, let > 0. Then x is not an upper bound for X and there exists N such that
x
N
> x . Then if n N,
x x
n
x
N
> x ,
giving [x x
n
[ < , so x
n
x.
We now recall some basic properties of limits. For one part we need a denition.
Denition 4.1.5. A sequence (x
n
) in a metric space X is bounded if there exists q X and
m R such that
d(p
n
, q) M for all n N .
Note that this is the same as saying that the set x
n
: n N is bounded.
Theorem 4.1.6 (Rudin, Theorem 3.2). Let (x
n
) be a sequence in a metric space X.
1. (x
n
) converges to x X if and only if every neighborhood of x contains x
n
for all but
nitely many n.
2. (Uniqueness of the limit) If x, y X and (x
n
) converges to both x and y, then x = y.
3. If (x
n
) converges then (x
n
) is bounded.
4. If E X and x is a limit point of E then there is a sequence (x
n
) in E converging to
x.
Proof. Part 1 is just a restatement of the denition of a limit. For the second part, suppose
that (x
n
) converges to x and to y. Let > 0, so that there exists N
1
and N
2
such that
if n N
1
then d(x
n
, x) < /2
and if n N
2
then d(x
n
, y) < /2 .
Using the triangle inequality, for N = maxN
1
, N
2
, we get
d(x, y) d(x, x
N
) + d(x
N
, y) < /2 + /2 = .
Thus d(x, y) < for all > 0; this is only possible if d(x, y) = 0 and thus x = y.
41
For part 3, suppose that (x
n
) converges to x X and let = 1. Then there exists N
such that if n N then d(x, x
n
) < 1. Now choose
r = max1, d(x, x
1
), . . . , d(x, x
N1
) .
It follows then that d(x, x
n
) < r for all n, and so the set x
n
: n N is contained in B
r
(x).
We now show part 4. Suppose x is a limit point of E. For each n, choose any point (call
it x
n
) in the set B
1/n
(x) E. We claim that this sequence of points (x
n
) converges to x. To
see this, let > 0 and pick
N =
_
1

_
.
Then if n N, d(x
n
, x) < 1/n 1/N < .
In the above, we see that a limit of a sequence is unique. This is in contrast to the limit
points (plural!) of a subset E of X. The points in E are in no particular order, and E may
have many limit points. But in a sequence, the points are ordered, and there can be at most
one limit as n runs through that chosen order.
In the case that the sequence is of real numbers, there is a nice compatibility with
arithmetic operations.
Properties of real sequences. Let (x
n
) and (y
n
) be real sequences such that x
n
x and
y
n
y.
1. x
n
+ y
n
x + y.
2. If c R then cx
n
cx.
3. x
n
y
n
xy.
4. If y ,= 0 and y
n
,= 0 for all n N then
xn
yn

x
y
.
Proofs of properties. Many of these are similar so we will prove only 1 and 3. Rudin contains
all of the proofs. Suppose rst that x
n
x and y
n
y. Given > 0 choose N
1
and N
2
such that
if n N
1
then [x
n
x[ < /2 and
if n N
2
then [y
n
y[ < /2 .
Letting N = maxN
1
, N
2
, then if n N we have
[x
n
+ y
n
(x + y)[ [x
n
x[ +[y
n
y[ < .
For the third part we write
[x
n
y
n
xy[ [y
n
[[x
n
x[ +[x[[y
n
y[ .
Now note that since (y
n
) converges, it is bounded. Therefore we can nd M > 0 such that
[x[ M and [y
n
[ M for all n. Given > 0 choose N such that if n N then both
[x
n
x[ /(2M) and [y
n
y[ /(2M) .
42
Then if n N,
[x
n
y
n
xy[ M/(2M) + M/(2M) = .
Note that for the last item above we required y
n
,= 0 for all n. This is not necessary for
the following reasons.
Lemma 4.1.7. If (y
n
) is a real sequence such that y
n
y and y ,= 0 then y
n
= 0 for at
most nitely many n N.
Proof. Suppose that y
n
y with y ,= 0 and let = [y[. Then there exists N N such that
if n N then [y
n
y[ < . By the triangle inequality, if n N then
[y
n
[ [y[ [y
n
y[ > 0 ,
giving y
n
,= 0.
The next lemma says that if we remove a nite number of terms from a convergent
sequence, this does not aect the limit.
Lemma 4.1.8. Let (y
n
) be a sequence in a metric space X. For a xed k N dene a
sequence (z
n
) by
z
n
= y
n+k
for n N .
Then (y
n
) converges if and only if (z
n
) does. If y
n
y then z
n
y.
Proof. Suppose y
n
y. If > 0 we can pick N N such that d(y
n
, y) < for n N. For
n N,
d(z
n
, y) = d(y
n+k
, y) < ,
since n + k N also. This means z
n
y.
Conversely, if z
n
y then given > 0 we can nd N N such that n N implies that
d(z
n
, y) < . Dene N

= N + k. Then if n N

, we have n k N and so
d(y
n
, y) = d(z
nk
, y) < .
Thus y
n
y.
Now we can change the last property of real sequences as follows. If (x
n
) and (y
n
) are
real sequences such that x
n
x and y
n
y with y ,= 0 then x
n
/y
n
x/y. To do this,
we use the rst lemma to nd k such that for all n, y
n+k
,= 0. Then we can consider the
sequences (x
n+k
) and (y
n+k
) and prove the property for them. Since they only dier from
(x
n
) and (y
n
) by a nite number of terms, the property also holds for (x
n
) and (y
n
).
We will mostly deal with sequences of real numbers (or elements of an arbitrary metric
space), but it is useful to understand convergence in R
k
, k 2. It can be reformulated
in terms of convergence of each coordinate. That is, if (x
n
) is a sequence in R
k
, we write
x
n
= (x
(1)
n
, . . . , x
(k)
n
). The sequence (x
n
) converges to x R
k
if and only if each coordinate
sequence (x
(j)
n
) converges to x
(j)
, the j-th coordinate of x.
43
Theorem 4.1.9. Let (x
n
) and (y
n
) be sequences in R
k
and (
n
) a sequence of real numbers.
1. (x
n
) converges to x R
k
if and only if x
(j)
n
x
(j)
(in R) for all j = 1, . . . , k.
2. If x
n
x, y
n
y in R
k
and
n
in R, then
x
n
+ y
n
x + y, x
n
y
n
x y, and
n
x
n
x ,
where

is the standard dot product in R


k
.
Proof. The second part follows from the rst part and properties of limits in R we discussed
above. To prove the rst, suppose that x
n
x and let j 1, . . . , k. Given > 0, let N
be such that n N implies that [x
n
x[ < . Then we have
[x
(j)
n
x
(j)
[ =
_
(x
(j)
n
x
(j)
)
2

_
(x
(1)
n
x
(1)
)
2
+ + (x
(k)
n
x
(k)
)
2
= [x
n
x[ < .
So x
(j)
n
x
(j)
.
For the converse, suppose that x
(j)
n
x
(j)
for all j = 1, . . . , k and let > 0. Pick
N
1
, . . . , N
k
such that for j = 1, . . . , k, if n N
j
then [x
(j)
n
x
(j)
[ < /

k. then for
N = maxN
1
, . . . , N
k
and n N, we have
[x
n
x[ =
_
(x
(1)
n
x
(1)
)
2
+ + (x
(k)
n
x
(k)
)
2
<
_

2
/d + +
2
/d = .
We nish this section with the idea of convergence to innity.
Denition 4.1.10. A real sequence (x
n
) converges to if for each M > 0 there exists
N N such that
n N implies x
n
> M .
It converges to if (x
n
) converges to .
As before we write x
n
(or x
n
) in this case. In this denition we think of
M as taking the role of from before and we imagine that (M, ) is a neighborhood of
innity.
Clearly a sequence that converges to innity needs to be unbounded. The converse is not
true. Consider (x
n
), dened by
x
n
=
_
1 n odd
n n even
.
This sequence does not converge to innity, but it is unbounded.
44
4.2 Subsequences, Cauchy sequences and completeness
We now move back to sequences in general metric spaces. Sometimes the sequence does not
converge, but if we remove many of the terms we can make it converge. Another way to say
this is that a sequence might not converge but it may have a convergent subsequence.
Denition 4.2.1. Let (x
n
) be a sequence in a metric space X. Given a monotonically
increasing sequence (n
k
) in N (that is, the sequence is such that n
1
< n
2
< ) then the
sequence (x
n
k
) is called a subsequence of (x
n
). If x
n
k
y as k then we call y a
subsequential limit of (x
n
).
Note that a sequence (x
n
) converges to x if and only if each subsequence of (x
n
) converges
to x. To prove this, suppose rst that x
n
x and let (x
n
k
) be a subsequence. Given > 0
we can nd N such that if n N then d(x
n
, x) < . Because (n
k
) is monotone increasing, it
follows that n
k
k for all k, so choose K = N. Then for k K, the element x
n
k
is a term
of the sequence (x
n
) with index at least equal to N, giving d(x
n
k
, x) < .
Conversely, suppose that each subsequence of (x
n
) converges to x. Then as (x
n
) is a
subsequence of itself, we also have x
n
x!
The next theorem is one of the most important in the course. It is a restatement of
compactness; in general topological spaces, it is called sequential compactness.
Theorem 4.2.2. Let (x
n
) be a sequence in a compact metric space X. Then some subse-
quence of (x
n
) converges to a point x in X.
Proof. It may be that the set of sequence elements x
n
: n N is nite. In this case, at
least one of these elements must appear in the sequence innitely often. That is, there exists
x x
n
: n N and a monotone increasing sequence (n
k
) such that x
n
k
= x for all k.
Clearly then x
n
k
x and the element x X because the sequence terms are.
Otherwise, the set x
n
: n N is innite. Because compactness implies limit point com-
pactness, there exists x X which is a limit point of this set. Then we build a subsequence
that converges to x as follows. Since d(x
n
, x) < 1 for innitely many n, we can pick n
1
such
that d(x
n
1
, x) < 1. Continuing in this fashion, at stage i we note that d(x
n
, x) < 1/i for
innitely many n, so we can pick n
i
> n
i1
such that d(x
n
i
, x) < 1/i. Because n
1
< n
2
< ,
the sequence (x
n
i
) is a subsequence of (x
n
). Further, given > 0, choose I > 1/, so that if
i I,
d(x
n
i
, x) 1/i 1/I < .
Corollary 4.2.3 (Bolzano-Weierstrass). Each bounded sequence in R
k
has a convergent
subsequence.
Proof. If (x
n
) is bounded in R
k
then we can t the set x
n
: n N into a k-cell, which is
compact. Now, viewing (x
n
) as a sequence in this compact k-cell, we see by the previous
theorem that it has a convergent subsequence.
45
The last topic of the section is Cauchy sequences. The motivation is as follows. Many
times we are in a metric space X that has holes. For instance, we may consider Q as a
metric space inside of R (that is, using the metric d(x, y) = [x y[ from R). In this space,
the sequence
(1, 1.4, 1.41, 1.414, . . .)
does not converge (it should only converge in R to

2, but this element is not in our


space). Although we cannot talk about this sequence converging; that is, getting close to
some limit x, we can do the next best thing. We can say that the terms of the sequence get
close to each other.
Denition 4.2.4. Let (x
n
) be a sequence in a metric space X. We say that (x
n
) is Cauchy
if for each > 0 there exists N N such that
if m, n N then d(x
m
, x
n
) < .
Just like before, the number N gives us a cuto in the sequence after which all terms
are close to each other. Each convergent sequence (x
n
) (with some limit x) is Cauchy, for if
> 0 then we can pick N such that if n N then d(x
n
, x) < /2. Then for m, n N,
d(x
n
, x
m
) d(x
n
, x) + d(x
m
, x) < /2 + /2 = .
One reason a Cauchy sequence might not converge was illustrated above; the limit may
not be in the space. This is not possible, though, in a compact space.
Theorem 4.2.5. If X is a compact metric space, then all Cauchy sequences in X converge.
Proof. Let X be compact and (x
n
) a Cauchy sequence. By the previous theorem, (x
n
) has
a subsequence (x
n
k
) such that x
n
k
x, some point of X. We will show that since (x
n
)
is already Cauchy, the full sequence must converge to x. The idea is to x some element
x
n
k

of the subsequence which is close to x. This term is chosen far enough along the initial
sequence so that all terms are close to it, and thus close to x.
Let > 0 and choose N such that if m, n N then d(x
m
, x
n
) < /2. Choose also some
K such that if k K then d(x
n
k
, x) < /2. Last, set N

= maxN, K. Because (n
k
) is
monotone increasing, we can x k

such that n
k
N

. Then for any n N

, we have
d(x
n
, x) d(x
n
, x
n
k

) d(x
n
k

, x) /2 + /2 = .
Denition 4.2.6. A metric space X in which all Cauchy sequences converge is said to be
complete.
The above theorem says that compact spaces are complete. This is also true of R
k
,
though it is not compact.
Theorem 4.2.7. R
k
(with the usual metric) is complete.
46
Proof. Let (x
n
) be a Cauchy sequence. We claim that it is bounded. The proof is almost
the same as that of the fact that a convergent sequence is bounded. We can nd N such
that if n, m N then d(x
n
, x
m
) < 1. Therefore d(x
n
, x
N
) < 1 for all n N. Putting
R = maxd(x
N
, x
1
), . . . , d(x
N
, x
N1
), 1, we then have d(x
j
, x
N
) < R for all j, so (x
n
) is
bounded.
Since (x
n
) is bounded, we can put it in a k-cell C. Then we can view the sequence
as being in the space C, which is compact. Now we use the fact that compact spaces are
complete, giving some x C such that x
n
x. But x R, so we are done.
4.3 Special sequences
Here we will list the limits Rudin gives in the book and prove a couple of them. One
interesting one is the fourth. Imagine taking = 10
10
and p = 10
10
. Then it says that
(1 + 10
10
)
n
eventually outgrows n
10
10
.
Theorem 4.3.1. The limits below evaluate as follows.
If p > 0 then n
p
0.
If p > 0 then p
1/n
1.
n
1/n
1.
If p > 0 and R then
n

(1+p)
n
0.
If [x[ < 1 then x
n
0.
Proof. For the rst limit, let > 0 and choose N = ,
1

1/p
| + 1. Then if n N we have
n >
1

1/p
and therefore n
p
< .
For the second, Rudin uses the binomial theorem:
Lemma 4.3.2. For x, y R and n N,
(x + y)
n
=
n

j=0
_
n
j
_
x
j
y
nj
.
Proof. The proof is by induction. For n = 0 we get
1 =
_
0
0
_
= 1 .
47
Assuming it holds for some n, we show it holds for n + 1. We have
(x + y)
n+1
= (x + y)(x + y)
n
= (x + y)
n

j=0
_
n
j
_
x
j
y
nj
=
n

j=0
_
n
j
_
x
j+1
y
nj
+
n

j=0
_
n
j
_
x
j
y
n+1j
=
n+1

j=1
_
n
j 1
_
x
j
y
n+1j
+
n

j=0
_
n
j
_
x
j
y
n+1j
=
_
n
0
_
y
n+1
+
_
n
n
_
x
n+1
+
n

j=1
__
n
j 1
_
+
_
n
j
__
x
j
y
n+1j
But now we use the identity
_
n
j1
_
+
_
n
j
_
=
_
n+1
j
_
, valid for n 0 and j = 1, . . . , n. This gives
y
n+1
+ x
n+1
+
n

j=1
_
n + 1
j
_
x
j
y
n+1j
,
which is

n+1
j=0
_
n+1
j
_
x
j
y
n+1j
.
Returning to the proof of the second limit, we rst assume that p > 1 and set y
n
= p
1/n
1.
Computing,
p = (y
n
+ 1)
n
1 + ny
n
,
where we have taken only the rst two terms from the binomial theorem. This means
0 y
n

p1
n
and letting n we get y
n
0, completing the proof in the case p > 1.
If 0 < p < 1 then we consider 1/p and see that (1/p)
1/n
1. Taking reciprocals, we get
p
1/n
1.
For the third limit, we use a dierent term in the binomial theorem. Set x
n
= n
1/n
1
and compute
n = (1 + x
n
)
n

_
n
2
_
x
2
n
=
n(n 1)
2
x
2
n
,
so 0 x
n

_
2
n1

_
4
n
if n 2. Since n
1/2
0 we are done.
The fourth limit is a bit more dicult. Choose any k and consider n > 2k. Then
(1 + p)
n

_
n
k
_
p
k
=
n(n 1) (n k + 1)
k!
p
k

(n/2)
k
p
k
k!
.
This gives
0
n

(1 + p)
n

(p/2)
k
k!
n
k
0
since < k.
The last limit is proved in Chapter 5 in the theorem on geometric series.
48
4.4 Exercises
1. For the following sequences, nd the limit and prove your answer (using an N
argument).
(a) x
n
=

n
2
+ 1 n, n N.
(b) x
n
= n2
n
, n N.
2. Determine whether or not the following sequence converges. If it does not, give a
convergent subsequence (if one exists).
x
n
= sin
_
n
2
_
+ cos(n), n N .
3. Let a
1
, . . . , a
k
be positive numbers. Show that
lim
n
_
a
n
1
+ + a
n
k
k
_
1/n
= maxa
1
, . . . , a
k
.
4. We have seen that if a metric space (X, d) is compact then it must be complete. In
this exercise we investigate the converse.
(a) Show that if X is complete then it need not be compact.
(b) We say that a metric space (X, d) is totally bounded if for each > 0 we can nd
nitely many points x
1
, . . . , x
n
such that
X
n
_
i=1
B

