Math 140A - Notes
Neil Donaldson
Fall 2023
Introduction
Analysis is one of the major sub-disciplines of mathematics, being concerned with continuous func-
tions, limits, calculus and accurate approximations.
Analytic ideas date back over 2000 years. For instance, Archimedes (c. 287–212 BC) used limit-type
approaches to approximate the circumference of a circle and compute the area under a parabola.
1
The philosophical objections to such calculations are just as old: how can it make sense to sum up
infinitely many infinitesimally small quantities? This was part of a deeper debate among the ancient
Greeks: is the matter comprising the natural world atomic (consisting of minute, discrete, indivisible
pieces) or continuous (arbitrarily and infinitesimally divisible). Several of Zeno’s famous paradoxes
(5
th
C. BC) grapple with these difficulties: Achilles and the Tortoise essentially argues that the infinite
series
n=1
1
2
n
=
1
2
+
1
4
+
1
8
+
1
16
+ ··· is meaningless.
1
2
1
2
2
1
2
3
1
2
4
···
0 1
1
2
3
4
···
As we’ll see, with modern definitions it makes sense for this sum to evaluate to 1.
The work of Newton and Leibniz in the late 1600s allowed the easy application of calculus to many
important problems in the sciences, but without properly addressing the ancient philosophical chal-
lenges. The logical development of calculus necessitated by this became the triumph of 18
th
–19
th
century mathematics. The critical notions of limit and continuity only became rigorous in the early
1800s, courtesy of Bolzano, Cauchy and Weierstrass (amongst others), with another 50 years before
Riemann provided a thorough description of the definite integral.
The Math 140A/B sequence introduces analysis by focusing on these ideas. In this course we pri-
marily consider sequences, limits, continuity and infinite series. Power series, differentiation and
integration are the focus of 140B. We start, however, with something even more basic: to numeri-
cally measure continuous quantities, we need to familiarize ourselves with the real numbers. Since a
concrete description is quite difficult, we build up to it using first the natural numbers and then the
rationals. . .
1
Archimedes’ circle calculation is reminiscent of the Riemann sum approach to integration, whereas his parabolic area
method required the evaluation of the infinite series
n=0
1
4
n
=
4
3
.
1 The Set N of Natural Numbers
You’ve been using the natural numbers N = {1, 2, 3, 4, 5, . . .} since you learned to count. In mathe-
matics, these must be axiomatically described. Here is one approach, known as Peano’s Axioms.
Axioms 1.1. The natural numbers are a set N satisfying the following properties:
1. (Non-emptiness) N is non-empty.
2. (Successor function) There exists a function f : N N. This function is usually denoted +1’
so that we may write,
n N = n + 1 N
3. (Initial element) f is not surjective. Otherwise said, there exists an element 1 range f which is
not the successor of any element.
2
4. (Unique predecessor/order) f is injective. If m and n have the same successor, then m = n.
5. (Induction) Suppose A N is a subset satisfying
(a) 1 A
(b) n A = n + 1 A.
Then A = N.
Axioms 1–4 are relatively straightforward, the natural numbers are defined by repeatedly adding 1
to the initial element; for instance
3 := f
f (1)
= (1 + 1) + 1
To see why Axiom 5 is so-named, compare an easy example with a standard induction argument.
Example 1.2. Prove that 7
n
4
n
is divisible by 3 for all n N.
Let A be the set of natural numbers for which 7
n
4
n
is divisible by 3. It is required to prove that
A = N.
(a) If n = 1, then 7
1
4
1
= 3, whence 1 A.
(b) Suppose n A. Then 7
n
4
n
= 3λ for some λ N. But then
7
n+1
4
n+1
= 7 ·7
n
4
n+1
= 7(3λ + 4
n
) 4
n+1
= 3 ·7λ + (7 4) ·4
n
= 3
7λ + 4
n
is divisible by 3. It follows that n + 1 A.
Appealing to axiom 5, we see that A = N, hence result.
The two arguments are precisely the familiar base case and induction step.
2
It is purely convention to denote the first natural number by 1; we could use 0, x, α , or any symbol you wish!
2
What about the integers? It should be clear that the integers satisfy axioms 1, 2 and 4, but not 3 and
5. For instance:
3. The function f : Z Z : n 7 n + 1 is surjective (indeed bijective/invertible). The number 1 is
the successor of 0.
We can reverse this observation to provide an explicit construction of the integers from the natural
numbers. Simply extend the function f so that every element has a unique predecessor: 0 is the
unique predecessor of 1, 1 the unique predecessor of 0, etc. In essence we are forcing f (n) = n + 1
to be bijective!
Exercises 1. Most of these exercises are to refresh your memory of mathematical induction. You can
use either the language of Peano’s axiom 5, or the (likely) more familiar base-case/induction-step
formulation.
1. Prove that 1
2
+ 2
2
+ ··· + n
2
=
1
6
n( n + 1)(2n + 1) for all natural numbers n.
2. Prove that 3 + 11 + ··· + (8n 5) = 4n
2
n for all n N.
3. (a) Guess a formula for 1 + 3 + ···+ (2n 1) by evaluating the sum for n = 1, 2, 3, and 4.
(For n = 1 the sum is simply 1)
(b) Prove your formula using mathematical induction.
4. Prove that 11
n
4
n
is divisible by 7 for all n N.
5. The principle of mathematical induction can be extended as follows. A list P
m
, P
m+1
, . . . of
propositions is true provided (i) P
m
is true, (ii) P
n+1
is true whenever P
n
is true and n m.
(a) Prove that n
2
> n + 1 for all integers n 2.
(b) Prove that n! > n
2
for all integers n 4.
(Recall that n! = n(n 1) ···2 ·1)
6. Prove (2n + 1) + (2n + 3) + (2n + 5) + ··· + (4n 1) = 3n
2
for all n N.
7. For each n N, let P
n
denote the assertion n
2
+ 5n + 1 is an even integer”.
(a) Prove that P
n+1
is true whenever P
n
is true.
(b) For which n is P
n
actually true? What is the moral of this exercise?
8. Show that Peano’s induction axiom is false for the set of integers Z by exhibiting a proper subset
A Z which satisfies conditions (a) and (b).
9. Consider Z
3
= {0, 1, 2} under addition modulo 3. That is,
0 + 1 = 1, 1 + 1 = 2, 2 + 1 = 0
Which of Peano’s axioms are satisfied?
3
2 The Set Q of Rational Numbers
There are several ways to define the rational numbers from the integers. For instance, we could
consider the set of relatively prime ordered pairs
Q =
(p, q) : p Z, q N, gcd(p, q) = 1
Z ×N
Things are more familiar once we write
p
q
instead of (p, q) and adopt the convention that
λp
λq
=
p
q
for
any non-zero λ Z. It is easy to define the usual operations (+, ·, etc.) consistently with those for
the integers (Exercise 7).
An alternative approach involves equations. Each linear equation qx p = 0 where p, q Z and
q = 0 corresponds to a rational number! For example
13x + 27 = 0 x =
27
13
Of course 26x + 54 = 0 also corresponds to the same rational number!
Extending this process, we might consider higher degree polynomials.
Definition 2.1. A number x is algebraic if it satisfies an equation of the form
3
a
n
x
n
+ a
n1
x
n1
+ ··· + a
1
x + a
0
= 0 ( )
for some integers a
0
, . . . , a
n
.
Examples 2.2. 1.
2 is algebraic since it satisfies the equation x
2
2 = 0.
2. x =
5
p
7 +
3 is also algebraic:
x
5
7 =
3 = (x
5
7)
2
= 3 = x
10
14x
5
+ 46 = 0
The next result is helpful for deciding whether a given number is rational and can assist with factor-
izing polynomials.
Theorem 2.3 (Rational Roots). Suppose that a
0
, . . . , a
n
Z and that x Q satisfies (). If x =
p
q
in
lowest terms, then p | a
0
and q | a
n
.
Proof. Since x satisfies the polynomial, we see that
a
n
p
q
n
+ ··· + a
1
p
q
+ a
0
= 0 = a
n
p
n
+ a
n1
p
n1
q + ··· + a
1
pq
n1
+ a
0
q
n
= 0
All terms except the last contain a factor of p, whence p | a
0
q
n
. Since gcd(p, q) = 1 it follows that
p | a
0
. The result for q is almost identical.
3
You should be alarmed by this! We have given up on constructing new numbers and instead are simply describing their
properties. No matter, a construction of the real numbers will come later.
4
Examples 2.4. 1. We show that
2 is irrational.
4
Plainly x =
2 satisfies the polynomial equation
x
2
2 = 0. If
2 =
p
q
were rational in lowest terms, then the rational roots theorem forces
p | 2 and q | 1 =
2 1, ±2}
Since none of the numbers ±1, ±2 satisfy x
2
2 = 0, we have a contradiction.
2. (
3 1)
1/3
is irrational. It satisfies (x
3
+ 1)
2
= 3, from which
x
6
+ 2x
3
2 = 0
By the theorem, if x =
p
q
were rational then p | 2 and q | 1, whence x = ±1, ±2, none of which
satisfies (x
3
+ 1)
2
= 3.
3.
4+
3
5
1/2
is irrational. It satisfies 5x
2
4 =
3, from which
25x
4
40x
2
+ 13 = 0
If x =
p
q
were rational, then p | 13 and q | 25. There are twelve possibilities in all:
x = ±1, ±13, ±
1
5
, ±
13
5
, ±
1
25
, ±
13
25
It is tedious to check all cases, but none satisfy the required polynomial.
With this example it is easier to bypass the theorem entirely: if x Q then
3 = 5x
2
4 would
also be rational!
4. We factorize the polynomial 3x
3
+ x
2
+ x 2 = 0. By the rational roots theorem, if x =
p
q
is a
rational root, then p | 2 and q | 3 which gives several possibilities:
x
±1, ±2, ±
1
3
, ±
2
3
It doesn’t take long to try these and observe that x =
2
3
is the only rational solution. The
polynomial has a factor of 3x 2 which we can extract by long division to obtain
3x
3
+ x
2
+ x 2 = (3x 2)(x
2
+ x + 1)
The quadratic has no real roots: absent complex numbers, the factorization is complete.
It is far from clear that there exist non-algebraic (or transcendental) numbers, of which e and π are
the most famous examples. These satisfy no polynomial equation with integer coefficients, though
demonstrating this is tricky.
4
Compare this to the standard proof of the irrationality of
2 as seen in a previous course. Note how easy it is to extend
our approach to
3,
29,
3
2,
5
8, etc.
5
Exercises 2. 1. Describe all the linear equations which correspond to the rational number
101
29
.
2. Show that
3,
5 and
24 are not rational numbers.
(Hint: what are the relevant polynomials?)
3. Show that 2
1/3
and 13
1/4
are not rational numbers.
4. Show that (2 +
2)
1/2
is not rational.
5. Show that (3 +
2)
2/3
is not rational.
6. Explain why 4 7b
2
must be rational if b is rational.
7. Given rational numbers (p, q) and (r, s) as ordered pairs, what are the rational numbers (p, q) +
(r, s) and (p, q) · (r, s)?
(Hint: what is
p
q
+
r
s
?)
8. Let n N. Use the rational roots theorem to prove that
n is rational if and only if it is an
integer.
6
3 Ordered Fields
Thus far, we have formally constructed the natural numbers and used them to (loosely) build the
integers and rational numbers. It is a significantly greater challenge to construct the real numbers.
We start by thinking about ordered fields, of which both Q and R are examples.
Axioms 3.1. A field F is a set together with two binary operations + and ·which satisfy the following
(for all a, b, c F),
5
Addition Multiplication
Closure a + b F ab F
Associativity a + (b + c) = (a + b) + c a( bc) = (ab)c
Commutativity a + b = b + a ab = ba
Identity 0 F such that a + 0 = a 1 F such that a ·1 = a
Inverse a F such that a + (a) = 0 If a = 0, a
1
F such that aa
1
= 1
Distributivity a( b + c) = ab + ac
A field F is ordered if we also have a binary relation which satisfies (again for all a, b, c F):
O1 a b or b a
O2 a b and b a = a = b
O3 a b and b c = a c
O4 a b = a + c b + c
O5 a b and 0 c = ac bc
For an ordered field, the symbol < is used in the usual way: x < y x y and x = y.