(x
i
) .
(Here, B
r
(x) is the neighborhood y X : d(x, y) < r.) Show that X is compact
if and only if X is both totally bounded and complete via the following steps.
i. Show that if X is compact then X is totally bounded.
ii. Assume that X is totally bounded and let E X be innite. We will try to
construct a limit point for E in X. Begin by nding x
(1)
1
, . . . , x
(1)
n
1
X such
that
X
n
1
_
i=1
B1
2
(x
(1)
i
) .
There must be k
1
1, . . . , n
1
such that
E
1
= B1
2
(x
(1)
k
1
) E is innite .
Continue, at stage n 2 choosing x
(n)
kn
X such that E
n
= B
2
n(x
(n)
kn
) E
n1
is innite. Show that (x
(1)
k
1
, x
(2)
k
2
, . . .) is a Cauchy sequence.
Hint. You may want to use the following fact (without proof). For n 1,
dene s
n
= 1/2 + 1/4 + + 1/2
n
. Then (s
n
) converges.
49
iii. Assume that X is totally bounded and complete and show that any E X
which is innite has a limit point in X. Conclude that X is compact.
(c) Show that if X is totally bounded then it need not be compact.
5. Sometimes we want to analyze sequences that do not converge. For this purpose we
dene upper and lower limits; numbers that exist for all real sequences. Let (a
n
) be a
sequence in R and for each n 1 dene
u
n
= supa
n
, a
n+1
, . . . and l
n
= infa
n
, a
n+1
, . . . .
(Here we write u
n
= if a
k
: k n is not bounded above and l
n
= if it is not
bounded below.)
(a) Show that (u
n
) and (l
n
) are monotonic. If (a
n
) is bounded, show that there exist
numbers u, l R such that u
n
u and l
n
l. We denote these numbers as the
limit superior (upper limit) and limit inferior (lower limit) of (a
n
) and write
limsup
n
a
n
= u, liminf
n
a
n
= l .
(b) Give reasonable denitions of limsup
n
a
n
and liminf
n
a
n
in the unbounded
case. (Here your denitions should allow for the possibilities .)
(c) Show that (a
n
) converges if and only if
limsup
n
a
n
= liminf
n
a
n
.
(You may want to separate into cases depending on whether the liminf and/or
limsup is nite or innite.)
6. Let (a
n
) be a real sequence and write E for the set of all subsequential limits of (a
n
);
that is,
E = x R : there is a subsequence (a
n
k
) of (a
n
) such that a
n
k
x as k .
Assume that (a
n
) is bounded and prove that limsup
n
a
n
= sup E. Explain how you
would modify your proof to show that liminf
n
a
n
= inf E. (These results are also
true if (a
n
) is unbounded but you do not have to prove that.)
7. Let a
0
, b
0
be real numbers with 0 b
0
a
0
. Dene sequences a
n
, b
n
by
a
n+1
=
a
n
+ b
n
2
and b
n+1
=
_
a
n
b
n
for n 0 .
(a) Prove the arithmetic-geometric mean inequality: b
1
a
1
. This trivially extends
to b
n
a
n
for all n.
(b) Prove that a
n
and b
n
converge, and to the same limit. This limit is called
the arithmetic-geometric mean of a
0
and b
0
.
50
8. This problem is not assigned; it is just for fun. Let x
0
= 1 and x
n+1
= sin x
n
.
(a) Prove x
n
0.
(b) Find lim
n

nx
n
.
51
5 Series
We now introduce series, which are special types of sequences. We will concentrate on them
for the next couple of lectures.
5.1 Denitions
Denition 5.1.1. Let (x
n
) be a real sequence. For each n N, dene the partial sum
s
n
= x
1
+ + x
n
=
n

j=1
x
j
.
We say that the series

x
n
converges if (s
n
) converges.
Just as before, the tail behavior is all that matters (we can chop o as many initial terms
as we want). In other words

n=1
x
n
converges i

n=N
x
n
converges for each N 1 .
The proof is the same as that for sequences.
We would like to characterize which series converge. We start with a simple criterion
that must be satised.
Theorem 5.1.2. If the series

x
n
converges then the terms (x
n
) converge to 0.
Proof. Let > 0 and suppose that

n
x
n
= s (that is, s
n
s). Then there exists N N
such that if n N then [s
n
s[ < /2. Now for n N + 1,
[s
n
s[ < /2 and [s
n1
s[ < /2 ,
implying that [s
n
s
n1
[ < . Therefore
[x
n
0[ = [s
n
s
n1
[ <
and we are done.
The above tells us that many series cannot converge. For instance,

x
n
diverges, where
x
n
= (1)
n
. However, it is not true that all series

x
n
with x
n
0 converge. For example,
Theorem 5.1.3. The harmonic series,

n=1
1/n, diverges.
Proof. To prove this, we give a lemma that allows us to handle series of non-negative terms
more easily.
Lemma 5.1.4. Let (x
n
) be a sequence of non-negative terms. Then

x
n
converges if and
only if the sequence of partial sums (s
n
) is bounded.
52
Proof. This comes directly from the monotone convergence theorem. If x
n
0 for all n,
then
s
n+1
= s
n
+ x
n+1
s
n
,
giving that (s
n
) is monotone, and converges if and only if it is bounded.
Returning to the proof, we will show that the partial sums of the harmonic series are
unbounded. Let M > 0 and choose n of the form n = 2
k
for k > 2M. Then we give a lower
bound:
s
n
= 1 +
1
2
+
1
3
+ +
1
2
k
> 1 +
_
1
2
+
1
3
_
+
_
1
4
+
1
5
+
1
6
+
1
7
_
+ +
_
1
2
k1
+ +
1
2
k
1
_
>
1
2
+ 2
_
1
4
_
+ 4
_
1
8
_
+ + 2
k1
_
1
2
k
_
=
1
2
(1 + 1 + 1 + + 1) =
k
2
> M .
So given any M > 0, there exists n such that s
n
> M. This implies that (s
n
) is unbounded
and we are done.
In the proof above, we used an argument that can be generalized a bit.
Theorem 5.1.5 (Comparison test). Let (x
n
) and (y
n
) be non-negative real sequences such
that x
n
y
n
for all n.
1. If

x
n
converges, then so does

y
n
.
2. If

y
n
diverges, then so does

x
n
.
Proof. The rst part is implied by the second, so we need only show the second. Write (s
n
)
and (t
n
) for the partial sums
s
n
= x
1
+ + x
n
and t
n
= y
1
+ + y
n
.
Since y
n
0 for all n we can use the above lemma to say that (t
n
) is unbounded, so given
M > 0 choose N such that n N implies that t
n
> M. Now for such n,
s
n
= x
1
+ + x
n
y
1
+ + y
n
= t
n
> M ,
so that (s
n
) is unbounded and diverges.
This test can be generalized in at least two ways:
1. We only need x
n
y
n
for n greater than some N
0
. This is because we can consider
convergence/divergence of

n=N
0
x
n
etc.
53
2. In the rst part, we do not even need y
n
0 as long as we modify the statement.
Suppose that (x
n
) is non-negative such that

x
n
converges and [y
n
[ x
n
for all n.
Then setting s
n
and t
n
as before, we can just show that (t
n
) is Cauchy. Since (s
n
) is,
given > 0 we can nd N such that if n > m N then [s
n
s
m
[ < . Then
[t
n
t
m
[ = [y
m+1
+ + y
n
[ [y
m+1
[ + +[y
n
[ x
m+1
+ + x
n
= [s
n
s
m
[ < .
To use the comparison test, let us rst introduce one of the simplest series of all time.
Theorem 5.1.6 (Geometric series). For a R dene a sequence (x
n
) by x
n
= a
n
. Then the
geometric series

n
x
n
converges if and only if [a[ < 1. Furthermore,

n=0
a
n
=
1
1 a
if [a[ < 1 .
Proof. The rst thing to note is that a
n
0 if [a[ < 1. We can prove this by showing that
[a[
n
0. So if 0 [a[ < 1 then the sequence [a[
n
is monotone decreasing:
[a[
n+1
= [a[[a[
n
[a[
n
and is bounded below so has a limit, say L. Then we get
L = lim
n
[a[
n
= [a[ lim
n
[a[
n1
= [a[L ,
but as [a[ , = 0 we have L = 0.
Now continue to assume that [a[ < 1 and compute the partial sum for n 1
s
n
+ a
n+1
= s
n+1
= 1 + a + a
2
+ + a
n+1
= 1 + a(1 + a + a
2
+ + a
n
) = 1 + as
n
.
Solving for s
n
, we nd s
n
(1 a) = 1 a
n+1
. Since a ,= 1,
s
n
=
1 a
n+1
1 a
.
We let n to get the result.
If [a[ 1 then the terms a
n
do not even go to zero, since [a[
n
[a[ ,= 0, so the series
diverges.
Now we can prove facts about the p-series.
Theorem 5.1.7. The series

n
p
converges if and only if p > 1.
54
Proof. For p 1 we have n
p
1/n and so the comparison test gives divergence. Suppose
then that p > 1. We can group terms as before: taking n = 2
k
1,
1 +
1
2
p
+
1
3
p
+ +
1
(2
k
1)
p
= 1 +
_
1
2
p
+
1
3
p
_
+
_
1
4
p
+
1
5
p
+
1
6
p
+
1
7
p
_
+ +
_
1
2
p(k1)
+ +
1
(2
k
1)
p
_
1 + 2
_
1
2
p
_
+ 4
_
1
4
p
_
+ + 2
k1
_
1
2
(k1)p
_
= 1 + 2
1p
+ 4
1p
+ + (2
k1
)
1p
= (2
1p
)
0
+ (2
1p
)
1
+ + (2
1p
)
k1

1
1 2
1p
< since p > 1 .
This means that if s
n
=

n
j=1
j
p
, then s
2
k
1

1
12
1p
for all k. Since (s
n
) is monotone, it
is then bounded and the series converges.
5.2 Ratio and root tests
We will continue to deal with series of non-negative terms. So far we have used the compar-
ison test to compare series to the geometric series, which we could solve for magically. We
will still do that, but in a more rened way.
Theorem 5.2.1 (Root test). Let (x
n
) be a real sequence and dene = limsup
n
[x
n
[
1/n
.
If < 1 then

x
n
converges.
If > 1 then

x
n
diverges.
If = 1 then

x
n
could converge or diverge.
Proof. Dene
u
n
= sup[x
n
[
1/n
, [x
n+1
[
1/n+1
, . . . .
In the homework set, limsup
n
x
n
was dened as lim
n
u
n
(at least in the bounded case
you can adapt this proof to the unbounded case). If < 1 this means that given p (, 1)
there exists N such that if n N then u
n
p, or
n N implies [x
n
[ p
n
.
Now we just use the comparison test. Since

p
n
converges (as 0 < p < 1), so does

x
n
.
Suppose now that > 1. Recall from the homework that given a real sequence (y
n
),
there always exists a subsequence (y
n
k
) such that y
n
k
limsup
n
y
n
. So we can nd an
increasing sequence (n
k
) such that [x
n
k
[
1/n
k
. Thus there exists K such that
k K implies [x
n
k
[ 1
n
k
= 1
and since (x
n
) does not converge to zero, we cannot have

x
n
convergent.
55
Last if = 1 we cannot tell anything. First (1/n)
1/n
1 but also (1/n
2
)
1/n
=
_
(1/n)
1/n
_
2
1. Since

1/n diverges and

1/n
2
converges, the root test tells us noth-
ing.
Applications.
1. The series

n
2
2
n
converges. We can see this by the root test.
limsup
n
_
n
2
2
n
_
1/n
= limsup
n
_
n
1/n
_
2
2
= 1/2 < 1 .
2. Power series. Let x R and for a given real sequence (a
n
), consider the series

n=0
a
n
x
n
.
We would like to know for which values of x this series converges. To solve for this, we
simply use the root test. Consider
limsup
n
[a
n
x
n
[
1/n
= [x[ limsup
n
[a
n
[
1/n
.
Setting = limsup
n
[a
n
[
1/n
, we nd that the series converges if [x[ < 1/ and
diverges if [x[ > 1/. So it makes sense to dene
R := 1/ as the radius of convergence of the power series

n=0
a
n
x
n
.
Of course we cannot tell from the root test what happens when x = R.
3. Consider the series

n=0
1
n!
. We have
_
1
n!
_
1/n

_
1
n(n 1) (,n/2|)
_
1/n

_
1
(n/2)
n/2
_
1/n
=
_
2
n
0 .
So the root test gives convergence.
Another useful test is the ratio test.
Theorem 5.2.2 (Ratio test). Let (x
n
) be a real sequence.
If limsup
n

x
n+1
xn

< 1 then

x
n
converges.
If x
n+1
x
n
> 0 for all n N
0
(a xed natural number) then

x
n
diverges.
56
Proof. Assume the limsup is < 1. Then as before, choosing p (, 1) we can nd N such
that if n N then [x
n+1
[ < p[x
n
[. Iterating this from n = N we nd
[x
N+k
[ < p
k
[x
N
[ for all k 1 .
Therefore if we set y
k
= x
N+k
then [y
k
[ Cp
k
with C a non-negative constant equal to x
N
.
This implies by the comparison test that

y
k
converges. This is the tail of

x
n
so this
converges as well.
Suppose on the other hand that x
n+1
x
n
> 0 for all n N
0
. Then by iteration,
x
N
0
+k
x
N
0
> 0 ,
and the terms do not even converge to 0. This implies that

x
n
diverges.
The ratio test can be inconclusive, but in more ways than can the root test. First if
limsup
n

x
n+1
xn

> 1 it can still be that the series converges (try to think of an example!).
Also if this limsup equals 1, we could have convergence or divergence.
Note however that if
lim
n

x
n+1
x
n

exists and is > 1


then we can apply the second criterion of the ratio test and conclude divergence of

x
n
.
Applications.
1. The series

x
n n
n
n!
converges if [x[ < 1/C, where C =

n=0
1/n!. To see this, set
b
n
= x
n n
n
n!
:

b
n+1
b
n

= [x[
(n + 1)
n+1
n
n
(n + 1)
= [x[
_
1 +
1
n
_
n
.
But
_
1 +
1
n
_
n
=
n

j=0
_
n
j
_
n
j
=
n

j=0
n(n 1) (n j + 1)
j!n
j

j=0
1
j!
= C .
So limsup
n

b
n+1
bn

C[x[ < 1.
2. Power series. Generally we can also test convergence of power series using the ratio
test. Considering the series

a
n
x
n
, we compute
limsup
n

a
n+1
x
n+1
a
n
x
n

= [x[ limsup
n

a
n+1
a
n

= [x[ ,
where = limsup
n

a
n+1
an

. So if [x[ < 1/ the series converges, whereas if [x[ 1/


we cannot tell. However, if = lim
n

a
n+1
an

exists then for [x[ > 1/ we have


divergence.
57
Remark (from class). The root and ratio tests can give dierent answers. Consider the
sequence (a
n
) given by
a
n
=
_
1 if n is even
2 if n is odd
.
Then
limsup
n

a
n+1
a
n

= 2 but limsup
n
([a
n
[)
1/n
= 1 .
Therefore if we consider the radius of convergence of

a
n
x
n
, the root test gives 1. If we
were to dene the radius of convergence using the ratio test (which we should not!) then
we would get 1/2, which is smaller, and is not accurate, since for x = 3/4, for instance, the
series converges. Generally speaking we have
limsup
n
([a
n
[)
1/n
limsup
n

a
n+1
a
n

.
See Rudin, Theorem 3.37.
5.3 Non non-negative series
We saw in the last couple of lectures that when dealing with series with non-negative terms,
we compare to the geometric series. This gave rise to the comparison test, the ratio test and
the root test. For other series we have basically one tool: summation by parts. It comes in
to the following theorem.
Theorem 5.3.1 (Dirichlet test). Let (a
n
) and (b
n
) be real sequences such that, setting A
n
=

n
j=0
a
j
, we have (A
n
) bounded. If (b
n
) is monotonic with b
n
0 then

a
n
b
n
converges.
Proof. Lets suppose that (b
n
) is monotone decreasing; the other case is similar. The idea is
to get a dierent representation for

a
n
b
n
by setting A
n
=

n
j=0
a
j
for n 0 and A
1
= 0.
Now
N

n=0
a
n
b
n
=
N

n=0
(A
n
A
n1
)b
n
=
N

n=0
A
n
b
n

n=0
A
n1
b
n
=
N

n=0
A
n
b
n

N1

n=0
A
n
b
n+1
=
N

n=0
A
n
(b
n
b
n+1
) + A
N
b
N
.
Now since (A
n
) is bounded and b
n
0 we have A
n
b
n
0. We can show this as follows.
Suppose that [A
n
[ M for all n and let > 0. Choose N
0
such that n N
0
implies that
[b
n
[ < /M. Then for n N
0
, [A
n
b
n
[ < M/M = .
58
Since A
N
b
N
0, the above representation gives that

a
n
b
n
converges if and only
if

A
n
(b
n
b
n+1
) converges. But now we use the comparison test: [A
n
(b
n
b
n+1
)[
M[b
n
b
n+1
[ = M(b
n
b
n+1
), where we have used monotonicity to get b
n
b
n+1
0. But
N

n=0
M(b
n
b
n+1
) = M
N

n=0
(b
n
b
n+1
) = M(b
0
b
N+1
) Mb
0
converges, so

A
n
(b
n
b
n+1
) converges, completing the proof.
Note in the previous proof that we used a technique similar to integration by parts.
Recall from calculus that the integral
_
b
a
u(x)v(x) dx can be written as
_
b
a
u(x)v(x) dx = U(b)v(b) U(a)v(a)
_
b
a
U(x)v

(x) dx ,
where U is an antiderivative of u. Here we are thinking of A
n
as the antiderivative of a
n
and b
n
b
n+1
as the derivative of b
n
. In the sum case above, we only have one boundary
term, because one of them corresponds to A
1
b
0
= 0.
Examples
1. Alternating series test. Let (a
n
) be a monotone non-increasing sequence converging
to 0. Then

(1)
n
a
n
converges. This is obtained by applying the Dirichlet test, noting
that the partial sums of

(1)
n
are bounded by 1. As an example, we have

(1)
n
/n converges although

1/n does not .