As with Peano’s axioms for the natural numbers, these are not worth memorizing. Instead you
should quickly check that you believe all of them for your current understanding of the real numbers;
you can’t prove anything since the real numbers haven’t yet been defined!
Example 3.2. It is worth considering the rational numbers in a little more detail. Recall (Section 2)
how Q may be defined as a set of ordered pairs
p
q
(p, q) Z ×N. It moreover inherits a natural
ordering from Z and N:
p
q
r
s
ps qr (remember that q, s > 0)
5
We write multiplication · as juxtaposition unless it is helpful for clarity. We also use the common shorthand a
2
= a · a.
If you know some abstract algebra:
The addition axioms say that
F, +
is an abelian group.
The multiplication axioms say that
F \ {0}, ·
is an abelian group.
The distributive axiom describes how addition and multiplication interact.
7
It is now possible, though tedious, to prove that each of the axioms of an ordered field holds for Q,
using only basic facts about multiplication, addition and ordering within the integers. For instance,
O3 Suppose a b and b c. Write a =
p
q
, b =
r
s
and c =
t
u
where all three denominators are
positive. By assumption,
ps qr and ru st = psu qru qst = pu qt
= a =
p
q
t
u
= c
Basic Results about ordered fields
As with the axioms of an ordered field, it is not worth memorizing these.
Theorem 3.3. Let F be a ordered field with at least two elements 0 = 1. Then:
1. a + c = b + c = a = b 2. a ·0 = 0
3. (a) b = (ab) 4. (a) (b) = ab
5. ac = bc and c = 0 = a = b 6. ab = 0 = a = 0 or b = 0
7. a b = b a 8. a b and c 0 = bc ac
9. 0 a and 0 b = 0 ab 10. 0 a
2
11. 0 < 1 12. 0 < a = 0 < a
1
13. 0 < a < b = 0 < b
1
< a
1
All of these statements should be intuitive for the fields Q and R. Try proving a few using only the
axioms; they are easiest done in the order presented. For instance, part 2 might be proved as follows:
a ·0 + 0 = a ·0 = a ·(0 + 0) = a ·0 + a ·0 (additive identity/distibutive axioms)
= 0 = a ·0 (part 1)
We finish with one final useful ingredient.
Definition 3.4. If F is an ordered field, then the absolute value of a F is
|
a
|
:=
(
a if a 0
a if a < 0
Theorem 3.5. In any ordered field:
1.
|
a
|
0
2.
|
ab
|
=
|
a
|
·
|
b
|
3.
|
a + b
|
|
a
|
+
|
b
|
(-inequality)
All three parts are immediate if you consider the ±-cases separately for a, b.
8
Exercises 3. 1. Which of the axioms of an ordered field fail for N? For Z?
2. Prove parts 11 and 13 of Theorem 3.3.
(Remember you can use any of the parts that come before. . . )
3. (a) Prove that
|
a + b + c
|
|
a
|
+
|
b
|
+
|
c
|
for all a, b, c R.
(Hint: Apply the triangle inequality twice. Don’t consider eight separate cases!)
(b) Use induction to prove
|
a
1
+ a
2
+ ··· + a
n
|
|
a
1
|
+
|
a
2
|
+ ··· +
|
a
n
|
for any a
1
, . . . , a
n
R.
4. (a) Show that
|
b
|
< a a < b < a.
(b) Show that
|
a b
|
< c b c < a < b + c
(c) Show that
|
a b
|
c b c a b + c
5. Let a, b R. Show that if a b
1
for every b
1
> b, then a b.
(Hint: draw a picture if you’re stuck. This is a very important example!)
6. Following Example 3.2, prove that Q satisfies axiom O5.
(Hint: if a =
p
q
, etc., what is meant by ac bc?)
7. (Hard!) The complex numbers C = {x + iy : x, y R} form a field. Consider the lexicographic
ordering of C defined by
x + iy p + iq
(
x < p or
x = p and y q
Which of the order axioms O1–O5 are satisfied by the lexicographic ordering?
(Don’t prove your claims if an axiom is satisfied, but provide a counter-example if not)
9
4 The Completeness Axiom
While we still haven’t provided an explicit definition of the real numbers, you should be comfortable
with the fact that both Q and R are ordered fields. The question remains of how to distinguish them?
Perhaps surprisingly, only one additional axiom is required: the completeness axiom or least upper
bound principle. To explain this we first need some terminology.
Definition 4.1 (Maxima, Minima & Boundedness). Let S R be non-empty.
1. S is bounded above if it has an upper bound M:
M R such that s S, s M
2. We write M = max S, the maximum of S, if M is an upper bound for S and M S.
3. S bounded below, a lower bound m, and the minimum min S are defined similarly.
4. S is bounded if it is bounded above and below. We say that S is bounded by M if
s S,
|
s
|
M
Examples 4.2. 1. If S is a finite set, then it is bounded and has both a maximum and a minimum.
For instance, S = {3, π, 12} has min S = 3 and max S = 12.
2. N has minimum 1, but no maximum. Z and Q have neither: both are unbounded.
3. The interval S = [0, 3) = {x R : 0 x < 3} is bounded, for example by M = 5, it has
minimum 0 and no maximum. While this last is likely intuitive, it worth giving an explicit
argument, in this case by contradiction.
Suppose x = max S exists. It is helpful to draw a picture to get the lay of the land. Since x S,
we’ve placed x inside the interval, away from 3.
0 3x
s =
x+3
2
The crux of the proof is to observe that there exists s S which is larger than x. The natural
choice is the average s :=
1
2
(x + 3). Now observe that
3 s = s x =
1
2
(3 x) > 0
In particular,
s S since it is non-negative and s < 3.
s > x.
Since S contains an element larger than x, it follows that x cannot be the maximum of S. In
conclusion, S has no maximum.
10
The following should be immediate: try proving them yourself.
Lemma 4.3. 1. If M is an upper bound for S, so is M + ε for any ε 0.
2. If max S exists, then it is unique.
3. A set is bounded if and only if it is bounded above and below. In particular, if m, M are
lower/upper bounds, then S is bounded by
s S,
|
s
|
max(
|
m
|
,
|
M
|
)
Example 4.4. Before introducing the key axiom, we consider a variation on the previous example.
We show that the following set has no maximum:
S = Q [0,
2) = {x Q : 0 x <
2}
The approach is similar to before: given a hypothetical maximum x, find an element s S between
x and
2. The challenge is that we can’t simply use the average
1
2
(x +
2) : this isn’t rational (why?)
and so doesn’t lie in S!
To fix this, we informally invoke sequences: this might seem quite hard at the moment, but will be
made rigorous later. The rough idea is to construct a sequence (s
n
) of elements of S which increases
to
2. Eventually one of these must be larger than x.
Define a sequence of rational numbers (s
n
) by s
n
=
1
10
n
10
n
2, where denotes the floor function.
6
The sequence simply recovers the first n decimal places of
2:
s
0
= 1, s
1
= 1.4 =
14
10
, s
2
= 1.41 =
141
100
, s
3
= 1.414 =
1414
1000
, . . .
and has the following properties:
s
n
S since any truncating decimal is rational and certainly 0 s
n
<
2.
2 s
n
< 10
n
follows since 10
n
2 10
n
2 < 1.
Now suppose x = max S exists. Since x S, we have x <
2. Choose N N large enough so that
10
N
<
2 x. Then s
N
S and
2 s
N
< 10
N
<
2 x = x < s
N
The hypothetical maximum x is not an upper bound for S: contradiction.
0 1 x
2
s
0
s
N1
s
N
10
N
6
y is the greatest integer less than or equal to y; informally round down. For example π = 3. This approach is a
something of a hack: it can be sped up enormously using the upcoming density of Q in R (Corollary 4.12); indeed the
Archimedean property on which it depends is necessary for y to be well-defined.
11
Suprema and Infima
We now generalize the idea of maximum and minimum values for bounded sets.
Example 4.5. The interval [2, 5) has least upper bound 5: among all upper bounds, 5 is the smallest.
Definition 4.6. Let S R be non-empty.
1. If S is bounded above, its supremum sup S is its least upper bound. Otherwise said,
(a) sup S is an upper bound: s S, s sup S,
(b) sup S is the least such: if M is an upper bound, then sup S M.
2. If S is bounded below, its infimum inf S is its greatest lower bound. Equivalently,
(a) inf S is a lower bound: s S, inf S s,
(b) inf S is the greatest such: if m is a lower bound, then m inf S.
sup Sinf S sm M
S
upper bounds
lower bounds
Example (4.5 cont). We verify the supremum and infimum for S = [2, 5); parts (a), (b) are the
properties in the above definition.
(a) Since s S 2 s < 5, we see that 5 is an upper bound and 2 a lower bound.
(b) Given x < 5, define
7
s := max{
1
2
(x + 5), 4}. Observe that x < s < 5 from which s S is larger
than x. It follows that x is not an upper bound for S, and that 5 is the least such.
Similarly, if y > 2, define t := min{
1
2
( y + 2), 4} to see that t S is smaller than y, which cannot
therefore be a lower bound for S.
We conclude that sup S = 5 and inf S = 2.
2 3 4 5
S
sup S
inf S
y
x
t s
We are assuming something quite important here!
Axiom 4.7 (Completeness of R). If S R is non-empty and bounded above, then sup S exists (and
is a real number!).
It is this property that distinguishes the real numbers from the rationals.
8
Note that every bounded
set S of rational numbers has a supremum; the issue is that sup S might not be rational!
7
The number 4 is merely an arbitrary element to make sure s S in case x were huge and negative!
8
If you’ve studied abstract algebra, then a more rigorous statement should make sense: every ordered field with 0 = 1
and which satisfies the completeness axiom is isomorphic to the real numbers.
12
Example (4.4 cont). The set S = Q [0,
2) has sup S =
2. We check conditions (a), (b) in
Definition 4.6.
(a) Certainly
2 is an upper bound for S, since every element is less than
2.
(b) If x <
2 is given, then our previous argument says there exists some s
N
S for which s
N
> x.
Plainly x isn’t an upper bound.
In conclusion,
2 is the smallest upper bound for S.
Consider the contrapositive of part (b) of Definition 4.6 after replacing M with x.
If x < sup S, then x is not an upper bound for S.
If we unpack this further, we recover a useful existence result. Indeed this is precisely what we did
in both previous examples.
Lemma 4.8. 1. If x < sup S, then s S such that s > x.
2. If y > inf S, then t S such that t < y.
sup S
inf S
st
y
x
S
This observation will be used repeatedly, so make sure it is well understood.
Examples 4.9. We state the following without proof or calculation. You should be able to justify all
these statements using the definition, or by mirroring the above examples.
1. A bounded set has many possible bounds, but only one supremum or infimum.
2. If S has a maximum, then max S = sup S. Similarly min S = inf S if a minimum exists.
3. S = Q (π, 4) has sup S = 4 and inf S = π.
4. S = {
1
n
: n N} = {. . . ,
1
4
,
1
3
,
1
2
, 1} has sup S = max S = 1, inf S = 0, and no minimum.
5. S =
S
n=1
[n, n +
1
2
) = [1, 1.5) [2, 2.5) [3, 3.5) ··· has inf S = 1. It is not bounded above.
6. S =
T
n=1
[
1
n
, 1 +
1
n
) has inf S = 1 = sup S since S = {1}.
The completeness axiom only asserts the existence of the supremum of a bounded set. By reflecting
across zero (see Exercise 9), we obtain the same thing for the infimum.
Theorem 4.10 (Existence of Infima). If S R non-empty and bounded below, then inf S R exists.
13
The Archimedean Property and the Density of the Rationals
We finish this section by discussing the distribution of the rational numbers among the real numbers.
Theorem 4.11 (Archimedean Property). If b > 0 is a real number, then n N such that n > b.
Equivalently:
9
a, b > 0 = n N such that an > b.