2. For n N, let f(n) be the largest value of k such that 2
k
n (this is the integer part
of log
2
n). Then

(1)
n
/f(n) converges .
3. A series

a
n
is said to converge absolutely if

[a
n
[ converges. If it does not converge
absolutely but does converge then we say

a
n
converges conditionally. It is a famous
theorem of Riemann that given any L R and a conditionally convergent series

a
n
,
there is a rearrangement (b
n
) of the terms of (a
n
) such that

b
n
= L. See the last
section of Rudin, Chapter 3 for more details.
5.4 Exercises
1. Determine if the following series converge.
(a)

n=1
n 3
n
2
6n + 10
59
(b)

n=1
n!
1 3 (2n 1)
2. Prove that the following series converges for all x R:

n=1
sin(nx)

n
.
Hint. Use Theorem 3.42, after multiplying by sin(x/2). Use the following identity,
which is valid for all a, b R:
sin a sin b =
1
2
(cos(a b) cos(a + b)) .
Although we have not dened sin x you can use the fact that [ sin x[ and [ cos x[ are
both bounded by 1.
3. Because

1/n diverges, any series

x
n
with x
n
1/n for all n must diverge, by the
comparison test. We might think then that if

x
n
converges and x
n
0 for all n
then x
n
is smaller than 1/n, in the sense that
lim
n
nx
n
= 0 .
(a) Show that this is false; that is, there exist convergent series

x
n
with x
n
0 for
all n such that nx
n
does not converge to 0.
(b) Show however that if x
n
is monotone non-increasing and non-negative with

x
n
convergent, then nx
n
0.
Hint. Use Theorems 3.23 and 3.27.
4. Here we give a dierent proof of the alternating series test. Let x
n
be a real sequence
that is monotonically non-increasing and x
n
0.
(a) For n N, let A
n
=

2n
i=1
(1)
i
x
i
and B
n
=

2n1
i=1
(1)
i
x
i
. Show that A
n
and
B
n
converge.
(b) Prove that A
n
B
n
0 and use this to show that

(1)
n
x
n
converges.
5. Suppose that (a
n
) and (b
n
) are real sequences such that b
n
> 0 for all n and

a
n
converges. If
lim
n
a
n
b
n
= L ,= 0
then must

b
n
converge? (Here L is a nite number.)
60
6. (a) For 0 k n, recall the denition of the binomial coecient
_
n
k
_
=
n!
k!(n k)!
.
For n 0, let a
n
=
_
2n
n
_
. Find lim
n
a
n+1
an
.
(b) Find R, the radius of convergence of

n=0
_
2n
n
_
x
n
. (Fun fact: this actually equals
1

14x
when it converges; it is the Taylor series of this function.)
(c) Show that if b
1
, b
2
, . . . , b
m
are non-negative real numbers then
(1 + b
1
)(1 + b
2
) (1 + b
m
) 1 + b
1
+ + b
m
.
Write c
n
= a
n
/4
n
and
c
0
cn
=
c
n1
cn

c
n2
c
n1

c
1
c
2

c
0
c
1
. Use this inequality to show that
_
2n
n
_
/4
n
0.
(d) Show that the sum
_
2n
n
_
x
n
converges when x = R. (Fun fact we will get to
later: it converges to the value of
1

14x
at x = R, which is
1

2
.)
(e) What can you say about convergence when x = R?
61
6 Function limits and continuity
6.1 Function limits
So far we have only talked about limits for sequences. Now we step it up to functions. Pretty
quickly, though, we will see that we can relate function limits to sequence limits. The point
at which we consider the limit does not even need to be in the domain of the function f. It
is important also to notice that x is not allowed to equal x
0
below.
Denition 6.1.1. Let (X, d
X
) and (Y, d
Y
) be metric spaces and E X with x
0
a limit point
of E. If f : E Y then we write
lim
xx
0
f(x) = L
if for each > 0 there exists > 0 such that whenever x E and 0 < d
X
(x, x
0
) < , it
follows that d
Y
(f(x), L) <
Why is it important that x
0
be allowed to be only a limit point of E (and therefore not
necessarily in E)? Consider f : (0, ) R dened by f(x) =
sin x
x
. Then f is not dened
at 0 but we know from calculus that it has a limit of 1 as x 0.
Here we are using as a measure of closeness just as before. We can imagine a dialogue
similar to what occurred for sequences: you say that f(x) approaches L as x approaches x
0
.
I say Well, can you get f(x) within .005 of L as long as x is close to x
0
? You say yes and
produce a value of = .01. You then qualify this by saying, As long as x is within = .01
of x
0
then f(x) will be within .005 of L. This goes on and on, and if each time I give you
an > 0 you manage to produce a corresponding > 0, then we say the limit equals L.
As promised, there is an equivalent formulation of limits using sequences. Note that the
statement below must hold for all sequences (x
n
) in E with x
n
x
0
but x
n
,= x
0
for all n.
Proposition 6.1.2. Let f : E Y and x
0
a limit point of E. We have lim
xx
0
f(x) = L if
and only if for each sequence (x
n
) in E such that x
n
x
0
with x
n
,= x
0
for all n, it follows
that f(x
n
) L.
Proof. Suppose rst that lim
xx
0
f(x) = L and let (x
n
) be a sequence in E such that x
n
x
0
and x
n
,= x
0
for all n. We must show that f(x
n
) L. So, let > 0 and choose > 0
such that whenever d
X
(x, x
0
) < , we have d
Y
(f(x), L) < . Now since x
n
x
0
we can pick
N N such that if n N then d
X
(x
n
, x
0
) < . For this N, if n N then
d
X
(x
n
, x
0
) < , so d
Y
(f(x
n
), L) < ,
and it follows that f(x
n
) L.
Suppose conversely that f(x
n
) L for all sequences (x
n
) in E such that x
n
x
0
and
x
n
,= x
0
for all n. By way of contradiction, assume that lim
xx
0
f(x) = L does not hold.
So there must be at least one > 0 such that for any > 0 we try to nd, there is always
a x

E with d
X
(x

, x
0
) < but d
Y
(f(x

), L) > . So create a sequence of these, using


= 1/n. In other words, for each n N, pick x
n
E x
0
such that 0 < d
X
(x
n
, x
0
) < 1/n
but d
Y
(f(x
n
), L) > . (This is possible in part because x
0
is a limit point of E.) Then clearly
x
n
x
0
with x
n
,= x
0
for all n but we cannot have f(x
n
) L. This is a contradiction.
62
One nice thing about the sequence formulation is that it allows us to immediately bring
over theorems about convergence for sequences. For instance,
Proposition 6.1.3. Let E R and x
0
a limit point of E. Let f, g : R R (with the
standard metric) and a, b R. If lim
xx
0
f(x) = L and lim
xx
0
g(x) = M exist. Then
1. lim
xx
0
af(x) = aL.
2. lim
xx
0
(f(x) + g(x)) = L + M.
3. lim
xx
0
f(x)g(x) = LM.
4. If M ,= 0 then
lim
xx
0
f(x)
g(x)
=
L
M
.
6.2 Continuity
We now give the denition of continuity. Note that x
0
must be an element of E, since f
needs to be dened there.
Denition 6.2.1. If E X and f : E Y with x
0
E then we say f is continuous at x
0
if
lim
xx
0
f(x) = f(x
0
) .
We say f is continuous on E if f is continuous at every x
0
E.
Here is the equivalent denition in terms of , . The function f is continuous at x
0
if for
each > 0 there is > 0 such that if x E satises d
X
(x, x
0
) < then d
Y
(f(x), f(x
0
)) < .
Note here that we are not restricting 0 < d
X
(x, x
0
) since we trivially have d
Y
(f(x
0
), f(x
0
)) <
for all > 0. This caveat (or lack thereof) carries over the the corollary:
Corollary 6.2.2. The function f is continuous at x
0
E if and only if for each sequence
(x
n
) in E with x
n
x
0
we have f(x
n
) f(x
0
).
Proof. This is just a consequence of the sequence theorem from last section.
There is yet another equivalent denition in terms of only open sets. This one is valid for
functions continuous on all of X (although there is a more technical one for continuity at a
point, but we will not get into that). To extend the theorem to functions that are continuous
on subsets E of X, one would need to talk about sets that are open in E.
Theorem 6.2.3. If f : X Y then f is continuous on X if and only if for each open set
O Y , the preimage
f
1
(O) = x X : f(x) O
is open in X.
63
Proof. Suppose that f is continuous on X and let O Y be open. We want to show that
f
1
(O) is open. So choose x
0
f
1
(O). Since f(x
0
) O (by denition) and O is open we
can nd > 0 such that B

(f(x
0
)) O. However f is continuous at x
0
so there exists a
corresponding > 0 such that if x X with d
X
(x, x
0
) < then d
Y
(f(x), f(x
0
)) < . So if
x B

(x
0
) then f(x) B

(f(x
0
)). As B

(f(x
0
)) was chosen to be a subset of O, we nd
if x B

(x
0
) then f(x) O ,
of B

(x
0
) f
1
(O). This means x
0
is an interior point of f
1
(O) and this set is open.
Suppose that for each open O Y the set f
1
(O) is open in X. To show f is continuous
on X we must show that f is continuous at each x
0
X. So let x
0
X and > 0. The set
B

(f(x
0
)) is open in Y , so f
1
(B

(f(x
0
))) is open in X. Because x
0
is an element of this
set (note that f(x
0
) B

(f(x
0
))) it must be an interior point, so there is a > 0 such that
B

(x
0
) f
1
(B

(f(x
0
))). Now if d
X
(x, x
0
) < then d
Y
(f(x), f(x
0
)) < , so f is continuous
at x
0
.
It is dicult to get intuition about this denition, but let us give an example to illustrate
how it may work. Consider the function f : R R given by
f(x) =
_
1 if x = 0
0 if x ,= 0
.
We know from calculus that f is not continuous because it is not continuous at 0:
lim
x0
f(x) = 0 ,= 1 = f(0) .
To see this in terms of the other denition, look at the open set (1/2, 3/2). Then
f
1
((1/2, 3/2)) = 0 ,
which is not open. This only proves, however, that f is not continuous everywhere.
Corollary 6.2.4. f is continuous on X if and only if for each closed C Y , the set f
1
(C)
is closed in X.
Proof. f is continuous on X if and only if for each open O Y , the set f
1
(O) is open in
X. If C Y is closed then C
c
is open in Y . Therefore
f
1
(C) =
_
f
1
(C
c
)
_
c
is closed inX .
To check this equality, we have x f
1
(C) i f(x) C i f(x) / C
c
i x (f
1
(C
c
))
c
.
The other direction is similar. If f
1
(C) is closed in X whenever C is closed in Y , let O
be an open set in Y . Then f
1
(O
c
) is closed in X, giving
f
1
(O) =
_
f
1
(O
c
)
_
c
open in X .
64
Examples.
1. The simplest. Take f : X X as f(x) = x. Then for each open O X, f
1
(O) = O
is open in X. So f is continuous on X.
2. Let f : R R be
f(x) =
_
1 if x Q
0 if x / Q
.
This function is continuous nowhere. If x R then suppose rst x is rational. Choose
a sequence of irrationals (x
n
) converging to x (this is possible by the fact that R Q
is dense in R, from the homework). Then lim
n
f(x
n
) = 0 ,= 1 = f(x). A similar
argument holds for irrational x and gives that f is continuous nowhere.
Note that this conclusion cannot be obtained by showing that some open set O has
f
1
(O) not open. (Take for instance O = (1/2, 1/2).) This would prove that f is not
continuous everywhere.
3. The last function was discontinuous at the rationals and irrationals. This one is a
nasty function that will be discontinuous only at the rationals. For any q Q write
q (m, n) if m/n is the lowest terms representation of q; that is, if m, n are the
unique numbers with m Z, n N and m, n have no common prime factors. Then
dene f : R R by
f(x) =
_
1
n
if x Q and x (m, n) for some m Z
0 if x / Q
.
It is not hard to see that f is discontinuous at rationals. Indeed, if x is rational then
f(x) > 0 but we can choose a sequence of irrationals (x
n
) such that x
n
x, giving
0 = lim
n
f(x
n
).
On the other hand it is a bit more dicult to show that f is continuous at the irra-
tionals. Let x R Q and let > 0. Choose N N such that 1/N < . Consider the
rational numbers in the interval (x 1, x + 1). There are only nitely many rationals
in this interval that are represented as (m, 1) for some m. There are also only nitely
many in this interval represented as (m, 2) for some m. Continuing, the set of rational
numbers X
N
in this interval that are represented as (m, n) for some n N is nite and
is therefore a closed set. Since x X
c
N
we can then nd > 0 such that B

(x) X
c
N
.
We claim that if [y x[ < then [f(x) f(y)[ < . To prove this, consider rst y
irrational. Then [f(x) f(y)[ = [0 0[ = 0 < . Next if y is rational in B

(x) then
x X
c
N
and so the representation of x is (m, n) with n > N. It follows that
[f(x) f(y)[ = [0 f(y)[ = f(y) =
1
n

1
N
< .
4. The last function was discontinuous exactly at the rationals. We will see in the home-
work that there is no function that is discontinuous exactly at the irrationals.
65
We saw last time that f : R R given by f(x) = x is continuous everywhere. We will
use this along with the following proposition to show that polynomials are also continuous.
Proposition 6.2.5. Let X be a metric space and f, g : X R be continuous at x
0
X. If
a R then the following functions are continuous as x
0
:
1. f + g, af, and fg,
2. f/g as long as g(x
0
) ,= 0.
Proof. These follow using the limit properties from before; for example,
lim
xx
0
(f + g)(x) = lim
xx
0
f(x) + lim
xx
0
g(x) = f(x
0
) + g(x
0
) = (f + g)(x
0
) ,
giving that f + g is continuous at x
0
. The others follow similarly.
The proposition implies:
Any polynomial function is continuous on all of R. That is, if f(x) = a
n
x
n
+ +a
1
x+a
0
then f is continuous on R.
Every rational function is continuous at each point for which the denominator is
nonzero. That is, if f(x) = g(x)/h(x), where g and h are polynomial functions, then
f is continuous at x
0
if and only if h(x
0
) ,= 0.
Another way to build continuous functions is through composition.
Theorem 6.2.6. Let X, Y, Z be metric spaces and f : X Y , g : Y Z functions. If f is
continuous at x
0
X and g is continuous at f(x
0
) Y then g f is continuous at x
0
.
Proof. Let > 0. Since g is continuous at f(x
0
), we can choose

> 0 such that if


d
Y
(y, f(x
0
)) <

then d
Z
(g(y), g(f(x
0
))) < . Since f is continuous at x
0
, we can choose
> 0 such that if d
X
(x, x
0
) < then d
Y
(f(x), f(x
0
)) <

. Putting these together, if


d
X
(x, x
0
) < then d
Y
(f(x), f(x
0
)) <

, giving d
Z
(g(f(x)), g(f(x
0
))) < . This means
d
Z
((g f)(x), (g f)(x
0
)) < and g f is continuous at x
0
.
6.3 Relations between continuity and compactness
Continuous functions and compact sets work well together. The rst basic theorem is:
Theorem 6.3.1. If f : X Y is continuous and E X is compact then the image
f(E) = f(x) : x E is compact.
Proof. We need to show that any open cover of f(E) can be reduced to a nite subcover, so
let ( be an open cover of f(E). Dene a collection (

of sets in X by
(

= f
1
(O) : O ( .
66
Because f is continuous, each set in (

is open. Furthermore (

covers E, as every point


in x E is mapped to an element in f(E), which is covered by some O (. This means
that (

is an open cover of E, and compactness of E allows to reduce it to a nite subcover


f
1
(O
1
), . . . , f
1
(O
n
). We claim that O
1
, . . . , O
n
is a nite subcover of f(E). To show
this, let y f(E) so that there exists some x E with f(x) = y. There exists k with
1 k n such that x f
1
(O
k
) and therefore y = f(x) O
k
.
This theorem has many consequences.
Corollary 6.3.2. Let f : X Y be continuous. If E X is compact then f is bounded on
E. That is, there exists y Y and R > 0 such that d
Y
(y, f(x)) M for all x E.
Proof. From the theorem, the set f(E) is compact and therefore bounded.
The next is for continuous functions to R.
Corollary 6.3.3 (Extreme value theorem). Let f : X R be continuous and E X
compact. Then f takes a maximum on E; that is, there exists x
0
E such that
f(x
0
) f(x) for all x E .
A similar statement holds for a minimum.
Proof. The set f(E) is closed and bounded, so it contains its supremum, y. Since y f(E)
there exists x
0
E such that f(x
0
) = y. Then f(x
0
) f(x) for all x E.
Continuous functions on compact sets actually satisfy a property that is stronger than
continuity. To explain this, consider the function f : (0, ) R given by f(x) = 1/x.
When we study continuity, what we are really interested in is how much a small change in x
will change the value of f(x). (Recall continuity says that if we change x by at most then
f(x) will change by at most .) Consider the eect of changing x by a xed amount, say .1,
for dierent values of x. If x is large, like 100, then changing x by .1 can change f so that
it lies anywhere in the interval (1/100.01, 1/99.99). If x is small, like .15, then this same
change in x changes f to lie in the interval (1/.25, 1/.05) = (4, 20). This is a much larger
interval, meaning that f is more unstable to changes when x is small compared to when x
is large.
This motivates the idea of uniform continuity. For a uniformly continuous function, the
measure of stability described above is uniform on the whole set. That is, there is an upper
bound to how unstable the function is to changes. This corresponds to a uniform > 0 over
all x for a given > 0:
Denition 6.3.4. A function f : X Y is uniformly continuous if given > 0 there exists
> 0 such that if x, y X satisfy d
X
(x, y) < then d
Y
(f(x), f(y)) < .
Theorem 6.3.5. If f : X Y is continuous and X is compact then f is uniformly contin-
uous.
67
Proof. The idea of the proof is as follows. Since f is continuous at each x, given > 0 we
can nd a
x
> 0 from the denition of continuity that works at x. These
x
-balls cover
X and by compactness we can nd only nitely many
x
i
s such that these balls still cover
X. Taking the minimum of these numbers will give us the required (positive) .
Let > 0. For each x X, since f is continuous at x, we can nd
x
> 0 such that if
d
X
(x, x

) <
x
then d
Y
(f(x), f(x

)) < /2. The collection


B
x/2
(x) : x X
is an open cover for X, so since X is compact, we can nd x
1
, . . . , x
n
X such that
X
i
B
x
i
/2
(x
i
). Let
= min
x
1
/2, . . . ,
xn
/2 ;
we claim that if x, y X satisfy d
X
(x, y) < then d
Y
(f(x), f(y)) < . To prove this, pick
such x and y. We can then nd i such that d
X
(x
i
, x) <
i
/2. By the triangle inequality we
then have
d
X
(x
i
, y) d
X
(x
i
, x) + d
X
(x, y) <
i
/2 +
i
.
This means by denition of
i
that
d
Y
(f(x), f(y)) d
Y
(f(x), f(x
i
)) + d
Y
(f(y), f(x
i
)) < /2 + /2 = .
Examples.
1. Not every continuous function on a non-compact set is uniformly continuous. If E
is any non-closed subset of R then there exists a continuous function on E that is
both unbounded and not uniformly continuous. Take x
0
to be any limit point of E
that is not in E. Then f(x) = (x x
0
)
1
is continuous but unbounded. Further f
is not uniformly continuous because there is no > 0 such that for all x, y E with
[x y[ < we have [f(x) f(y)[ < 1. If there were, we could just choose some y E
with [y x
0
[ < /2 and then deduce that all points z within distance of y have
f(z) f(y) + 1. But this is impossible.
2. If E is an unbounded subset of R then there is an unbounded continuous function on
E: just take f(x) = x.
3. The only polynomials that are uniformly continuous on all of R are those of degree at
most 1. Indeed, take f(x) = a
n
x
n
+ a
1
x + a
0
with a
n
,= 0 and n 2 and assume
that there exists > 0 such that if [x y[ < then [f(x) f(y)[ < 1. Then consider
points x, y of the form x, x + /2: you can check that
[f(x) f(x + /2)[ : x R
is unbounded, giving a contradiction. (The problem here is that for xed , the quantity
[f(x) f(x + /2)[ grows to innity as x . This is not the case if n = 0 or 1.)
There are other ways that functions can fail to be uniformly continuous. We will see later,
however, that any dierentiable function with bounded derivative is uniformly continuous.
68
6.4 Connectedness and the IVT
We would like to prove the intermediate value theorem from calculus and the simplest way to
do this is to see that it is a consequence of a certain property of intervals in R. Specically,
an interval is connected. The denition of connectedness is somewhat strange so we will try
to motivate it. Instead of trying to envision what connectedness is, we will try to capture
what it is not. That is, we want to call a metric space disconnected if we can write it as a
union of two sets that do not intersect. There is a problem with this attempt at a denition,
as we can see by considering R. Certainly we can write it as (, 1/2) [1/2, ) and these
sets do not intersect, but we still want to say that R is connected. The issue in this example
is that the sets are not separated enough from each other. That is, one set contains limit
points of the other. This problem is actually resolved if we require that both sets are open.
(But you have to think about how this resolves the issue.)
Denition 6.4.1. A metric space X is disconnected if there exist non-empty open sets O
1
and O
2
in X such that X = O
1
O
2
but O
1
O
2
= . If X is not disconnected we say it is
connected.
Connectedness and continuity also go well with each other.
Theorem 6.4.2. Let X, Y be metric spaces and f : X Y be continuous. If X is connected
then the image set f(X), viewed as a metric space itself, is connected.
Proof. As stated above, we view f(X) Y as a metric space itself, using the metric it
inherits from Y . To show that f(X) is a connected space we will assume it is disconnected
and obtain a contradiction. So assume that we can write f(X) = O
1
O
2
with O
1
and O
2
nonempty, disjoint, and open (in the space f(X)). We will produce from this a disconnection
of X and obtain a contradiction.
Now consider U
1
= f
1
(O
1
) and U
2
= f
1
(O
2
). These are open sets in X since f is
continuous. Further they do not intersect: if x is in their intersection, then f(x) O
1
O
2
,
which is empty. Last, they are nonempty because, for example, if y O
1
(which is nonempty
by assumption) then because O
1
f(X), there exists x X such that f(x) = y. This x is
in f
1
(O
1
).
So we nd that X is disconnected, a contradiction. This means f(X) must have been
connected.
Let X be a discrete metric space. If X consists of at least two points then X is
disconnected. This is because we can let O
1
= x for some x X and O
2
= O
c
1
. All
subsets of X are open, so these are open, disjoint, nonempty sets whose union is X.
Every interval in R is connected. You will prove this in exercise 6.
Theorem 6.4.3 (Intermediate value theorem). Let f : [a, b] R for a < b be continuous.
Suppose that for some L R,
f(a) < L < f(b) .
Then there exists c (a, b) such that f(c) = L.
69
Proof. Since f is continuous, the space f([a, b]) is connected. Since
O
1
:= (, L) f([a, b]) and O
2
:= (L, ) f([a, b])
are both nonempty (because f(a) O
1
and f(b) O
2
), open in f([a, b]), and disjoint, it
cannot be that their union is equal to f([a, b]). Therefore L f([a, b]) and there exists
x X with f(x) = L.
6.5 Discontinuities
Let us spend a couple of minutes on types of discontinuities for real functions. Let E R,
f : E R and x
0
E. (Draw some pictures.)
x
0
is a removable discontinuity of f if lim
xx
0
f(x) exists but it not equal to f(x
0
).
x
0
is a simple discontinuity of f if lim
xx

0
f(x) exists, as does lim
xx
+
0
f(x), but they
are not equal. Here the rst limit is a left limit; that is, we are considering f as being
dened on the metric space E(, x
0
] and taking the limit in this space. The second
is a right limit, and we consider the space as E [x
0
, ). This corresponds to saying,
for example, that
lim
xx

0
f(x) = L
if for each > 0 there exists > 0 such that if x
0
< x < x
0
then [f(x) L[ < .
x
0
can be a discontinuity but not captured above. Consider f : R R given by
f(x) =
_
sin(1/x) if x ,= 0
0 if x = 0
.
Here there is not even a limit as x 0. This is because we can nd a sequences (x
n
)
converging to 0 such that (f(x
n
)) does not have a limit. Take
x
n
= 2/(n) .
6.6 Exercises
1. Let f : [a, b] R be continuous with f(x) > 0 for all x [a, b]. Show there exists
> 0 such that f(x) for all x [a, b].
2. Determine if the following functions are continuous at x = 0. Prove your answer. (You
may use standard facts about trigonometric functions although we have not introduced
them rigorously.)
(a)
f(x) =
_
x cos
1
x
if x ,= 0
0 if x = 0
.
70
(b)
g(x) =
_
sin
1
x
if x ,= 0
0 if x = 0
.
3. (a) Let f, g : R R be continuous. Show that h : R R is continuous, where h is
given by
h(x) = maxf(x), g(x) .
(b) Let ( be a set of continuous functions from R to R . For each x, assume that
f(x) : f ( is bounded above and dene F : R R by
F(x) = supf(x) : f ( .
Must F be continuous?
4. In this problem we will show that there is no real-valued function that is continuous
exactly at the rationals. Fix any f : R R.
(a) Show that for each n N, the set A
n
is open, where
A
n
=
_
x : > 0 such that [f(z) f(y)[ <
1
n
for all y, z (x , x + )
_
.
(b) Prove that the set of points at which f is continuous is equal to
nN
A
n
.
(c) Prove that
nN
A
n
cannot equal Q.
Hint. Argue by contradiction and enumerate the rationals as q
1
, q
2
, . . .. Dene
B
n
= A
n
q
n
and obtain a contradiction using exercise 5 of Chapter 3.
5. Find metric spaces X, Y , a continuous function f : X Y , and a Cauchy sequence
x
n
in X such that f(x
n
) is not Cauchy in Y .
6. Read the last section of Chapter 4 in Rudin on limits at innity. Prove that the
function f : (0, ) R given by f(x) = 1/x has lim
x
f(x) = 0.
7. Prove that any interval I R is connected.
Hint. Consider I as a metric space with the standard metric from R. Suppose that
I = O
1
O
2
where O
1
O
2
= and both O
i
s are nonempty and open in I. Then
there must be a point x
1
O
1
and a point x
2
O
2
. Suppose that x
1
< x
2
and dene
I
1
= r x
1
: [x
1
, r] O
1
.
What can you say about sup I
1
?
8. Let f : (a, b) R be uniformly continuous. Prove that f has a unique continuous
extension to [a, b). That is, there is a unique g : [a, b) R which is continuous and
agrees with f everywhere on (a, b). Show by example that it is not enough to assume
f is only continuous, or even both continuous and bounded.
71
9. Show that the function f given by f(x) = 1/x is uniformly continuous on [1, ).
10. Show that the function f given by f(x) =

x is uniformly continuous on [0, ).


Hint. Use the fact that for a, b 0 we have

a +

a + b.
11. Show that the function f given by f(x) = sin(1/x) is not uniformly continuous on
(0, 1).
12. Suppose that f : [0, ) R is continuous and has a nite limit lim
x
f(x). Show
that f is uniformly continuous.
13. Give an example of functions f, g : [0, ) R that are uniformly continuous but the
product fg is not.
14. Let f : R R be continuous with f(f(x)) = x for all x. Show there exists c R such
that f(c) = c.
15. Let p be a polynomial with real coecients and odd degree. That is,
p(x) = a
n
x
n
+ + a
1
x + a
0
with a
n
,= 0 and n odd .
(a) Show there exists c such that p(c) = 0.
(b) Let L R. Show there exists c such that p(c) = L.
16. If E R then a function f : E R is called Lipschitz if there exists M > 0 such that
[f(x) f(y)[ M[x y[ for all x, y E .
The smallest number such that the above inequality holds for all x, y E is called the
Lipschitz constant for f.
(a) Show that if f : E R is Lipschitz then it is uniformly continuous. Does the
converse hold?
(b) Show that the function f : R R given by f(x) =

x
2
+ 4 is Lipschitz on R.
What is the Lipschitz constant?
(c) Is f : [0, ) R given by f(x) =

x Lipschitz?
17. Let I be a closed interval. Let f : I I and assume that f is Lipschitz with Lipschitz
constant A < 1.
(a) Prove that there is a unique y I with the following property. Choose x
1
I
and dene x
n+1
= f(x
n
) for all n N. Then x
n
y. This holds independently
of the choice of x
1
.
(b) Show by counterexample that for (a) to work, we need I to be closed.
72
(c) Choose a
1
, a
2
, . . . , a
k
Q with a
i
> 0 for all i and with a
1
a
k
> 1. Starting
from any x
1
> 0, dene a sequence x
n
by the continued fraction
x
n
=
1
a
1
+
1
a
2
+
1
+
1
a
k
+ x
n1
.
Prove that x
n
converges. Prove that its limit is the root of a quadratic poly-
nomial with coecients in Q. In older books this is stated: an innite periodic
continued fraction is a quadratic surd. The devil is the eternal surd in the
universal mathematic. C. S. Lewis, Perelandra.
18. Let f : I R for some (open or closed) interval I R. We say that f is convex if for
all x, y I and [0, 1],
f(x + (1 )y) f(x) + (1 )f(y) .
(a) Reformulate the above condition in terms of a relation between the graph of f
and certain line segments.
(b) Suppose that f : R R is convex and let x < z < y. Choose =
yz
yx
to show
that
f(z) f(x)
z x

f(y) f(x)
y x
.
Interpret this inequality in terms of the graph of f. Argue similarly to show that
f(y) f(x)
y x

f(y) f(z)
y z
.
Combine these two to get
f(z) f(x)
z x

f(y) f(z)
y z
and interpret this inequality in terms of the graph of f.
(c) Suppose that f : [a, b] R is convex. Show that f is continuous on (a, b).
Hint. Let [c, d] be a subinterval of (a, b). Use the last inequality from (b) to show
that f is Lipschitz on [c, d] with Lipschitz constant bounded above by
max
_
[f(c) f(a)[
[c a[
,
[f(b) f(d)[
[b d[
_
.
19. Suppose that f : R R is continuous and satises
f(x + y) = f(x) + f(y) for all x, y R .
(a) Show that there exists c R such that for all x Z, f(x) = cx.
(b) Show that there exists c R such that for all x Q, f(x) = cx.
(c) Show that there exists c R such that for all x R, f(x) = cx.
73
7 Derivatives
7.1 Introduction
Continuous functions are nicer than most functions. However we have seen that they can still
be rather weird (recall the function that equals 1/q at a rational expressed in lowest terms
as p/q). So we move on to study functions that are even nicer, and for this we henceforth
restrict to functions from R to R. We could start at the very bottom, rst studying constant
functions f(x) = c and then linear functions f(x) = ax+b, then quadratics, etc. But I trust
you learned about these functions earlier. Noting that constant functions are just special
cases of linear ones, we set out to study functions that are somehow close to linear functions.
The idea we will pursue is that even if a function f is wild, it may be that very close to
a particular point x
0
, it may be well represented by a linear function. For a good choice of
a linear function L, it would make sense to hope that
lim
xx
0
(f(x) L(x)) = 0 .
If f is already continuous then this is not much of a requirement: we just need L(x
0
) = f(x
0
).
So this just means that L(x) can be written as L(x) = a(x x
0
) + f(x
0
).
We will look for a stronger requirement on the speed at which this dierence converges
to zero. It should go to zero at least as fast as x x
0
does (as x x
0
). In other words, we
will require that
lim
xx
0
f(x) L(x)
x x
0
= 0 or in shorthand, f(x) L(x) = o(x x
0
) .
Plugging in our form of L, this means
lim
xx
0
_
f(x) f(x
0
)
x x
0
a
_
= 0 ,
or
lim
xx
0
f(x) f(x
0
)
x x
0
= a .
Rewriting this with the notation above, we get
f(x) = f(x
0
) + a(x x
0
) + o(x x
0
) ,
or setting x = x
0
+ h,
f(x
0
+ h) = f(x
0
) + ah + o(h)
as h 0. Again, the symbol o(h) represents some term such that if we divide it by h and
take h 0, it goes to 0.
Denition 7.1.1. Let f : (a, b) R. We say that f is dierentiable at x
0
(a, b) if
lim
xx
0
f(x) f(x
0
)
x x
0
exists .
In this case we write f

(x
0
) for the limit.
74
7.2 Properties
Proposition 7.2.1. Let f : (a, b) R be dierentiable at x
0
. Then f is continuous at x
0
.
Proof.
lim
xx
0
f(x) = f(x
0
) + lim
xx
0
[f(x) f(x
0
)] = f(x
0
) + lim
xx
0
f(x) f(x
0
)
x x
0
lim
xx
0
(x x
0
) = f(x
0
) .
The converse is not true. Consider the function f : R R given by f(x) = [x[. Then
f(x) = maxx, 0 minx, 0 ,
so it is continuous. However, trying to compute the derivative at x = 0, we get
lim
x0
[x[
x
,
which does not exist (it has a right limit of 1 and left limit of -1).
We will now play the same game as we did for continuity, trying to nd which functions
are dierentiable. Here are some examples.
1. f(x) = x:
lim
xx
0
f(x) f(x
0
)
x x
0
= 1 .
So f

(x) exists for all x and equals 1.


2. f(x) = x
n
for n N:
lim
xx
0
x
n
x
n
0
x x
0
= lim
xx
0
(x x
0
)(x
n1
+ x
n2
x
0
+ + xx
n2
0
+ x
n1
0
)
x x
0
= lim
xx
0
_
x
n1
+ + x
n1
0

= nx
n1
0
.
So f

(x) exists for all x and equals nx


n1
.
Again we look at how to build dierentiable functions from others.
Proposition 7.2.2. Let f, g : (a, b) R be dierentiable at x. Then the following functions
are dierentiable with derivatives:
1. (f + g)

(x) = f

(x) + g

(x)
2. (fg)

(x) = f

(x)g(x) + f(x)g

(x).
3. (f/g)

(x) =
f

(x)g(x)f(x)g

(x)
g
2
(x)
if g(x) ,= 0.
75
Proof. For the rst we just use properties of limits:
lim
yx
(f + g)(y) (f + g)(x)
y x
= lim
yx
f(y) f(x)
y x
+ lim
yx
g(y) g(x)
y x
= f

(x) + g

(x) .
For the second, we write
(fg)(y) (fg)(x) = (f(y) f(x))g(y) + f(x)(g(y) g(x)) ,
divide by y x and take a limit:
lim
yx
(fg)(y) (fg)(x)
y x
= lim
yx
f(y) f(x)
y x
lim
yx
g(y) + f(x) lim
yx
g(y) g(x)
y x
.
As g is dierentiable at x, it is also continuous, so g(y) g(x) as x y. This gives the
formula.
The last property can be derived in a similar fashion:
(f/g)(y) (f/g)(x) =
1
g(y)g(x)
[f(y)g(x) f(x)g(y)]
=
1
g(y)g(x)
[g(x)(f(y) f(x)) f(x)(g(y) g(x))] .
Dividing by y x and taking the limit gives the result.
Again, from this proposition, we nd that all polynomials are dierentiable everywhere
as are rational functions wherever the denominator is nonzero. The next way to build
dierentiable functions is to compose:
Theorem 7.2.3 (Chain rule). Let f : (a, b) (c, d) be dierentiable at x
0
and g : (c, d) R
be dierentiable at f(x
0
). Then g f is dierentiable at x
0
with derivative
(g f)

(x
0
) = f

(x
0
)g

(f(x
0
)) .
Proof. We will want to use a division by f(y) f(x
0
) for y ,= x
0
, so we must rst deal with
the case that this could be 0. If there exists a sequence (x
n
) in (a, b) with x
n
x
0
but
x
n
,= x
0
for all n with f(x
n
) = f(x
0
) for innitely many n, we would have
f

(x
0
) = lim
yx
f(y) f(x
0
)
y x
0
= lim
n
f(x
n
) f(x
0
)
x
n
x
0
= 0 ,
so the right side of the equation in the theorem would be 0. The left side would also be zero
for a similar reason:
lim
yx
0
(g f)(y) (g f)(x
0
)
y x
0
= lim
n
g(f(x
n
)) g(f(x
0
))
x
n
x
0
= 0 .
76
In the other case, every sequence (x
n
) in (a, b) with x
n
x
0
and x
n
,= x
0
has f(x
n
) =
f(x
0
) for at most nitely many n. Then as f is continuous at x
0
, we have f(x
n
) f(x
0
)
with a
n
,= f(x
0
) for all n and so
lim
yx
(g f)(y) (g f)(x)
y x
= lim
n
g(f(x
n
)) g(f(x
0
))
x
n
x
0
= lim
n
g(f(x
n
)) g(f(x
0
))
f(x
n
) f(x
0
)
lim
n
f(x
n
) f(x
0
)
x
n
x
0
= g

(f(x
0
))f

(x
0
) .
Examples.
1. We know f(x) = [x[ is continuous but not dierentiable. To go one level deeper,
consider
f(x) =
_
x
2
x 0
x
2
x < 0
.
The derivative at 0 is
lim
h0
f(0 + h)
h
= 0 ,
and the derivative elsewhere is
f

(x) =
_
2x x > 0
2x x < 0
.
Note that f

is continuous. Then we say f C


1
(or f is in class C
1
). However the
second derivative does not exist.
2. The function
f(x) =
_
x
3
x 0
x
3
x < 0
is in class C
2
, as it has two continuous derivatives. But it is not three times dieren-
tiable.
3. Generally, the function
f(x) =
_
x
n
x 0
x
n
x < 0
, n 1
is in class C
n1
, meaning that it has n1 continuous derivatives. But it is not n times
dierentiable.
77
7.3 Mean value theorem
We begin by looking at local extrema.
Denition 7.3.1. For X a metric space, let f : X R. We say that x
0
X is a local
maximum for f if there exists r > 0 such that for all x B
r
(x
0
) we have f(x) f(x
0
).
Similarly x
0
is a local minimum for f if there exists r > 0 such that for all x B
r
(x
0
) we
have f(x) f(x
0
).
In the case that X is R, if f is dierentiable at a local extreme point, then the derivative
must be zero.
Proposition 7.3.2. Let f : (a, b) R and suppose that c (a, b) is a local extreme point
for f. If f

(c) exists then f

(c) = 0.
Proof. Let c be a local max such that f

(c) exists. Then there exists r > 0 such that for all
y with [y c[ < r, we have f(y) f(c). Therefore, looking at only right limits,
lim
yc
+
f(y) f(c)
y c
0 .
Looking only at left limits,
lim
yc

f(y) f(c)
y c
0 .
Putting these together, we nd f

(c) = 0. The argument for local min is similar.


Theorem 7.3.3 (Rolles theorem). For a < b, let f : [a, b] R be continuous such that f
is dierentiable on (a, b). If f(a) = f(b) then there exists c (a, b) such that f

(c) = 0.
Proof. If f is constant on the interval then clearly the statement holds. Otherwise for some
d (a, b) we have f(d) > f(a) or f(d) < f(a). Let us consider the rst case; the second is
similar. By the extreme value theorem, f takes a maximum on [a, b] and since f(d) > f(a)
this max cannot occur at a or b. So it occurs at some c (a, b). Then c is a local max as
well, so we can apply the previous proposition to nd f

(c) = 0.
An important corollary is the following.
Corollary 7.3.4 (Mean value theorem). For a < b let f : [a, b] R be continuous such that
f is dierentiable on (a, b). There exists c (a, b) such that
f

(c) =
f(b) f(a)
b a
.
Proof. Dene L(x) to be the line that connects the points (a, f(a)) and (b, f(b)):
L(x) =
f(b) f(a)
b a
(x a) + f(a) .
78
Then the function g = f L satises g(a) = g(b) = 0. It is also continuous on [a, b]
and dierentiable on (a, b). Therefore by Rolles theorem, we can nd c (a, b) such that
g

(c) = 0. This gives


0 = g

(c) = f

(c) L

(c) = f

(c)
f(b) f(a)
b a
,
implying the corollary.
The mean value theorem has a lot of consequences. It is one of the central tools to
analyze derivatives.
Corollary 7.3.5. Let f : (a, b) R be dierentiable.
1. If f

(x) 0 for all x (a, b) then f is non-decreasing.