In this result we assume nothing about R except that is an ordered field satisfying the completeness
axiom and 0 = 1. The natural numbers in this context are defined as the subset
N = {1, 1 + 1, 1 + 1 + 1, . . .} R
Proof. Suppose the result were false. Then b > 0 such that n b for all n N; that is, N is bounded
above! By completeness, sup N exists, and we trivially see that
0 < 1 = sup N < sup N + 1 = sup N 1 < sup N
By Lemma 4.8, n N such that n > sup N 1. But then sup N < n + 1 which is clearly a natural
number! Thus sup N is not an upper bound for N: contradiction.
The use of completeness is necessary: there exist non-Archimedean ordered fields!
Corollary 4.12 (Density of Q in R). Between any two real numbers, there exists a rational number.
The idea is simple: given a < b, stretch the interval by an integer factor n until it contains an integer
m, before dividing by n to obtain a <
m
n
< b. The Archimedean property shows the existence of m, n.
Proof. WLOG suppose 0 a < b. The Archimedean property applied to
1
ba
> 0 says
n N such that n >
1
ba
A second application says k N such that k > an. Now consider
J := {j N : an < j k}
and define m = min J: this exists since J is a finite non-empty set of natural numbers.
10
0
a b an bn
k
m
m
n
J
Clearly m > an > m 1, since m = min J. But then m an + 1 < bn. We conclude that
an < m < bn = a <
m
n
< b
It is immediate that any interval (a, b) now contains infinitely many rational numbers.
9
Just replace b with
b
a
.
10
This part of the argument is needed because, in this context, we haven’t established the well-ordering property of N
(equivalent to Peano’s fifth axiom).
14
Exercises 4. 1. Decide if each set is bounded above and/or below. If it is, state its supremum and/or
infimum (no working is required).
(a) (0, 1) (b) {2, 7} (c) {0}
(d)
S
n=1
[2n, 2n + 1] (e)
1
1
3
n
: n N
(f) {r Q : r
2
< 2}
(g)
S
n=1
1
1
n
, 1 +
1
n
(h) {
1
n
: n N and n is prime} (i) {cos(
nπ
3
) : n N}
2. Modelling Example 4.4, sketch an argument that S = Q (π, 4] has no minimum.
(Hint: let s
n
be π rounded up to n decimal places)
3. Let S be a non-empty, bounded subset of R.
(a) Prove that inf S sup S.
(b) What can you say about S if inf S = sup S?
4. Let S and T be non-empty subsets of R with the property that s t for all s S and t T.
(a) Prove that S is bounded above and T bounded below.
(b) Prove that sup S inf T.
(c) Give an example of such sets S, T where S T is non-empty.
(d) Give an example of such sets S, T where S T is empty, and sup S = inf T.
5. Prove that if a > 0 then there exists n N such that
1
n
< a < n.
6. Let I = R \Q be the set of irrational numbers. Given real numbers a < b, prove that there exists
x I such that a < x < b.
(Hint: First show {r +
2 : r Q} I)
7. Let A, B be non-empty bounded subsets of R, and let S be the set of all sums
S := {a + b : a A, b B}
(a) Prove that sup S = sup A + sup B.
(b) Prove that inf S = inf A + inf B.
8. Show that sup{r Q : r < a} = a for each a R.
9. We prove Theorem 4.10 on the existence of the infimum.
Let S R be non-empty and let m be a lower bound for S. Define T = {t R : t S} by
reflecting S across zero.
0
S
T
inf S = sup T
m
sup T
m ss
(a) Prove that m is an upper bound for T.
(b) By completeness (Axiom 4.7), sup T exists. Prove that inf S = sup T by verifying Defini-
tion 4.6 parts 2(a) and (b).
15
5 The Symbols ±
Thus far the only subsets of the real numbers that have a supremum are those which are non-empty
and bounded above. In this very short section, we introduce the -symbol to provide all subsets of the
real numbers with both a supremum and an infimum.
Definition 5.1. Let S R be any subset. If S is bounded above/below, then sup S/inf S are as in
Definition 4.6. Otherwise:
1. We write sup S = if S is unbounded above, that is
x R, s S such that s > x
2. We write inf S = if S is unbounded below,
y R, t S such that t < y
3. By convention, sup := and inf := , though these will rarely be of use to us.
The symbols ± have no other meaning (as yet): in particular, they are not numbers! If one is willing
to abuse notation and write x < and y > for any real numbers x, y, then the conclusions of
Lemma 4.8 are precisely statements 1 & 2 in the above definition!
Examples 5.2. 1. sup R = sup Q = sup Z = sup N = , since all are unbounded above. We also
have inf R = inf Q = inf Z = (recall that inf N = min N = 1).
2. If a < b, then any interval [a, b], (a, b), [a, b) or (a, b] has supremum b and infimum a, even if one
end is infinite. For example,
S = (7, ) = {x R : x > 7}
has sup S = and inf S = 7.
3. Let S = {x R : x
3
4x < 0}. With a little factorization, we see that
x
3
4x = x(x 2)(x + 2) < 0 x < 2 or 0 < x < 2
It follows that S = (, 2) (0, 2), from which sup S = 2 and inf S = .
Exercises 5. 1. Give the infimum and supremum of each of the following sets:
(a) {x R : x < 0} (b) {x R : x
3
8}
(c) {x
2
: x R} (d) {x R : x
2
< 8}
2. Let S R be non-empty, and let S = {s : s S}. Prove that inf S = sup(S).
3. Let S, T R be non-empty such that S T. Prove that inf T inf S sup S sup T.
4. If sup S < inf S, what can you say about S?
16
6 A Development of R (non-examinable)
The comment in footnote 8 essentially constitutes a synthetic definition of the real numbers: there
is essentially just one set with the required properties. It is nice, however, to be able to provide an
explicit construction. The following approach uses Dedekind cuts.
First one defines N, Z and Q. Use Peano’s axioms and proceed as in sections 1 and 2. The operations
+, · and are defined, first on N and then for Z and Q building on the concepts for the integers.
Definition 6.1. A Dedekind cut α
is a non-empty proper subset of Q with the following properties:
1. If r α
and s Q with s < r, then s α
.
2. α
has no maximum.
Define R to be the set of all Dedekind cuts!