2. If f

(x) 0 for all x (a, b) then f is non-increasing.


3. If f

(x) = 0 for all x (a, b) then f is constant.


Proof. Suppose rst that f

(x) 0 for all x (a, b). To show f is non-decreasing, let c < d


in (a, b). By the mean value theorem, there exists x
0
(c, d) such that
f

(x
0
) =
f(d) f(c)
d c
.
But this quantity is nonnegative, giving f(d) f(c). The second follows by considering f
instead of f. The third follows from the previous two.
7.4 LHopitals rule
For the proof of LHopitals rule, we need a generalized version of the mean value theorem.
Lemma 7.4.1 (Generalized MVT). If f, g : [a, b] R are continuous and dierentiable on
(a, b) then there exists c (a, b) such that
(f(b) f(a))g

(c) = (g(b) g(a))f

(c) .
Proof. The proof is exactly the same as that of the MVT but using the function h : [a, b] R
given by
h(x) = (f(b) f(a))g(x) (g(b) g(a))f(x) .
Indeed, h(a) = (f(b) f(a))g(a) (g(b) g(a))f(a) = h(b), so applying Rolles theorem, we
nd c (a, b) such that h

(c) = 0.
79
Theorem 7.4.2 (LHopitals rule). Suppose f, g : (a, b) R are dierentiable with g

(x) ,= 0
for all x, where a < b < . Suppose that
f

(x)
g

(x)
A as x a .
If f(x) 0 and g(x) 0 as x a or if g(x) + as x a, then
f(x)
g(x)
A as x a .
Proof. We will suppose that A, a ,= ; otherwise the argument is similar. We consider
two cases. First suppose that f(x) 0 and g(x) 0 as x a. Then let > 0 and choose
> 0 such that if x (a, a + ) then

(x)
g

(x)
A

< /2 .
We will now show that if x (a, a + ) then also

f(x)
g(x)
A

< . Indeed, choose such an x


and then pick any y (a, x). From the generalized MVT, there exists c (y, x) such that
f(x) f(y)
g(x) g(y)
=
f

(c)
g

(c)
.
Note that the denominator is nonzero since g is injective (just use the MVT). But since
c (a, a + ), we have

f(x) f(y)
g(x) g(y)
A

< /2 .
Let y a and we nd the result.
In the second case, we suppose that g(x) + as x a. Again for > 0 pick
1
> 0
such that if x (a, a +
1
) then

(x)
g

(x)
A

< /2 .
Fix x
0
= a +
1
. By the generalized MVT, as before, for all x (a, x
0
),
A /2 <
f(x) f(x
0
)
g(x) g(x
0
)
< A + /2 . (4)
Notice that since g(x) as x a,
lim
xa
x(a,x
0
)
g(x) g(x
0
)
g(x)
= 1 .
80
Therefore using equation (4), there exists
2
<
1
such that if x (a, a +
2
) then
A 3/4 <
f(x) f(x
0
)
g(x) g(x
0
)

g(x) g(x
0
)
g(x)
< A + 3/4 . (5)
Also since g(x) as x a,
lim
xa
x(a,x
0
)
f(x
0
)
g(x)
= 0 .
Therefore using (9) we can nd
3
<
2
such that if x (a, a +
3
) then
A <
f(x) f(x
0
)
g(x) g(x
0
)

g(x) g(x
0
)
g(x)
+
f(x
0
)
g(x)
< A + .
But this means
A <
f(x)
g(x)
< A + for all x (a, a +
3
) .
This proves that
f(x)
g(x)
A as x a.
7.5 Power series
We will derive some results about power series because they will help us on the problem set
to dene trigonometric functions. Let f(x) =

n=0
a
n
x
n
be a power series with radius of
convergence R > 0. We wish to show that
f is dierentiable on (R, R).
The power series

n=0
na
n
x
n1
also has radius of convergence R.
For all x (R, R), f

(x) =

n=0
na
n
x
n1
.
Step 1. The power series

n=0
na
n
x
n1
also has radius of convergence R. To show this, we
need a lemma.
Lemma 7.5.1. Suppose that (x
n
) and (y
n
) are non-negative real sequences such that x
n

x > 0. Then
limsup
n
x
n
y
n
= x limsup
n
y
n
.
Proof. We will use the denition from the homework that limsup
n
b
n
is the supremum
of all subsequential limits of (b
n
). Let S be the set of subsequential limits of (y
n
) and T the
corresponding set for (x
n
y
n
). We will prove the case that S and T are bounded above; the
other case is left as an exercise.
We claim that
xS = T , where xS = xs : s S .
To prove this, let a xS. Then there exists a subsequence (y
n
k
) such that y
n
k
a/x.
Now x
n
k
y
n
k
xa/x = a, giving that a T. Conversely, let b T so that there exists a
81
subsequence (x
n
k
y
n
k
) such that x
n
k
y
n
k
b. Then y
n
k
= x
n
k
y
n
k
/x
n
k
b/x. This means
that b = xb/x xS.
To nish the proof we show that sup T = x sup S. First if t T we have t/x S, so
t/x sup S. Therefore t x sup S and sup T x sup S. Conversely if s S then xs T,
so xs sup T, giving s (1/x) sup T. This means sup S (1/x) sup T and therefore
sup T x sup S.
To nd the radius of convergence of

n=0
na
n
x
n1
, we use the root test:
limsup
n
(n[a
n
[)
1/n
= limsup
n
n
1/n
[a
n
[
1/n
.
Since n
1/n
1 we can use the previous lemma to get a limsup of 1/R, where R is the radius
of convergence of

n=0
a
n
x
n
. This means the radius of convergence of the new series is also
R.
Step 2. The function f given by f(x) =

n=0
a
n
x
n
is dierentiable at x = 0.
To prove this, we use 0 < [x[ < R/2 and compute
f(x) f(0)
x 0
=

n=0
a
n
x
n
a
0
x
=

n=1
a
n
x
n1
.
Pulling o the rst term,

f(x) f(0)
x 0
a
1

n=2
a
n
x
n1

= [x[

n=2
a
n
x
n2

.
We can use the triangle inequality for the last sum to get

f(x) f(0)
x 0
a
1

[x[

n=2
[a
n
[[x[
n2
[x[

n=2
[a
n
[(R/2)
n2
.
By the ratio test, the last series converges, so setting C equal to it, we nd

f(x) f(0)
x 0
a
1

C[x[ .
Now we can take the limit as x 0 and nd
lim
x0

f(x) f(0)
x 0
a
1

= 0 , or lim
x0
f(x) f(0)
x 0
= a
1
.
This means f

(0) = a
1
.
Step 3. We will now prove that f is dierentiable at all [x[ < R. So take such an x
0
and use
the binomial theorem:
f(x) =

n=0
a
n
(x x
0
+ x
0
)
n
=

n=0
a
n
_
n

j=0
_
n
j
_
x
nj
0
(x x
0
)
j
_
=

n=0

j=0
1
nj
a
n
_
n
j
_
x
nj
0
(x x
0
)
j
. (6)
We now state a lemma.
82
Lemma 7.5.2. Let a
m,n
, m, n 0 be a double sequence. If

n=0
[

m=0
[a
m,n
[] converges
then

n=0
_

m=0
a
m,n
_
=

m=0
_

n=0
a
m,n
_
.
Proof. Let > 0 and write S for the left side above and T for the right side above. For
M, N N, dene
S
M,N
=
N

n=0
M

m=0
a
m,n
and T
M,N
=
M

m=0
N

n=0
a
m,n
.
Clearly S
M,N
= T
M,N
for all M, N N. We claim that there exists M
0
, N
0
such that if
M M
0
and N N
0
then both [S S
M,N
[ and [T T
M,N
[ are less than /2. We need to
only verify this for S because the same argument works for T. Once we show that, we have
[S T[ [S S
M,N
[ +[S
M,N
T
M,N
[ [T
M,N
T[ < ,
and since is arbitrary, this means S = T.
To prove the claim rst use the fact that

n=0
[

m=0
[a
m,n
[] converges to pick N
0
such
that if n N
0
then

n=N
0
+1
[

m=0
[a
m,n
[] < /4. Next because each sum

m=0
[a
m,0
[,

m=0
[a
m,1
[, . . . ,

m=0
[a
m,N
0
[
converges, we can pick M
0
such that if

N
0
n=0

m=0
[a
m,n
[ < /4. This gives for M M
0
and N N
0
,
[S S
M,N
[
N

n=0
_

m=M+1
[a
m,n
[
_
+

n=N+1
_

m=0
[a
m,n
[
_
=
N
0

n=0
_

m=M+1
[a
m,n
[
_
+
N

n=N
0
+1
_

m=M+1
[a
m,n
[
_
+

n=N+1
_

m=0
[a
m,n
[
_

N
0

n=0
_

m=M+1
[a
m,n
[
_
+

n=N
0
+1
_

m=0
[a
m,n
[
_

N
0

n=0
_

m=M
0
+1
[a
m,n
[
_
+

n=N
0
+1
_

m=0
[a
m,n
[
_
< /2 .
We now want to apply the lemma to the sum in (6). To do this, we must verify that

n=0
_

j=0
1
nj
[a
n
[
_
n
j
_
[x
0
[
nj
[x x
0
[
j
_
83
converges. But using the binomial theorem again, this sum equals

n=0
[a
n
[([x
0
[ +[x x
0
[)
n
,
which converges as long as [x
0
[ +[x x
0
[ < R. So pick such an x and we can exchange the
order of summation:
f(x) =

j=0
_

n=j
a
n
_
n
j
_
x
nj
0
_
(x x
0
)
j
.
We can view this as a power series in x x
0
by setting g(x) = f(x +x
0
) and seeing that for
[x[ < R [x
0
[,
g(x) =

j=0
b
j
x
j
, with b
j
=

n=j
a
n
_
n
j
_
x
nj
0
.
Taking the derivative of this at x = 0 gives by the previous computation
f

(x
0
) = g

(0) = b
1
=

n=1
na
n
x
n1
0
.
7.6 Taylors theorem
Note that the theorem on power series actually gives that if f(x) =

n=0
a
n
x
n
then f has
innitely many derivatives. (Just apply the theorem over and over.) Then we ask: is it true
that if a function has innitely many derivatives then it is equal to some power series?
Denition 7.6.1. A function f : (a, b) R is called analytic if it equals some power series

n=0
a
n
x
n
.
The question now becomes: is every f C

actually analytic? To try to answer this


question we look at the derivatives of a power series: if f(x) =

n=0
a
n
x
n
, then
f(0) = a
0
, f

(0) = a
1
, f

(0) = 2a
2
, f

(0) = 6a
3
, . . .
So we can rewrite a power series as
f(x) =

n=0
f
(n)
(0)
n!
x
n
.
The sum on the right is called the Taylor series for f.
To try to go the other way (to try to build a power series from a function), suppose for
simplicity that f : R R and a < b. If f is dierentiable on (a, b) and continuous on [a, b],
the mean value theorem gives c
1
(a, b) such that
f(b) = f(a) + f

(c
1
)(b a) .
84
We can then ask, if f is twice dierentiable, can we nd c
2
(a, b) such that
f(b) = f(a) + f

(a)(b a) +
f

(c
2
)
2
(b a)
2
,
or a c
3
(a, b) such that
f(b) = f(a) + f

(a)(b a) +
f

(a)
2
(b a)
2
+
f

(c
3
)
6
(b a)
3
?
The answer is yes, and in fact we can keep going to any order we like. For its statement,
derivatives at a and b are understood as right and left derivatives, respectively.
Theorem 7.6.2 (Taylors theorem). Suppose that f : [a, b] R has n 1 continuous
derivatives on [a, b] and is n times dierentiable on (a, b). There exists c (a, b) such that
f(b) =
n1

j=0
f
(j)
(a)
j!
(b a)
j
+
f
(n)
(c)
n!
(b a)
n
.
Proof. See the proof in Rudin, Thm. 5.15. It is a repeated application of the mean value
theorem.
We get from this a corollary:
Corollary 7.6.3. Suppose that f : [a, b] R has innitely many derivatives; that is, f
C

([a, b]). Set


M
n
= sup
c(a,b)
f
(n)
(c) .
If
Mn
n!
(b a)
n
0 then
f(b) =

n=0
f
(n)
n!
(b a)
n
.
We can see that in this corollary it is necessary to have this bound on M
n
. Take for
example f : [0, ) R given by
f(x) =
_
e
1/x
if x > 0
0 if x = 0
.
In this case, you can check that f
(n)
(0) = 0 for all n. However, if f(x) =

n=0
a
n
x
n
this
would imply that a
n
= 0 for all n, giving f(x) = 0 for all x.
This means in particular that we must not have the required growth on f
(n)
(x) to apply
the corollary. If you compute the n-th derivative, you can try to see why the corollary does
not apply; that is, why f is not analytic. For instance, we have
f

(x) = e
1/x
_
1
x
2
_
, f

(x) = e
1/x
_
1
x
4

2
x
3
_
for x > 0
85
and n-th derivative can be written as
f
(n)
(x) = e
1/x
P(1/x) ,
where P is a polynomial in 1/x of degree 2n. For any given r > 0, you can show that
sup
x[0,r]
f
(n)
(x)
n!
r
n
,
so that the corollary cannot apply.
7.7 Exercises
1. Prove that for any c R, the polynomial equation x
3
3x + c = 0 does not have two
distinct roots in [0, 1].
2. Suppose that f : R R is dierentiable and there exists C < 1 such that [f

(x)[ C
for all x.
(a) Show that there exists a unique xed point; that is, an x such that f(x) = x.
(b) Show that if f(0) > 0 then the xed point is positive.
3. Let f : R R be continuous. Suppose that for some a < b, both of the following two
conditions hold:
f(a) = f(b) = 0 and
f si dierentiable at both a and b with f

(a)f

(b) > 0.
Show there exists c (a, b) such that f(c) = 0.
4. Assume f on [a, b] is continuous, and that f

exists and is everywhere continuous and


positive on (a, b). Let [c, d] be the image of f. Prove that f has an inverse function
f
1
: [c, d] [a, b] and that the derivative of f
1
is continuous on (c, d).
5. Let f : (a, a) R. Assume there is a C R such that for all x (a, a), we have
[f(x) x[ Cx
2
. Does f

(0) exist? If so, what is it?


6. Use the Mean Value Theorem to prove that for x ,= 0
1 +
x
2

1 + x
<

1 + x < 1 + x/2 .
7. If I is an open interval and f : I R is dierentiable, show that [f

(x)[ is bounded on
I by a constant M if and only if f is Lipschitz on I with Lipschitz constant bounded
above by (this same) M.
86
8. Read example 5.6 in Rudin. Dene f : R R by
f(x) =
_
x
200
sin
1
x
x ,= 0
0 x = 0
.
(a) For which n N does f
(n)
(0), the n-th derivative of f at 0, exist?
(b) For which n N does lim
x0
+ f
(n)
(x) exist?
(c) For which n N is f C
n
(R)?
9. Let I R be an open interval. Assume f : I R is continuous on I and is dieren-
tiable on I except perhaps at c I. Suppose further that lim
xc
f

(x) exists. Prove


that f is dierentiable at c and that f

is continuous at c.
10. (Weierstrass M-test) Let I be any interval. For each n N, let f
n
: I R be
continuous and assume that there is a constant M
n
such that [f
n
(x)[ M
n
for all x.
Assume further that

M
n
converges.
(a) Show that for each x I, the sum

n
f
n
(x) converges. Call this number f(x).
We say f
n
converges pointwise to f.
(b) Show that f : I R given in the rst part is continuous.
Hint. Given > 0, nd N N such that

N
n=1
f
n
(x) f(x)

< /2 for all x I.


Then use the fact that

N
n=1
f
n
is continuous.
Remark. The condition above with /2 is called uniform convergence. Precisely,
we say a family f
n
of functions from R to R converges uniformly to f if for
each > 0 there exists N such that n N implies that [f
n
(x) f(x)[ < for all
x. This problem is a special case of a more general theorem: if a family f
n
of
continuous functions converges uniformly then the limit f is continuous. Try to
think up an example where a family of continuous functions converges pointwise
to f, but does not converge uniformly, and where f is not continuous.
11. (From J. Feldman.) In this problem we will construct a function that is continuous
everywhere but dierentiable nowhere. Dene g : R R by rst setting for x [0, 2],
g(x) =
_
x x [0, 1]
2 x x [1, 2]
.
Then for x / [0, 2], dene g(x) so that it is periodic of period 2; that is, set g(x) = g( x)
for the unique x [0, 2) such that x = x +2m for some m Z. (The graph of g forms
a sequence of identical triangles with the x-axis, each of height 1 and base 2. Clearly
g is continuous.) For each n N, dene f
n
: [0, 1] R by f
n
(x) =
_
3
4
_
n
g(4
n
x).
(a) Make a sketch of f
1
and f
2
on [0, 1]. (Optional: use a computer algebra package
to graph f
1
, f
1
+ f
2
, f
1
+ f
2
+ f
3
, etc.)
87
(b) Prove that the formula f(x) =

n=1
f
n
(x) denes a continuous function on [0, 1].
(c) Complete the following steps to show that f is not dierentiable at any x.
i. Let x [0, 1] and for each m N, dene h
m
to be either number in the set
_
x
1
2
4
m
, x +
1
2
4
m
_
such that there is no integer strictly between 4
m
x and
4
m
h
m
. Show that
if n > m then
f
n
(h
m
) f
n
(x)
h
m
x
= 0 .
ii. Show that
if n = m then

f
n
(h
m
) f
n
(x)
h
m
x

= 3
m
.
iii. Show that
if n < m then

f
n
(h
m
) f
n
(x)
h
m
x

3
n
.
Putting these three cases together, show that

f(h
m
) f(x)
h
m
x

1
2
(3
m
+ 3)
and deduce that f is not dierentiable at x.
12. Dene for x R,
sin x =

n=0
(1)
n
x
2n+1
(2n + 1)!
and cos x =

n=0
(1)
n
x
2n
(2n)!
.
(a) Show that for any x, both series converge absolutely and dene continuous func-
tions. Show that cos 0 = 1 and sin 0 = 0.
(b) Show that the derivative of sin x is cos x and the derivative of cos x is sin x.
(c) Show that for any x, sin
2
x + cos
2
x = 1.
Hint. Take the derivative of the left side.
(d) For a given a R nd the Taylor series of both f(x) = sin(a + x) and g(x) =
cos(a + x) centered at x = 0.
(e) Use the previous part to show the identities
sin(x + y) = sin x cos y + cos x sin y and cos(x + y) = cos x cos y sin x sin y .
13. Dene the set
S = x > 0 : cos x = 0 .
88
(a) Show that S is nonempty.
Hint. Assume it is empty. Since cos 0 = 1, show that then cos x would be
positive for all x > 0 and therefore sin x would be strictly increasing. As sin x is
bounded, it would have a limit as x . Deduce then that cos x would also
have a limit L. Show that L = 2L
2
1 and that we must have L = 1. Argue that
this implies sin x is unbounded.
(b) Dene
= 2 inf S .
Show that cos