The rough idea is that a real number α corresponds to the Dedekind cut α
of all rational numbers
less than α. While this is the idea, it doesn’t stand up as a definition due to circular logic: α cannot be
defined in terms of itself!
Examples 6.2. 1. For any rational number r, the corresponding real number is the Dedekind cut
r
= {x Q : x < r}
For instance 4
= {x Q : x < 4} is the Dedekind cut definition of the real number 4.
2. It is a little trickier to explicitly define Dedekind cuts corresponding to irrational numbers,
though some are relatively straightforward. For instance the real number
2 would be the
Dedekind cut
2
= {x Q : x < 0 or x
2
< 2}
It remains to prove that the set of Dedekind cuts satisfies all the axioms of a complete ordered field.
The full details are too much for us, so here is a rough overview.
Define the ordering of Dedekind cuts via
α
β
α
β
One can now prove axioms O1–O3 and that the ordering corresponds to that of Q.
Define addition of cuts via
α
+ β
:= {a + b : a α
, b β
}
This suffices to prove the addition axioms and O4: a careful definition of α
is required.
Multiplication is horrible: if α
, β
0 then
α
β
:= {ab : a 0, a α
, b 0, b β
}{q Q : q < 0}
which may be carefully extended to cover situations when α
or β < 0. Once can then prove
the multiplication axioms, the final order axiom O5, and the distributive axiom.
17
The completeness axiom must also be verified, though it comes almost for free! If A R (so
that A is a set of Dedekind cuts), then the supremum of A is
sup A =
[
α
A
α
An alternative approach to R using sequences of rational numbers will be given later in the course.
Exercises 6. 1. Show that if α
, β
are Dedekind cuts, then so is
α
+ β
= {r
1
+ r
2
: r
1
α
, r
2
β
}
2. Let α
, β
be Dedekind cuts and define the ‘product’:
α
· β
= {r
1
r
2
: r
1
α
, r
2
β
}
(a) Calculate some ‘products’ using the cuts 0
, 1
and (1)
.
(b) Discuss why this definition of ‘product’ is unsatisfactory for defining multiplication in R.
3. We verify the Archimedean property (Theorem 4.11) using the Dedekind cut definition of R (it
is somewhat easier since the unboundedness of N and Q are baked in).
(a) Explain why every cut β
is bounded above by some rational number.
(Hint: if β
satisfies Definition 6.1 parts 1 & 2 but is unbounded above, then what is it?)
(b) If β
> 0
is a positive cut bounded above by
p
q
with p, q N, show that n := p + 1
corresponds to a cut for which n
> β
.
18
7 Limits of Sequences
Sequences are the fundamental tool in our approach to analysis.
Definition 7.1. A sequence of real numbers is a list indexed by the natural numbers
( s
n
) = (s
1
, s
2
, s
3
, . . .)
We call s
1
the initial term/element.
This is strictly the definition of an infinite sequence; finite sequences don’t appear in this course. Other
letters may be used (a
n
, b
n
, etc.), though s
n
is most common in the abstract. It is also common to have
sequences which start with a different initial term (n = 0 is particularly common). If you need to be
explicit, describe the range of indices with sub/superscripts, e.g. (s
n
)
n=0
.
Examples 7.2. 1. Explicit sequences are often defined by providing a formula for the n
th
term. For
instance, s
n
=
1 +
1
n
n
defines a sequence whose first three terms are
s
1
= 2, s
2
=
9
4
, s
3
=
64
27
, . . .
Since each term is a rational number, (s
n
) could be described as a rational sequence.
2. Sequences can be defined inductively. For instance t
1
= 1 and t
n+1
= 3t
n
1 together define
the sequence
( t
n
) = (1, 2, 5, 14, 41, . . .)
3. u
n
=
1
n
2
4
defines a sequence with initial term u
3
=
1
5
:
( u
n
)
n=3
=
1
5
,
1
12
,
1
21
, . . .
Limits In analysis we are typically interested in what happens to the terms of a sequence (s
n
) when
n gets large (as such, it is common to be non-explicit as to the initial term). In elementary calculus,
you should have become used to writing expressions such as
11
lim
2n
2
+ 3n 1
3n
2
2
=
2
3
which encapsulates the idea that the expression s
n
=
2n
2
+3n1
3n
2
2
gets close to
2
3
when n is large. We can
easily convince ourselves of this with a calculator/computer: to 4 decimal places, we have
( s
n
) = (4, 1.3, 1.04, 0.9348, 0.8767, 0.8396, 0.8138, 0.7947, . . .), s
1000
= 0.6677
Our primary business is to make this idea logically watertight. In the next section we will do so
by developing the formal definition of limit. Before seeing this, we quickly refresh a few simple
examples from elementary calculus. At the moment, all these rely on your intuition and experience.
This is a good thing to practice: in analysis it is often essential to have a good idea of the correct
answer before you try to prove it!
11
If there are multiple letters in your expression, then for clarity it can be helpful to write lim
n
with a subscript.
19
Examples 7.3. 1. lim
1
n
= 0. Our instinct is s
n
=
1
n
becomes arbitrarily small as n becomes large.
2. lim
7n+9
2n4
=
7
2
. To convince yourself of this, you might write
7n+9
2n4
=
7+
9
n
2
4
n
and observe that the
1
n
terms become tiny as n increases.
3. The sequence with n
th
term s
n
= (1)
n
does not converge to anything (it diverges). Indeed
( s
n
)
n=0
= (1, 1, 1, 1, 1, 1, . . .)
isn’t getting closer to any real number.
4. If c
n
=
1
n
cos
πn
6
, then lim c
n
= 0. To see this, observe that the cosine term lies between ±1,
while
1
n
has limit 0.
5. The sequence defined inductively by s
0
= 2, s
n+1
:=
1
2
s
n
+ 3 begins
( s
n
) =
2, 4, 5,
11
2
,
23
4
,
47
8
, . . .
This appears to have limit lim s
n
= 6. Indeed it is not hard to spot the pattern s
n
= 6
4
2
n
which
is easily verified by induction: for the induction step, simply observe that
1
2
s
n
+ 3 =
1
2
6
4
2
n
+ 3 = 6
4
2
n+1
Exercises 7. 1. Decide whether each sequence converges; if it does, give the limit. No proofs are
required; if you’re unsure what’s going on, try writing out the first few terms.