2
= 0, sin

2
= 1. Then prove that sin(x + 2) = sin x and
cos(x + 2) = cos x.
(c) Dene tan x =
sin x
cos x
for all x such that cos x ,= 0. Show that tan

4
= 1.
14. Please continue to use only the facts about trigonometry established in problems 9
and 10.
(a) Show that the derivative of tan x is sec
2
x, where we dene sec x = 1/ cos x.
(b) From now on, restrict the domain of tan x to (/2, /2). Show that tan x is
strictly increasing on this domain. Show that its image is R. Therefore tan x has
an inverse function arctan x mapping R (/2, /2). By problem 1, arctan x
is of class C
1
, and in particular continuous.
(c) Show that sec
2
(arctan x) = 1+x
2
for all x R. (It is not rigorous to draw a little
right triangle with an angle = arctan x in one corner. Problems 910 involve no
notion of angle or two-dimensional geometry.)
(d) By the denition of inverse function, tan(arctan x) = x for all x R. Use the
Chain Rule to show the derivative of arctan x is
1
1+x
2
.
(e) In the geometric series 1 + x + x
2
+ x
3
+ , substitute x
2
for x. Show that
this power series converges to
1
1+x
2
for x (1, 1). (Aside: is this uniform
convergence?)
(f) Consider the power series
A(x) = x
x
3
3
+
x
5
5

x
7
7
+
Show that this denes an analytic function on (1, 1). Show that A(x) and
arctan x have the same derivative. Therefore A(x)arctan x is a constant. Check-
ing at x = 0 to see what this constant is, show that A(x) = arctan x on (1, 1).
(g) Show that arctan x is uniformly continuous on R.
(h) Since A(x) equals arctan x on (1, 1), it is uniformly continuous on that open
interval. By the last problem set, it has a unique continuous extension to [1, 1].
Conclude that

4
= 1
1
3
+
1
5

1
7
+
89
15. Abels limit theorem. Suppose that f : (1, 1] R is a function such that (a) f
is continuous at x = 1 and (b) for all x (1, 1), f(x) =

n=0
a
n
x
n
for some power
series that converges for all x (1, 1). If, in addition,

a
n
converges, prove that

n=0
a
n
= f(1) .
Hint. For x (1, 1) write f
n
(x) =

n
k=0
a
k
x
k
and A
n
=

n
k=0
a
k
. Show that
f
n
(x) = (1 x)(A
0
+ + A
n1
x
n1
) + A
n
x
n
.
Let n to get a dierent representation for f(x). Next denote A =

k=0
a
k
and
write
f(x) A = f(x) (1 x)

n=0
Ax
n
.
Use the representation of f(x) above to bound this dierence for x near 1.
90
8 Integration
The standard motivation for integration is to nd the area under the graph of a function.
There are other very important reasons to study integration and one is that integration is
a smoothing operation: the (indenite) integral of a function has more derivatives than the
original function does. Other motivations can be seen in abstract measure theory and the
application to, for instance, probability theory.
8.1 Denitions
We will start at the bottom and try to nd the area under a graph. We will place boxes
under the graph and sum the area in these boxes. The x-coordinates of the sides of these
boxes form an (ordered) partition. Although we have used this word before, it will take a
new meaning here.
Denition 8.1.1. A partition T of the interval [a, b] is a nite set x
1
, . . . , x
n
such that
a = x
1
< x
2
< < x
n
= b .
Given a partition and a bounded function f we can construct an upper sum and a lower
sum. To do this, we consider a subinterval [x
i
, x
i+1
] and let
m
i
= inf
x[x
i
,x
i+1
]
f(x) and M
i
= sup
x[x
i
,x
i+1
]
f(x) .
A box with base [x
i
, x
i+1
] and height M
i
contains the entire area below f in this interval,
whereas the box with the same base but height m
i
is contained in this area. (Here we are
thinking of f 0, so these statements are slightly dierent otherwise.) Counting up the
area of these boxes, we get the following denitions.
Denition 8.1.2. Given a partition T = x
1
< < x
n
of [a, b] and a bounded function
f : [a, b] R we dene the upper and lower sums of f relative to the partition T as
U(f, T) =
n1

i=1
M
i
(x
i+1
x
i
) and L(f, T) =
n1

i=1
m
i
(x
i+1
x
i
) .
There is a useful monotonicity property of upper and lower sums. To state this, we use
the following term. A partition Q of [a, b] is said to be a renement of T if T Q. This
means that we have just thrown in extra subintervals to T to form Q.
Lemma 8.1.3. Let f : [a, b] R be bounded and Q a renement of T. Then
U(f, Q) U(f, T) and L(f, Q) L(f, T) .
91
Proof. By iteration (or induction) it suces to show the inequalities in the case that Q has
just one more point than T. So take T = x
1
< < x
n
and Q = x
1
< < x
k
< t <
x
k+1
< < x
n
. Since most intervals are unchanged,
U(f, T) U(f, Q) = M
k
(x
k+1
x
k
)
_
sup
y[x
k
,t]
f(y)
_
(y x
k
)
_
sup
z[t,x
k+1
]
f(z)
_
(x
k+1
y)
M
k
(x
k+1
x
k
) M
k
(y x
k
) M
k
(x
k+1
y)
= 0 .
The argument for lower sums is similar.
The above lemma says that upper sums decrease and lower sums increase when we add
more points into the partition. Since we are thinking of taking very ne partitions, we dene
the upper and lower integrals
_
b
a
f(x) dx = inf
P
U(f, T) and
_
b
a
f(x) dx = sup
P
L(f, T)
for bounded f : [a, b] R. Note that these are dened for all bounded f.
Denition 8.1.4. If f : [a, b] R then f is integrable (written f 1([a, b])) if
_
b
a
f(x) dx =
_
b
a
f(x) dx .
In this case we write
_
b
a
f(x) dx for the common value.
Note the following property of upper and lower sums and integrals.
For any partition T of [a, b] and bounded function f : [a, b] R,
L(f, T)
_
b
a
f(x) dx
_
b
a
f(x) dx U(f, T) .
Proof. The only inequality that is not obvious is the one between the integrals. To
show this, we rst let > 0. By denition of the upper and lower integrals, there exist
partitions T
1
and T
2
of [a, b] such that
L(f, T) >
_
b
a
f(x) dx /2 and U(f, Q) <
_
b
a
f(x) dx + /2 .
Taking T

to be the common renement of T and Q (that is, their union), we can use
the previous lemma to nd
_
b
a
f(x) dx < L(f, T) + /2 L(f, T

) + /2 U(f, T

) + /2
U(f, Q) + /2 <
_
b
a
f(x) dx + .
Taking 0 we are done.
92
There is an equivalent characterization of integrability. It is useful because the condition
involves only one partition, whereas when dealing with both upper and lower integrals one
would need to approximate using two partitions.
Theorem 8.1.5. Let f : [a, b] R be bounded. f is integrable if and only if for each > 0
there is a partition T of [a, b] such that U(f, T) L(f, T) < .
Proof. Suppose rst that f is integrable and let > 0. Then the upper and lower integrals
are equal. Choose T
1
such that L(f, T
1
) >
_
b
a
f(x) dx/2 and U(f, T
2
) <
_
b
a
f(x) dx+/2.
Taking T to be the common renement of T
1
and T
2
we nd
L(f, T) L(f, T
1
) >
_
b
a
f(x) dx /2
and
U(f, T) U(f, T
2
) <
_
b
a
f(x) dx + /2 .
Combining these two gives U(f, T) L(f, T) < .
Conversely suppose that for each > 0 we can nd a partition T such that U(f, T)
L(f, T) < . Then
_
b
a
f(x) dx U(f, T) < L(f, T) +
_
b
a
f(x) dx + .
Since > 0 is arbitrary, we nd
_
b
a
f(x) dx
_
b
a
f(x) dx. The other inequality is obvious, so
the upper and lower integrals are equal. In other words, f 1.
Using this we can show that all continuous functions are integrable.
Theorem 8.1.6. Let f : [a, b] R be continuous. Then f is integrable.
Proof. Since [a, b] is compact, f is uniformly continuous. Then given > 0 we can nd > 0
such that if x, y [a, b] with [xy[ < then [f(x)f(y)[ < /(2(ba)). Now construct any
partition T of [a, b] such that, writing T = x
1
< x
2
< < x
n
, we have [x
i
x
i+1
[ <
for all i = 1, . . . , n 1. Then in each subinterval [x
i
, x
i+1
], we have
[f(x) f(y)[ < /2 for all x, y [x
i
, x
i+1
] .
This gives M
i
m
i
/(2(b a)) < /(b a). Therefore
U(f, T) L(f, T) =
n1

i=1
(M
i
m
i
)(x
i+1
x
i
) < /(b a)
n1

i=1
(x
i+1
x
i
) = .
Using the last theorem, we are done.
93
So we know now that all continuous functions are integrable. There are some other
questions we need to resolve.
1. Which other functions are integrable?
2. Which functions are not integrable?
3. How do we compute integrals?
Examples.
Let f be the indicator function of the rationals:
f(x) =
_
1 if x Q
0 if x / Q
.
We will now show that f is not integrable on any [a, b].Indeed, let T be any partition
of [a, b], written as x
1
< x
2
< < x
n
. Then for each subinterval [x
i
, x
i+1
], we have
M
i
= sup
x[x
i
,x
i+1
]
f(x) = 1 and m
i
= 0 .
Therefore U(f, T) L(f, T) =

n1
i=1
(M
i
m
i
)(x
i+1
x
i
) = b a. Choosing any > 0
that is less than ba, we see that there is no partition T such that U(f, T)L(f, T) < .
Therefore f / 1.
Every monotone function is integrable. Indeed, take f : [a, b] R to be nondecreasing.
If T
n
is the partition
T
n
=
_
a < a +
b a
n
< a +
2(b a)
n
< < b
_
,
then we can compute
U(f, T
n
) L(f, T
n
) =
n1

i=0
(M
i
m
i
)(1/n)
=
1
n
n1

i=1
_
f
_
a +
i(b a)
n
_
f
_
a +
(i 1)(b a)
n
__
=
1
n
(f(b) f(a)) .
Then given > 0 take n such that (f(b) f(a))/n < . This shows that f 1.
All functions with countably many discontinuities are integrable. One example will be
in the problem set. It is actually possible to show that some functions with uncountably
many discontinuities are integrable, but we will not address this.
94
Let us prove a simple example, the function f : [0, 1] given by
f(x) =
_
0 x 1/2
1 x > 1/2
.
Given > 0 we construct a partition containing a very small subinterval around the
discontinuity. Let T = 0 < 1/2 /3 < 1/2 + /3 < 1. Then
U(f, T) L(f, T) =
2

i=1
(M
i
m
i
)(x
i+1
x
i
)
= 0(1/2 /3) + 1(2/3) + 0(1/2 2/3) = 2/3 < .
In this example we did not need to care about subintervals away from the discontinuity
because the function is constant there (and thus has M
i
= m
i
). In general we would
have to have construct a partition with somewhat more complicated parts there too
(possibly using continuity).
Let us now give an example of computing an integral by hand. Consider f : [0, 1] R
given by f(x) = x
2
. Take a partition T
n
to be
T
n
=
_
0 <
1
n
<
2
n
< <
n 1
n
< 1
_
.
The upper sum is
U(f, T
n
) =
n1

i=0
f
_
i + 1
n
_
(1/n) = (1/n)
n1

i=0
_
i + 1
n
_
2
=
1
n
3
n

i=1
i
2
=
n(n + 1)(2n + 1)
6n
3
1/3.
Similarly, L(f, T
n
) 1/3. This means that
_
1
0
f(x) dx 1/3 and
_
1
0
f(x) dx 1/3, giving
_
1
0
x
2
dx = 1/3 .
8.2 Properties of integration
Here we state many properties of the integral. Because of the third item, we dene
_
a
b
f(x) dx =
_
b
a
f(x) dx
and the third item remains valid for any a, b, d.
95
Proposition 8.2.1. Let f, g : [a, b] R be integrable and c R.
1. The functions f + g and cf are integrable with
_
b
a
(f + g)(x) dx =
_
b
a
f(x) dx +
_
b
a
g(x) dx
and
_
b
a
(cf)(x) dx = c
_
b
a
f(x) dx .
2. If f(x) g(x) for all x [a, b] then
_
b
a
f(x) dx
_
b
a
g(x) dx .
3. If d (a, b) then f is integrable on [a, d] and on [d, b] with
_
b
a
f(x) dx =
_
d
a
f(x) dx +
_
b
d
f(x) dx .
Proof. Let us show item 1 rst. For > 0, take T and Q to be partitions such that
L(f, T)
_
b
a
f(x) dx U(f, T) < L(f, T) + /2
and
L(g, Q)
_
b
a
g(x) dx U(g, Q) < L(g, Q) + /2
Let T

be their common renement so that


L(f, T

) + L(g, T

)
_
b
a
f(x) dx +
_
b
a
g(x) dx < L(f, T

) + L(g, T

) + .
On the other hand you can check that
L(f, T

) + L(g, T

) L(f + g, T

) U(f + g, T

) U(f, T

) + U(g, T

) .
(Here we have used that for bounded functions h
1
and h
2
and any set S R, inf
xS
(h
1
(x) +
h
2
(x)) inf
xS
h
1
(x) + inf
xS
h
2
(x) and the corresponding statement for suprema.) So we
nd both
U(f + g, T

) L(f + g, T

) <
and

L(f + g, T

)
_
b
a
f(x) dx
_
b
a
g(x) dx

< .
96
The rst statement implies that f +g is integrable and

L(f + g, T

)
_
b
a
(f + g)(x) dx

< .
Combining this with the second statement gives

_
b
a
(f + g)(x) dx
_
b
a
f(x) dx
_
b
a
g(x) dx

< 2 .
Since is arbitrary this gives the result.
If c R suppose rst that c 0. Then for any set S R and bounded function
h : S R we have
sup
xS
(ch)(x) = c sup
xS
h(x) and inf
xS
(ch)(x) = c inf
xS
h(x) .
Therefore for any partition T of [a, b],
U(cf, T) = cU(f, T) and L(cf, T) = cL(f, T) .
So given that f is integrable and > 0, we can choose a partition T such that U(f, T)
L(f, T) < /c. Then U(cf, T) L(cf, T) < , proving that cf is integrable. Furthermore,
L(cf, T) = cL(f, T) c
_
b
a
f(x) dx cU(f, T = U(cf, T) L(cf, T) + ,
giving

c
_
b
a
f(x) dx L(cf, T)

< . However we already know that


L(cf, T)
_
b
a
(cf)(x) dx U(cf, T) < L(cf, T) + ,
giving

_
b
a
(cf)(x) dx L(cf, T)

< . Combining these two and taking 0 proves


_
b
a
(cf)(x) dx = c
_
b
a
f(x) dx.
If instead c < 0 then we rst prove the case c = 1. Then we have for any partition T of
[a, b] that U(f, T) = L(f, T) and L(f, T) = U(f, T). Thus is U(f, T) L(f, T) <
we also have U(f, T) L(f, T) < , proving that f is integrable. Further, as above,
L(f, T)
_
b
a
(f)(x) dx < L(f, T) +
and
U(f, T)
_
b
a
f(x) dx < U(f, T) + .
Combining these and taking 0 gives
_
b
a
(f)(x) dx =
_
b
a
f(x) dx. Last, for any c < 0
we note that if f is integrable, so is f and since c > 0, so is (c)(f) = cf. Further,
_
b
a
(cf)(x) dx =
_
b
a
((cf))(x) dx =
_
b
a
(cf)(x) dx = (c)
_
b
a
f(x) dx
= c
_
b
a
f(x) dx .
97
For the second item, we just use the fact that for every partition T of [a, b], U(f, T)
U(g, T) whenever f(x) g(x) for all x [a, b]. So given > 0, choose T such that
U(g, T) <
_
b
a
g(x) dx + . Now
_
b
a
f(x) dx U(f, T) U(g, T) <
_
b
a
g(x) dx + .
This is true for all > 0 so we deduce that
_
b
a
f(x) dx
_
b
a
g(x) dx.
We move to the third item. Given > 0 choose a partition T of [a, b] such that U(f, T)
L(f, T) < . Now rene T to a partition Q by adding the point d. Call T
1
the partition of
[a, d] obtained from the points of Q up to d and T
2
the remaining points of Q (including d)
that form a partition of [d, c]. Then
U(f, T
1
) L(f, T
1
) =

i:x
i
<d
(M
i
m
i
)(x
i+1
x
i
) U(f, T) L(f, T) < .
This means f is integrable on [a, d]. Similarly it is integrable on [d, c]. Furthermore, we have
L(f, T
1
)
_
d
a
f(x) dx L(f, T
1
) + ,
and
L(f, T
2
)
_
c
d
f(x) dx L(f, T
2
) + .
Combining these with
L(f, T
1
) + L(f, T
2
) = L(f, Q)
_
b
a
f(x) dx L(f, T
1
) + L(f, T
2
) + ,
We nd

_
b
a
f(x) dx
_
d
a
f(x) dx
_
c
d
f(x) dx

< 3 .
Taking to zero gives the result.
Let us give one more important property of the integral.
Proposition 8.2.2 (Triangle inequality for integrals). Let f : [a, b] R be integrable. Then
so is [f[ and

_
b
a
f(x) dx

_
b
a
[f(x)[ dx .
Proof. Let > 0 and choose a partition T of [a, b] such that U(f, T) L(f, T) < . For the
proof we use the fact (which you can check using the triangle inequality) that for any set
S R and bounded function g : S R,
sup
xS
[f(x)[ inf
xS
[f(x)[ sup
xS
f(x) inf
xS
f(x) .
98
This implies that
U([f[, T) L([f[, T) U(f, T) L(f, T) < ,
so [f[ 1.
To prove the inequality in the proposition, note that f(x) [f(x)[ for all x, so
_
b
a
f(x) dx
_
b
a
[f(x)[ dx. Similarly f(x) [f(x)[, so
_
b
a
f(x) dx =
_
b
a
(f(x)) dx
_
b
a
[f(x)[ dx.
Combining these gives the inequality.
In fact this is an instance of a more general theorem, stated in Rudin. We will not prove
it; the proof is similar to the above (but more complicated).
Theorem 8.2.3. Suppose that f : [a, b] [c, d] is integrable and : [c, d] R is continuous.
Then f is integrable.
Proof. See Rudin, Thm. 6.11.
From this theorem we nd more integrable functions:
If f is integrable on [a, b] then so is f
2
. This follows by taking (x) = x
2
in the above
theorem.
If f and g are integrable on [a, b] then so is fg. This follows by writing
fg =
1
4
_
(f + g)
2
(f g)
2

.
8.3 Fundamental theorems
Of course we do not always have to compute integrals by hand. As we learn in calculus, we
can compute an integral if we know the antiderivative of the function. Stated precisely,
Theorem 8.3.1 (Fundamental theorem of calculus part I). Let f : [a, b] R be integrable
and F : [a, b] R a continuous function such that F

(x) = f(x) for all x (a, b). Then


F(b) F(a) =
_
b
a
f(x) dx .
Proof. Since f is integrable, given > 0 we can nd a partition T such that U(f, T)
L(f, T) < . We will use the mean value theorem to relate values of f in the subintervals to
values of F. That is, writing T = x
1
< < x
n
, we can nd for each i = 1, . . . , n 1 a
point c
i
(x
i
, x
i+1
) such that
F(x
i+1
) F(x
i
) = f(c
i
)(x
i+1
x
i
) .
Then we have
L(f, T)
n1

i=1
f(c
i
)(x
i+1
x
i
) L(f, T) + .
99
Furthermore
L(f, T)
_
b
a
f(x) dx L(f, T) + .
Using the equation derived by the mean value theorem above,
n1

i=1
f(c
i
)(x
i+1
x
i
) =
n1

i=1
[F(x
i+1
) F(x
i
)] = F(b) F(a) .
Combining with the above,

_
b
a
f(x) dx [F(b) F(a)]

<
and we are done.
As we learn in calculus, we are able now to say, for example, that
_
b
a
cos x dx = sin(b) sin(a)
and
_
b
a
x
n
dx =
1
n + 1
_
b
n+1
a
n+1

.
There is a second fundamental theorem of calculus. Whereas the rst is about integrating
a derivative, the second is about dierentiating an integral. Both of them say that integration
and dierentiation are inverse operations. For example, in the rst, when we start with F
and dierentiate to get a function f, we integrate back to get F (in a sense).
Theorem 8.3.2 (Fundamental theorem of calculus part II). Let f : [a, b] R be continuous.
Dene F : [a, b] R by
F(x) =
_
x
a
f(t) dt .
Then F is dierentiable on [a, b] with F

(x) = f(x) for all x.