(a) a
n
=
1
3n + 1
(b) b
n
=
3n + 1
4n 1
(c) c
n
=
n
3
n
(d) d
n
= sin
nπ
4
2. Repeat the previous question for sequences whose n
th
term is as follows:
(a)
n
2
+ 3
n
2
3
(b) 1 +
2
n
(c) 2
1/n
(d) (1)
n
n (e)
7n
3
+ 8n
2n
3
31
(f) sin
nπ
2
(g) sin
2nπ
3
(h)
2
n+1
+ 5
2
n
7
(i)
1 +
1
n
2
(j)
6n + 4
9n
2
+ 7
3. Give an example of:
(a) A sequence (x
n
) of irrational numbers having a limit lim x
n
that is a rational number.
(b) A sequence (r
n
) of rational numbers having a limit lim r
n
that is an irrational number.
4. Prove by induction that the sequence defined in Example 7.2.2 has n
th
term t
n
=
1
2
3
n1
+ 1
.
5. In future courses, you’ll meet sequences of functions. For instance, we could define a sequence
( f
n
) of functions f
n
: R R inductively via
f
0
(x) 1, f
n+1
(x) := 1 +
Z
x
0
f
n
( t) dt
Compute the functions f
1
, f
2
and f
3
. The sequence ( f
n
) should seem familiar if you think back
to elementary calculus; why?
20
8 The Formal Definition of Limit
Definition 8.1. A sequence (s
n
) converges to a limit s R, if
12
ϵ > 0, N such that n > N =
|
s
n
s
|
< ϵ
We write lim s
n
= s or simply s
n
s; both are read s
n
approaches (or tends to) s.”
A sequence converges if it has a limit, and diverges otherwise.
This isn’t as hard as it looks! The best way to understand it is to work a lot of examples. . .
Example 8.2. We show that the sequence with n
th
term s
n
= 2
1
n
converges to s = 2.
If we plot the sequence like a function, we see how ϵ controls the distance from s
n
is to the limit s; the
definition requires us to show that no matter how small we make ϵ, there is some tail of the sequence
(all s
n
with n > N) whose terms are less than a distance ϵ from the limit.
0
1
s
n
0 20 40 60 80 100
s + ϵ
s ϵ
N
s = 2
n
. . . guarantees that s
n
lives here
n being larger than this. . .
To verify a ‘for all, there exists’ statement requires an argument with a specific structure:
Suppose ϵ > 0 has been provided and describe N, dependent on ϵ (ϵ smaller means N larger).
Verify algebraically that n > N =
|
s
n
s
|
< ϵ.
Scratch work. To find a suitable N, start with what you want to be true and let it inspire you.
We want
|
s
n
s
|
=
2
1
n
2
=
1
n
< ϵ (equivalently n >
1
ϵ
2
) whenever n > N.
Choosing N =
1
ϵ
2
should be enough to complete the proof!
Warning! We do not yet have a proof: N =
1
ϵ
2
is not the correct conclusion! We finish by rear-
ranging our scratch work to make it clear that the definition is satisfied.
Formal argument. Suppose ϵ > 0 is given, and let N =
1
ϵ
2
. Then
n > N =
|
s
n
s
|
=
2
1
n
2
=
1
n
<
1
N
= ϵ
Thus s
n
2, as required.
12
N can be quantified as either a real or a natural number, the definitions being equivalent by the Archimedean property:
if N R satisfies the definition, then
e
N N such that
e
N N; but then n >
e
N = n > N . . . It tends to be easier to use
R for convergence and N when directly proving divergence (see Definition 8.5).
21
The last three lines are all we need—think of them as the concert performance after much rehearsal!
With practice, you might be able to do simple ϵN arguments like these without scratch work, though
even experts usually require some.
Before seeing more examples, we prove a hopefully intuitive result.
Lemma 8.3 (Uniqueness of Limit). If (s
n
) converges, then its limit is unique.
The proof structure should be familiar from other uniqueness arguments: assume there are two limits
s = t and obtain a contradiction. The picture explains the strategy: by choosing ϵ =
|
st
|
2
in the defini-
tion we obtain a tail of the sequence (all terms s
n
coming after some N) which must be simultaneously
close to both limits.
st
s + et e
s+t
2
For all n > N, s
n
must lie both here and here!
Proof. Suppose s = t are two limits. Take ϵ =
|
st
|
2
and apply Definition 8.1 twice: N
1
, N
2
such that
n > N
1
=
|
s
n
s
|
<
|
s t
|
2
and n > N
2
=
|
s
n
t
|
<
|
s t
|
2
Define N := max{N
1
, N
2
}. Then,
n > N =
|
s t
|
=
|
s s
n
+ s
n
t
|
|
s
n
s
|
+
|
s
n
t
|
(-inequality)
<
|
s t
|
2
+
|
s t
|
2
=
|
s t
|
Contradiction.
Examples 8.4. We give several more examples of using the limit definition. Remember that only the
formal arguments needs to be presented; some scratch work is included to show the thought process.
1. For any k R
+
, we prove that lim
1
n
k
= 0.
Scratch work. Given ϵ > 0, we want to choose N such that
n > N =
1
n
k
< ϵ
This amounts to having n >
1
ϵ
1/k
, so it is enough to choose N to be the right hand side.
Formal argument. Let ϵ > 0 be given, and let N =
1
ϵ
1/k
. Then
n > N =
1
n
k
0
=
1
n
k
<
1
N
k
= ϵ
We conclude that
1
n
k
0, as required.
22
2. We prove that lim
n + 4
n
= 0.
Scratch work. Everything follows from a (hopefully) familiar algebraic trick for manipulating
surd expressions:
n + 4
n =
4
n + 4 +
n
<
4
2
n
=
2
n
Formal argument. Let ϵ > 0 be given, and let N =
4
ϵ
2
. Then
n > N =
n + 4
n
=
4
n + 4 +
n
<
4
2
n
=
2
n
<
2
N
= ϵ
Thus lim
n + 4
n
= 0, as required.
3. We prove that lim
3n+1
n7
= 3.