Proof. Let x [a, b); the case of x = b is similar and is calculated as a left derivative. For
h > 0,
F(x + h) F(x)
h
=
1
h
__
x+h
a
f(t) dt
_
x
a
f(t) dt
_
=
1
h
_
x+h
x
f(t) dt .
Let > 0. Since f is continuous at x we can nd > 0 such that if [t x[ < then
[f(t) f(x)[ < . This means that if 0 < h < then

F(x + h) F(x)
h
f(x)

=
1
h

_
x+h
x
f(t) dt
_
x+h
x
f(x) dt

=
1
h

_
x+h
x
(f(t) f(x)) dt

1
h
_
x+h
x
[f(t) f(x)[ dt
(1/h)h = .
100
In other words,
lim
h0
+

F(x + h) F(x)
h
f(x)

= 0 .
A similar argument works for the left limit (in the case that x ,= a), using
F(x h) F(x)
h
=
1
h
_
xh
x
f(t) dt .
and completes the proof.
8.4 Change of variables, integration by parts
We will now prove the u-substitution rule for integrals. As you know from calculus, this
is a valuable tool to solve for the value of many denite integrals. The proof is essentially
a combination of the chain rule and the fundamental theorem of calculus. Note that in its
statement, the range of f is a closed interval. This follows from the fact that f is continuous
on a closed interval. Indeed, the image must be connected and compact, therefore a closed
interval as well.
Theorem 8.4.1 (Substitution rule). Let f : [a, b] R be C
1
and write [c, d] for the range
of f. If g : [c, d] R is continuous then
_
f(b)
f(a)
g(t) dt =
_
b
a
g(f(x))f

(x) dx .
Proof. Dene a function F : [c, d] R by
F(x) =
_
x
f(a)
g(t) dt .
Then because g is continuous, by the fundamental theorem of calculus II, F is dierentiable
and F

(x) = g(x) (giving actually F C


1
). Furthermore as f is dierentiable, the function
F f : [a, b] R is dierentiable with (F f)

(x) = F

(f(x))f

(x). Last, F

is continuous
and f is integrable, so by Theorem 8.2.3, F

f is integrable. Since f

is continuous, it is
also integrable, so the product of F

f and f

is integrable. By the fundamental theorem


of calculus I,
F(f(b)) F(f(a)) =
_
b
a
F

(f(x))f

(x) dx .
Plugging in,
_
f(b)
f(a)
g(t) dt =
_
b
a
g(f(x))f

(x) dx .
Just as the substitution rule is related to the chain rule, integration by parts is related
to the product rule.
101
Theorem 8.4.2 (Integration by parts). Let f, g : [a, b] R be C
1
. Then
_
b
a
f(x)g

(x) dx = f(b)g(b) f(a)g(a)


_
b
a
f

(x)g(x) dx .
Proof. This follows from the product rule since both of f

g and fg

is integrable.
8.5 Exercises
1. Let f : [0, 1] R be continuous.
(a) Suppose that f(x) 0 for all x and that
_
1
0
f(x) dx = 0. Show that f is identically
zero.
(b) Suppose that f is not necessarily non-negative but that
_
b
a
f(x) dx = 0 for all
a, b [0, 1] with a < b. Show that f is identically zero.
2. Let f : [0, 1] R be continuous. Show that
lim
n
_
1
0
x
n
f(x) dx = 0 .
Hint. For c near 1, consider [0, c] and [c, 1] separately.
3. Let f : [0, 1] R be continuous. Prove that
lim
n
__
1
0
[f(x)[
n
dx
_
1/n
= max
x[0,1]
[f(x)[ .
4. Dene f : [0, 1] R by
f(x) =
_
0 if x / Q
1
n
if x =
m
n
Q, where m and n have no common divisor
.
Use Theorem 6.6 in Rudin to prove that f is Riemann integrable.
5. Let f and g be continuous functions on [0, 1] with g(x) 0 for all x. Show there exists
c [0, 1] such that
_
1
0
f(x)g(x) dx = f(c)
_
1
0
g(x) dx .
6. (a) Show that the Euler-Mascheroni constant
= lim
n
_
n

k=1
1
k
log n
_
exists .
102
Hint. Write the above quantity as
1
n
+
n1

k=1
1
k

_
n
1
dx
x
=
1
n
+
n1

k=1
_
k+1
k
_
1
k

1
x
_
dx .
Show the last sum converges.
(b) Use the last part to nd the limit
lim
n
_
1
n
+ +
1
2n
_
.
7. Let f
n
be a sequence of continuous functions on [0, 1]. Suppose that f
n
converges
uniformly to a function f. Recall from last problem set that this means that for any
> 0 there exists N such that n N implies that [f
n
(x) f(x)[ < for all x [0, 1].
Show that
lim
n
_
1
0
f
n
(x) dx =
_
1
0
f(x) dx .
Give an example to show that we cannot only assume f
n
f pointwise (meaning that
for each xed x [0, 1], f
n
(x) f(x)).
Hint. Use the inequality

_
1
0
g(x) dx


_
1
0
[g(x)[ dx, valid for any integrable g.
8. Suppose that f
n
is a sequence of functions in C
1
([0, 1]) and that the sequence f

converges uniformly to some function g. Suppose there exists some c [0, 1] such that
the sequence f
n
(c) converges. By the fundamental theorem of calculus, we can write
for x [0, 1]
f
n
(x) = f
n
(c) +
_
x
c
f

n
(t) dt .
(a) Show that f
n
converges pointwise to some function f.
(b) Show that f is dierentiable and f

(x) = g(x) for all x. (You will need to use


Theorem 7.12 in Rudin.)
Remark. The above result gives a method to prove the form of a derivative of a power
series. Suppose that f(x) =

n=0
a
n
x
n
has radius of convergence R > 0. Setting
f
n
(x) =
n

j=0
a
j
x
j
and g(x) =

j=1
ja
j
x
j1
,
one can show using the Weierstrass M-test that for any r with 0 < r < R, f

n
g
uniformly on (r, r). We can then conclude that f

(x) = g(x).
103
9. You can solve either this question or the next one. In this problem we will
show part of Stirlings formula. It states that
lim
n
n!
n
n
e
n

2 .
We will only show the limit exists.
(a) Show that
log
_
n
n
n!
_
=
n1

k=1
__
k+1
k
log(x/k) dx
_
+ n 1 log n .
Use a change of variable u = x/k and continue to show that this equals
n1

k=1
_
k
_
1/k
0
[log(1 + u) u] du +
1
2k
_
+ n 1 log n .
(b) Prove that
n!
n
n
e
n

n
converges if and only if
lim
n
n1

k=1
_
k
_
1/k
0
[log(1 + u) u] du
_
exists .
(c) Show that for u [0, 1],
u
2
/2 log(1 + u) u 0
and deduce that the limit in part (b) exists.
Hint. Use Taylors theorem.
10. You can solve either this question or the previous one. In this question, you
will work out an alternate derivation of existence of the limit in Stirlings formula.
(a) Dene a continuous function g such that g(n) = log n for n N and g(x) is linear
in each interval [n, n + 1]. Show that for n large enough,
log n +
x n
n + 1
g(x) log x log n +
x n
n
for x [n, n + 1] .
(b) Let S
n
=
_
n
1
[log x g(x)] dx. Use part (a) to show that (S
n
) is Cauchy and thus
converges. Compute directly that
_
n
1
log x dx = nlog n n + 1
and
_
n
1
g(x) dx = log n!
1
2
log n. Conclude that the limit in Stirlings formula
exists.
104
A Real powers
The question is the following: we know what 2
2
or 2
3
means, or even 2
2/3
, the number whose
cube equals 2
2
. But what does 2

2
mean? We will give the denition Rudin has in the
exercises of Chapter 1. We will only use the following facts for r, s > 0, n, m Z:
r
n+m
= r
n
r
m
.
(r
n
)
m
= r
mn
.
(rs)
n
= r
n
s
n
.
if r > 1 and m n then r
m
r
n
. If r < 1 and m n then r
m
r
n
.
if s < r and n > 0 then s
n
< r
n
. If s < r and n < 0 then s
n
> r
n
.
A.1 Natural roots
We rst dene the n-th root of a real number, for n N.
Theorem A.1.1. For any r > 0 and n N there exists a unique positive real number y
such that y
n
= r.
Proof. The proof is Theorem 1.21 in Rudin. The idea is to construct the set
S = x > 0 : x
n
r
and to show that S is nonempty, bounded above, and thus has a supremum. Calling y this
supremum, he then shows y
n
= r. The proof of this is somewhat involved and is similar to
our proof (from the rst lecture) that a Q : a
2
< 2 does not have a greatest element.
To show there is only one such y, we note that 0 < y
1
< y
2
implies that y
n
1
< y
n
2
and so
if y
1
,= y
2
are positive then y
n
1
,= y
n
2
.
This denition extends to integer roots.
Denition A.1.2. If r > 0 and n N we dene r
1/n
as the unique positive real number y
such that y
n
= 1/r.
A.2 Rational powers
The above denitions allow us to dene rational powers.
Denition A.2.1 (Preliminary denition of rational powers). If r > 0 and m, n N we
dene r
m/n
to be the unique positive real number y such that y
n
= r
m
. Also r
m/n
is dened
as (1/r)
m/n
.
Because a rational number can have more than one representation m/n we need to show
this is well dened.
105
Proposition A.2.2. If a positive a Q can be represented by m/n and p/q for m, n, p, q N
then for all r > 0,
r
m/n
= r
p/q
.
Proof. First note that (r
m/n
)
nq
= ((r
m/n
)
n
)
q
= r
mq
and (r
p/q
)
nq
= ((r
p/q
)
q
)
n
= r
pn
. However
as m/n = p/q we have pn = mq and so these numbers are equal. There is a unique nq-th
root of this number, so r
m/n
= r
p/q
.
Note that the above proof applies to negative rational powers: suppose that r > 0 and
a Q is negative such that a = m/n = p/q. Then
r
m/n
= (1/r)
m/n
= (1/r)
p/q
= r
p/q
.
Denition A.2.3 (Correct denition of rational powers). If r > 0 and a > 0 is rational
we dene r
a
= r
m/n
for any m, n N such that a = m/n. If a < 0 is rational we dene
r
a
= (1/r)
a
.
Properties of rational powers. Let a, b Q and r, s > 0.
If a = m/n for m Z and n N then r
a
is the unique positive number such that
(r
a
)
n
= r
m
.
Proof. For m 0 this is the denition. For m < 0, this is because (r
a
)
n
= ((1/r)
a
)
n
=
(1/r)
m
= r
m
and if s is any other positive number satisfying s
n
= r
m
then uniqueness
of n-th roots gives s = r
a
.
r
a+b
= r
a
r
b
.
Proof. Choose m, p Z and n, q N such that a = m/n and b = p/q. Then a + b =
mq+np
nq
and therefore r
a+b
is the unique positive number such that (r
a+b
)
nq
= r
mq+np
.
But we can just compute
(r
a
r
b
)
nq
= ((r
a
)
n
)
q
((r
b
)
q
)
n
= r
mq
r
np
= r
mq+np
.
And by uniqueness we get r
a
r
b
= r
a+b
.
(r
a
)
b
= r
ab
.
Proof. Write a = m/n and b = p/q for m, p Z and n, q N. Then r
ab
is the unique
positive number such that (r
ab
)
nq
= r
mp
. But
((r
a
)
b
)
nq
= (((r
a
)
b
)
q
)
n
= ((r
a
)
p
)
n
= ((r
a
)
n
)
p
= (r
m
)
p
= r
mp
,
giving (r
a
)
b
= r
ab
.
(rs)
a
= r
a
s
a
.
106
Proof. Again write a = m/n for m Z and n N. Then (rs)
a
is the unique positive
number such that ((rs)
a
)
n
= (rs)
m
. But
(r
a
s
a
)
n
= (r
a
)
n
(s
a
)
n
= r
m
s
m
= (rs)
m
.
If r > 1 and a b then r
a
r
b
. If r < 1 and a b are rational then r
a
r
b
.
Proof. Suppose rst that r > 1 and a 0 with a = m/n for m, n N. Then if r
a
< 1,
we nd r
m
< 1
n
= 1, a contradiction, as r
m
> 1. So r
a
> 1. Next if a b then
a b 0 so r
ab
1. This gives r
a
= r
ab
r
b
r
b
.
If r < 1 then r
a
(1/r)
a
= 1
a
= 1, so r
a
= (1/r)
a
. Similarly r
b
= (1/r)
b
. So since
1/r > 1 we get r
a
= (1/r)
a
(1/r)
b
= r
b
. Multiplying both sides by r
a
r
b
we get
r
a
r
b
.
If s < r and a > 0 then s
a
< r
a
. If s < r and a < 0 then s
a
> r
a
.
Proof. Let a = m/n with m, n N. Then if s
a
r
a
we must have s
m
= (s
a
)
n

(r
a
)
n
= r
m
. But this is a contradiction since s
m
< r
m
. This proves the rst statement.
For the second, 1/s > 1/r so s
a
= (1/s)
a
> (1/r)
a
= r
a
.
A.3 Real powers
We dene a real power as a supremum of rational powers.
Denition A.3.1. Given r > 1 and t R we set
r
t
= supr
a
: a Q and a t .
If 0 < r < 1 then dene r
t
= (1/r)
t
.
Proposition A.3.2. If a Q then for r > 0, the denition above coincides with the rational
denition.
Proof. For this proof, we take r
a
to be the dened as in the rational powers section.
Suppose rst that r > 1. Clearly r
a
r
b
: b Q and b a. So to show it is the
supremum we need only show it is an upper bound. This follows from the fact that b a
implies r
b
r
a
(proved above).
If 0 < r < 1 then r
a
(r
1
)
a
= (1/r)
a
so the denitions coincide here as well.
Properties of real powers. Let t, u R and r, s > 0.
r
t+u
= r
t
r
u
.
107
Proof. We will use the following statement, proved on the homework. If A and B are
nonempty subsets of [0, ) which are bounded above then dene AB = ab : a
A, b B. We have
sup(AB) = sup Asup B . (7)
It either of the sets consists only of 0, then the supremum of that set is 0 and both
sides above are 0. Otherwise, both sets (and therefore also AB) contain positive
elements. For any element c AB we have c = ab for some a A, b B. Therefore
c = ab sup Asup B and therefore this is an upper bound for AB. As sup(AB) is the
least upper bound, we get sup(AB) sup Asup B. Assuming now for a contradiction
that we have strict inequality, because sup A > 0 we also have sup(AB)/ sup A < sup B.
Thus there exists b B such that sup(AB)/ sup A < b. As b must be positive, we also
have sup(AB)/b < sup A and there exists a A such that sup(AB)/b < a, giving
sup(AB) < ab. This is clearly a contradiction.
Now to prove the property, suppose rst that r > 1. By the statement we just proved,
we need only show that
r
b
: b Q and b t + u = AB ,
where A = r
c
: c Q and c t and B = r
d
: d Q and d u. (This is because
these are sets of non negative numbers.) This holds because each rational b t + u
can be written as a sum of two rationals c, d such that c t and d u.
For 0 < r < 1 we have r
t+u
= (1/r)
(t+u)
= (1/r)
(t)+(u)
= (1/r)
t
(1/r)
u
= r
t
r
u
.
(rs)
t
= r
t
s
t
.
Proof. We rst note that
r
t
= 1/r
t
. (8)
This is true because r
t
r
t
= r
0
= 1, so r
t
= 1/r
t
.
For the property, if r, s > 1 then we can just use equation (7), noting that (rs)
a
:
a Q and a t = AB, where A = r
a
: a Q and a t and B = s
a
: a
Q and a t. If 0 < r < 1 but s > 1 with rs > 1 we get (rs)
t
/r
t
= (rs)
t
(1/r)
t
.
We now use equation (7) again, noting that s
a
: a Q and a t = AB, where
A = (rs)
a
: a Q and a t and B = (1/r)
a
: a Q and a t. This gives
(rs)
t
/r
t
= s
t
. This same proof works if r > 1 but 0 < s < 1 with rs > 1. If 0 < r < 1
but s > 1 with rs < 1 we consider s
t
/(rs)
t
= s
t
(1/(rs))
t
and use the above argument.
This also works in the case r > 1 but 0 < s < 1 with rs < 1. Finally, if 0 < r < 1 and
0 < s < 1 then (rs)
t
= (1/(rs))
t
= ((1/r)(1/s))
t
= (1/r)
t
(1/s)
t
= r
t
s
t
.
(r
t
)
u
= r
tu
.
108
Proof. We will rst show the equality in the case r > 1 and t, u > 0. We begin with
the fact that (r
t
)
u
is an upper bound for r
a
: a Q and a tu. So let a tu be
rational and assume further that a > 0. In this case we can write a = bc for b, c Q
and b t, c u. By properties of rational exponents, we have r
a
= (r
b
)
c
. As r
b
r
t
(by denition) we get from monotonicity that (r
b
)
c
(r
t
)
c
. But this is an element of
the set (r
t
)
d
: d Q and d u, so (r
t
)
c
(r
t
)
u
. Putting these together,
r
a
= (r
b
)
c
(r
t
)
c
(r
t
)
u
.
This shows that (r
t
)
u
is an upper bound for r
a
: a Q and 0 < a tu. For the case
that a < 0 we can use monotonicity to write r
a
r
0
(r
t
)
u
. Putting this together
with the case a > 0 gives that (r
t
)
u
is an upper bound for r
a
: a Q and a tu
and therefore r
tu
(r
t
)
u
.
To prove that (r
t
)
u
r
tu
we must show that r
tu
is an upper bound for (r
t
)
a
: a
Q and a u. For this we observe that r
t
> 1. This holds because t > 0 and therefore
we can nd some rational b with 0 < b < t. Thus r
t
r
b
> r
0
= 1. Now let a be
rational with 0 < a u; we claim that (r
t
)
a
r
tu
. Proving this will suce since if
a < 0 then (r
t
)
a
< (r
t
)
0
= 1 r
tu
. To show the claim, note that if we show that
r
t
(r
tu
)
1/a
we will be done. This is by properties of rational exponents: we would
then have
(r
t
)
a

_
(r
tu
)
1/a
_
a
= r
tu
.
So we are reduced to proving that
supr
b
: b Q and b t (r
tu
)
1/a
,
which follows if we show that for each b Q such that b t, we have r
b
(r
tu
)
1/a
.
Again, this is true if r
ab
r
tu
because then r
b
= (r
ab
)
1/a
(r
tu
)
1/a
. But a t and
b u so r
ab
r
tu
. This completes the proof of (r
t
)
u
= r
tu
in the case r > 1 and
t, u > 0.
In the case r > 1 but t > 0 and u < 0, we can use (8):
(r
t
)
u
= 1/(r
t
)
u
= 1/r
tu
= r
tu
.
If instead r > 1 but t < 0 and u > 0,
(r
t
)
u
= (1/r
t
)
u
= 1/(r
t
)
u
= 1/r
tu
= r
tu
.
Here we have used that for s > 0 and x R, (1/s)
x
= 1/s
x
, which can be veried as
1 = (s(1/s))
x
= s
x
(1/s)
x
. Last if r > 1 but t < 0 and u < 0, we compute
(r
t
)
u
= ((1/r)
t
)
u
= 1/(r
t
)
u
= 1/r
tu
= r
tu
,
completing the proof in the case r > 1.
If 0 < r < 1 then
(r
t
)
u
= ((1/r)
t
)
u
= (1/r)
tu
= r
tu
.
109
If r > 1 and u t then r
u
r
t
. If 0 < r < 1 and u t then r
u
r
t
.
Proof. Assume r > 1. If u = 0 and t > 0 then we can nd a rational b such that
0 < b t, giving r
t
r
b
> r
0
= 1. For general u t we note 1 r
tu
, so multiplying
both sides by the (positive) r
u
we get the result.
If 0 < r < 1 then r
u
= (1/r)
u
(1/r)
t
= r
t
.
If s < r and t > 0 then s
t
< r
t
. If s < r and t < 0 then s
t
> r
t
.
Proof. First consider the case that s = 1. Then r > 1 and for any t > 0 we can nd a
rational b such that 0 < b < t. Therefore r
t
r
b
> r
0
= 1. For general s < r we write
r
t
= s
t
(r/s)
t
> s
t
. If t < 0 then s
t
= (1/s)
t
> (1/r)
t
= r
t
.
B Logarithm and exponential functions
B.1 Logarithm
We will use the integral denition of the natural logarithm. For x > 0 dene
log x =
_
x
1
1
t
dt .
This is dened because 1/x is continuous on (0, ).
Properties of logarithm.
log 1 = 0.
log is C

on (0, ).
log is strictly increasing and therefore injective.
Proof. The derivative is 1/x, which is positive.
For x, y > 0, log(xy) = log x + log y. Therefore log(1/x) = log x.
Proof. For a xed y > 0 dene f(x) = log(xy) log y. We have
f

(x) =
y
xy
=
1
x
=
d
dx
log x .
Therefore f(x) log x has zero derivative and must be a constant. Taking x = 1, we
get
f(1) log 1 = log y log y = 0 ,
so f(x) = log x. This completes the proof.
110
The range of log is R.
Proof. We rst claim that lim
x
log x = . Because log is strictly increasing, it
suces to show that the set log x : x R is unbounded above. Note that
log 2 =
_
2
1
1
t
dt
_
2
1
1
2
dt = 1/2 .
Therefore log(2
n
) = nlog 2 n/2. This proves the claim.
Because log is continuous and approaches innity as x , the intermediate value
theorem, combined with the fact that log 1 = 0, implies that the range of log includes
[0, ). Using log(1/x) = log x, we get all of R.
B.2 Exponential function
Because log is strictly increasing and dierentiable, exercise 1, Chapter 7 implies that the
inverse function of log exists and is dierentiable. We dene the inverse to be the exponential
function:
for x R, e
x
is the number such that log(e
x
) = x .
Its derivative can by found using the chain rule:
x = log(e
x
), so 1 =
1
e
x
d
dx
e
x
,
or
d
dx
e
x
= e
x
.
Properties of exponential.
e
0
= 1.
e
x
is C

on R.
For x, y R, e
x+y
= e
x
e
y
.
Proof. From properties of log,
log(e
x+y
) = x + y = log(e
x
) + log(e
y
) = log(e
x
e
y
) .
Since log is injective, this shows e
x+y
= e
x
e
y
.
e
x
> 0 for all x. Therefore the exponential function is strictly increasing.
Proof. Because e
x
is the inverse function of log x, which is dened on (0, ), its range
is (0, ), giving e
x
> 0.
111
For any x,
e
x
=

n=0
x
n
n!
.
Proof. This follows from Taylors theorem. For any x, the n-th derivative of the ex-
ponential function evaluated at x is simply e
x
. Therefore expanding at x = 0, for any
N 1,
e
x
=
N1

n=0
f
(n)
(0)
n!
x
n
+
f
N
(c
N
)
N!
x
N
=
N1

n=0
x
n
n!
+
e
c
N
N!
x
N
,
with c
N
some number between 0 and x. This remainder term is bounded by
e
c
N
N!
x
N
e
x
x
N
N!
0 ,
because x
N
/N! 0 as N . This follows because the ratio test gives convergence
of

x
n
/n!, so the n-th term must go to 0. By the corollary to Taylors theorem, we
get e
x
=

n=0
x
n
n!
.
Writing e = e
1
, the exponential function is the x-th power of e (dened earlier in terms
of suprema).
Proof. For ease of reading, write exp(x) for the function we have dened here and e
x
for the x-th power of e, dened in terms of suprema. Then for x = m/n Q with
n N, we have
(exp(x))
n
= exp(m/n)
n
= exp(m) = exp(1)
m
= e
m
.
Because e
m/n
was dened as the unique positive number y such that y
n
= e
m
, we have
exp(x) = e
x
. Generally for x R we dened
e
x
= supe
q
: q Q and q < x .
(This was the denition for exponents whose bases are 1, which is true in our case
because e
1
e
0
= 1.) Using equivalence over rationals,
e
x
= supexp(q) : q Q and q < x .
However exp is an increasing function, so writing S for the set whose supremum we take
above, exp(x) sup S = e
x
. On the other hand, because exp is continuous at x, we can
pick any sequence q
n
of rationals converging up to x and we have exp(q
n
) exp(x).
This implies that exp(x) r is not an upper bound for S for any r > 0 and therefore
exp(x) = sup S = e
x
.
112
We now show that the exponential function can be attained by the standard limit
e
x
= lim
n
_
1 +
x
n
_
n
.
First we use the binomial formula
_
1 +
x
n
_
n
=
n

j=0
_
n
j
_
_
x
n
_
j
=
n

j=0
n!
j!(n j)!
_
x
n
_
j
=
n

j=0
n(n 1) (n j + 1)
n
j
x
j
j!
.
To show the limit, let > 0. By convergence of

j=0
|x|
j
j!
, we may choose J such that

j=J+1
[x[
j
j!
< /3 .
Because

J
j=0
n(n1)(nj+1)
n
j
x
j
j!
is a nite sum and the j-th term approaches
x
j
j!
as n ,
we can pick N such that if n N then

j=0
n(n 1) (n j + 1)
n
j
x
j
j!

J

j=0
x
j
j!

< /3 .
Thus by the triangle inequality, we nd for n N,

_
1 +
x
n
_
n

j=0
x
j
j!

j=J+1
x
j
j!

j=0
n(n 1) (n j + 1)
n
j
x
j
j!

J

j=0
x
j
j!

j=J+1
n(n 1) (n j + 1)
n
j
x
j
j!

< 2/3 +
n

j=J+1
[x[
j
j!
< .
B.3 Sophomores dream
We end this appendix with a strange identity that is for some reason called the Sophomores
dream. It is
_
1
0
x
x
dx =

n=1
n
n
.
113
To prove this, we need to dene the function x
x
. It is given by x
x
= exp(x log x). So
the identity reads
_
1
0
e
xlog x
dx =

n=1
n
n
.
For this integral to make sense, as the integrand is not dened at x = 0, we must use the
right limit. Note that by lHopitals rule (which we didnt cover but it is in Rudin),
lim
x0
+
x log x = lim
x0
+
log x
1/x
= lim
x0
+
1/x
1/x
2
= 0 .
Using continuity of the exponential function, we nd
lim
x0
+
x
x
= 1 ,
so we can continuously extend x
x
to 0 by dening 0
0
= 1. Thus x
x
is then continuous on
[0, 1] and integrable.
To nd the integral we will use a power series expansion of e
x
:
e
x
=

n=0
x
n
n!
,
which has radius of convergence R = . Therefore by the remark after exercise 6, Chapter
8, for any M > 0, this series converges uniformly for x in [M, M]. (The proof uses the
Weierstrass M-test.) Because the number [x log x[ is bounded by e
1
on the interval [0, 1]
(do some calculus),
e
xlog x
=

n=0
(x log x)
n
n!
converges uniformly on [0, 1] .
We now use exercise 5, Chapter 8, which says that if (f
n
) is a sequence of continuous functions
that converges uniformly on [0, 1] to a function f then
_
1
0
f
n
(x) dx
_
1
0
f(x) dx. Noting
that an innite series of functions is just a limit of the sequence of partial sums (which
converges uniformly in our case), we get
_
1
0
x
x
dx =

n=0
_
1
0
(x log x)
n
n!
dx =

n=0
1
n!
_
1
0
(x log x)
n
dx .
Now we compute the integral
_
1
0
(x log x)
n
dx using integration by parts. We take
u = (log x)
n
and dv = x
n
dx to get du = (1)
n
n(log x)
n1
/x dx and v = x
n+1
/(n + 1):
_
1
0
(x log x)
n
dx =
(log x)
n
x
n+1
n + 1

1
0
(1)
n
n
n + 1
_
1
0
x
n
(log x)
n1
dx
=
n
n + 1
_
1
0
x
n
(log x)
n1
dx .
114
Repeating this, we nd
_
1
0
(x log x)
n
dx =
n!
(n + 1)
n+1
.
So plugging back in, we nd
_
1
0
x
x
dx =

n=0
1
(n + 1)
n+1
=

n=1
n
n
.
C Dimension of the Cantor set
In this section we will discuss how to assign a dimension to the Cantor set. One way is
through the use of Hausdor dimension. We will start with denitions and examples. This
treatment is based on notes of J. Shah from UChicago.
C.1 Denitions
For any set S R write [S[ for the diameter of S:
[S[ = sup[x y[ : x, y S .
For example, we have [[0, 1][ = 1, [Q [0, 1][ = 1 and [(0, 1) (2, 3)[ = 3.
Denition C.1.1. Let S R. A countable collection C
n
of subsets of R is called a
countable cover of S if
S

n=1
C
n
.
Note that the sets in a countable cover can be any sets whatsoever. For example, they
do not need to be open or closed.
Denition C.1.2. If C
n
is a countable collection of sets in R and > 0, the -total
length of C
n
is

n=1
[C
n
[

.
If > 1 then it has the eect of increasing the diameter (that is, [C
n
[

> [C
n
[) when
[C
n
[ is large (bigger than 1) and decreasing it when [C
n
[ is small (less than 1).
Example 1. Consider the interval [0, 1]. Let us build a very simple cover of this set by
xing n and choosing our (nite) cover C
1
, . . . , C
n
by
C
i
=
_
i 1
n
,
i
n
_
.
For instance, for n = 4 we have
[0, 1/4], [1/4, 1/2], [1/2, 3/4] and [3/4, 1] .
115
Computing the -total length of this cover:
n

i=1

_
i 1
n
,
i
n
_

=
n
n

.
The limit as n approaches is
_

_
if < 1
1 if = 1
0 if > 1
.
This result gives us some hint that the dimension of a set is related to the -total length of
countable covers of the set. Specically we make the following denition:
Denition C.1.3. If S R has [S[ < and > 0 we dene the -covered length of S as
H

(S) = inf
_

n=1
[C
n
[

: C
n
is a countable cover of S
_
.
The Hausdor dimension is dened as
dim
H
(S) = inf > 0 : H

(S) = 0 .
It is an exercise to show that for all 0 < < dim
H
(S), we have H

(S) > 0. Also, setting


0
0
= 1 then H
0
(S) > 0 for all S Thus we could dene the Hausdor dimension as
sup 0 : H

(S) > 0 .
Note that example 1 shows that dim
H
([0, 1]) 1. To show the other inequality, we must
show that for all < 1, H

([0, 1]) > 0. To do this, let C


n
be a countable cover of [0, 1].
We may replace the C
n
s by D
n
= C
n
[0, 1], since the D
n
s will still cover [0, 1] and will
have smaller -length. For < 1 we then have

n=1
[D
n
[

n=1
[D
n
[ ,
because [D
n
[ 1. Now it suces to show.
Lemma C.1.4. If D
n
is a countable cover of [0, 1] then

n=1
[D
n
[ 1 .
Proof. The proof is an exercise.
116
Assuming the lemma, we have H

([0, 1]) 1 for all < 1 and therefore dim


H
([0, 1]) = 1.
If the concept of Hausdor dimension is to agree with our current notion of dimension
it had better be that each subset of R has dimension no bigger than 1. This is indeed the
case; we can argue similarly to before. If S R has [S[ < then we can nd M > 0 such
that S [M, M]. Now for each n dene a cover C
1
, . . . , C
n
by
C
i
=
_
M + 2M
i 1
n
, M + 2M
i
n
_
.
As before, for > 1, the -total length of C
1
, . . . , C
n
is
n
_
2M
n
_

0 as n .
Therefore H

(S) = 0 and dim


H
(S) 1.
Example 2. Take S to be any countable set with nite diameter (for instance the rationals
in [0, 1]). We claim that dim
H
(S) = 0. To show this we must prove that for all > 0,
H

(S) = 0. Let > 0 and dene a countable cover of S by rst enumerating the elements
of S as s
1
, s
2
, . . . and for i N, letting C
i
be any interval containing s
i
of length (/2
n
)
1/
(note that this is a positive number). Then the -total length of the cover is

n=1

2
n
= ;
therefore H

(S) . This is true for all > 0 so H

(S) = 0.
C.2 The Cantor set
Let S be the Cantor set. To remind you, the construction is as follows. We start with
S
0
= [0, 1]. We remove the middle third of S
0
to get S
1
= [0, 1/3] [2/3, 1]. In general, at
the k-th step we have a set S
k
which is a union of 2
k
intervals of length 3
k
. We then remove
the middle third of each interval to get S
k+1
. The denition of S is
S =

k=0
S
k
.
Theorem C.2.1. The Hausdor dimension of the Cantor set is
dim
H
(S) =
log 2
log 3
= log
3
2 .
Proof. Set = log 2/ log 3. We rst prove that dim
H
(S) . For this we must show that
if > then H

(S) = 0. Pick k 0 and let I


1
, . . . , I
2
k be the intervals of length 3
k
that
comprise S
k
, the set at the k-th level of the construction of the Cantor set. Since S S
k
,
this is a cover of S. We compute the -total length of the cover. It is
2
k

j=1
[I
j
[

=
2
k

j=1
3
k
= e
k[log 2 log 3]
,
117
and this approaches zero as k . Note that we have used above that, for example
2
k
= e
k log 2
. Therefore H

(S) = 0 and dim


H
(S) .
For the other direction (to prove dim
H
(S) ) we will show that H

(S) > 0. Let C


n

be a countable cover of S. We will give a bound on the -total length of C


n
. As before,
we may assume that each C
n
is actually a subset of [0, 1]. By compactness one can show the
following:
Lemma C.2.2. Given > 0 there exist nitely many open intervals D
1
, . . . , D
m
such that

n=1
C
n

m
j=1
D
j
and
m

j=1
[D
j
[

<

n=1
[C
n
[

+ .
Proof. The proof is an exercise. The idea is to rst replace the C
n
s by closed intervals and
then slightly widen them, while making them open. Then use compactness.
Now choose k such that
_
1
3
_
k
min[D
j
[ : j = 1, . . . , m .
For l = 1, . . . , k let N
l
be the number of sets D
j
such that 3
l
[D
j
[ < 3
l+1
. Using
= log 2/ log 3 and the denition of k, we nd
m

j=1
[D
j
[

l=1
N
l
3
l
=
k

l=1
N
l
2
l
, (9)
so we will give a lower bound for the right side. Suppose that D
j
has 3
l
[D
j
[ < 3
l+1
.
Then D
j
can intersect at most 2 of the intervals in S
l
, the l-th step in the construction of
the Cantor set. Since each of these intervals produces 2
kl
subintervals at the k-th step of
the construction, we nd that D
j
contains at most 2 2
kl
subintervals at the k-th step of
the construction. But there are only 2
k
subintervals at the k-th step so we nd
2
k

l=1
N
l
2 2
kl
or
1
2

k

l=1
N
l
2
l
.
Combining this with (9),
m

j=1
[D
j
[

1/2 .
Now using the previous lemma with = 1/4,

n=1
[C
n
[

> 1/4 and H

(S) 1/4. Thus


dim
H
(S) .
118
C.3 Exercises
1. Prove Lemma C.1.4.
2. Prove Lemma C.2.2.
3. Prove that if S R with [S[ < has nonempty interior then show that dim
H
(S) = 1.
4. What is the Hausdor dimension of a modied Cantor set where we remove the middle
1/9-th of our intervals?
5. What is the Hausdor dimension of the modied Cantor set from exercise 15, Chapter
3?
119

You might also like