Scratch work. Given ϵ > 0, we want to choose N such that
n > N =
3n + 1
n 7
3
=
(3n + 1) 3(n 7)
n 7
=
22
n 7
< ϵ ()
For large n (n > 7) everything is positive, so it is sufficient for us to have
n 7 >
22
ϵ
or equivalently n > 7 +
22
ϵ
Formal argument 1. Let ϵ > 0 be given, and let N = 7 +
22
ϵ
. Then
n > N =
3n + 1
n 7
3
=
22
n 7
<
22
N 7
= ϵ
The absolute values are dropped since n > 7. We conclude that lim
3n+1
n7
= 3.
Scratch work (cont). An alternative approach is available if we play with () a little. By insisting
that n 14, we may simplify the denominator
n 7
1
2
n =
22
n 7
22
1
2
n
=
44
n
Formal argument 2. Let ϵ > 0 be given, and let N = max
14,
44
ϵ
. Then
n > N =
3n + 1
n 7
3
=
22
n 7
22
1
2
n
=
44
n
(since n 14)
<
44
N
ϵ (since N
44
ϵ
)
We again conclude that lim
3n+1
n7
= 3.
23
The plot illustrates the two choices of N as functions of ϵ. Observe how the second is always
larger than the first: if N = N
1
( ϵ) works in a proof, then any larger choice N
2
( ϵ) will also,
n > N
2
N
1
= n > N =
|
s
n
s
|
< ϵ
Use this to your advantage to produce simpler arguments.
0
100
N
0 1 2 3 4 5
N
1
= 7 +
22
ϵ
N
2
= max
14,
44
ϵ
ϵ
4. Given s
n
=
2n
4
3n+1
3n
4
+n
2
+4
, we prove that lim s
n
=
2
3
.
Scratch work. We want to conclude that
n > N =
2n
4
3n + 1
3n
4
+ n
2
+ 4
2
3
=
2n
2
9n 5
3( 3n
4
+ n
2
+ 4)
< ϵ
Attempting to solve for n (as in the first method previously) is crazy! Instead we simplify the
fraction by observing that since n 1, we have
2n
2
9n 5
3( 3n
4
+ n
2
+ 4)
16n
2
3( 3n
4
+ n
2
+ 4)
(1 n n
2
and the -inequality)
<
16n
2
9n
4
<
2
n
2
(n
2
+ 4 > 0 = 3n
4
+ n
2
+ 4 > 3n
4
)
The final simplification is merely for additional tidying.
Formal argument. Let ϵ > 0 be given, and let N =
q
2
ϵ
. Then
n > N =
s
n
2
3
=
2n
2
9n 5
3( 3n
4
+ n
2
+ 4)
<
16n
2
9n
4
(since n 1)
<
2
n
2
<
2
N
2
= ϵ
Other choices of N are feasible (see e.g. Exercise 5); everything depends on how you want to
simplify things in your scratch work.
24
Divergent sequences
By negating Definition 8.1, we obtain a new definition.
Definition 8.5. A sequence (s
n
) does not converge to s R if,
ϵ > 0 such that N, n > N with
|
s
n
s
|
ϵ
A sequence is divergent if it does not converge to any limit s R. Otherwise said,
s R, ϵ > 0 such that N, n > N with
|
s
n
s
|
ϵ
Examples 8.6. 1. We prove that the sequence with s
n
=
7
n
does not converge to s = 2.
Visualization. We intuitively know that s
n
0. If ϵ is anything smaller than 2, then the terms
s
n
will eventually be further than this from s = 2.
0
4
6
s
n
0 5 10 15
s = 2
s + ϵ
s ϵ
n
Every tail of the sequence contains
elements s
n
that do not lie here
Direct argument. Let ϵ = 1. Since we are only concerned with large values of n, we see that
|
s
n
s
|
=
7
n
2
= 2
7
n
ϵ = 1
7
n
1 n 7
Given N N, let
13
n = max{7, N + 1}. But then
|
s
n
s
|
=
7
n
2
ϵ, from which we
conclude that s
n
2.
Contradiction argument. For an alternative approach, we suppose s
n
2 and let ϵ = 1 in
Definition 8.1. Then N such that
n > N =
7
n
2
< 1 = 1 <
7
n
< 3 =
7
3
< n < 7
Regardless of the value of N, this cannot hold for all n > N: in particular n := max{7, N + 1}.
Contradiction.
The two arguments are very similar, though consider that a significant advantage of the contra-
diction approach is that you only have to remember one definition!
13
This is why we prefer to let N be a natural number when proving divergence. If N R, then we’d have to use the
ceiling function (n = max{14, N + 1}), or resort to the Archimedean property on which it depends (n > max{14, N}).
Either way is ugly and potentially confusing, so better avoided.
25
2. We prove that the sequence defined by s
n
= (1)
n
is divergent.
Suppose, for contradiction, that s
n
s. The picture shows the case s 0 and strongly suggests
that ϵ = 1 will lead to a contradiction (why?).
1
1
2
s
n
10 20
s + ϵ
s ϵ
N
s
n
Regardless of N, every tail
of the sequence after here. . .
. . . contains some elements
s
n
that do not lie here
Let ϵ = 1 in the definition of limit. Then N N such that
n > N =
|
( 1)
n
s
|
< 1
One each of the values {n
e
, n
o
} = {N, N + 1} is even and the other odd. There are two cases:
If s 0 then
|
( 1)
n
o
s
|
=
|
1 s
|
= s + 1 1 = ϵ.
If s < 0 then
|
( 1)
n
e
s
|
=
|
1 s
|
= 1 s 1 = ϵ.
Either way we have a contradiction. We conclude that (s
n
) is divergent.
3. We show that the sequence defined by s
n
= ln n is divergent.
14
Our intuition from calculus is that logarithms increase unboundedly. For any s R, letting
ϵ = 1 should be enough, for eventually ln n s + 1. This time we prove directly using the
definition of divergence (8.5).
Suppose s R, let ϵ = 1, and suppose that N N is given. Define n = max{N + 1, e
s+1
} and
observe that. Then
n > N and ln n ln(e
s+1
) = s + 1 (ln is increasing, and so respects inequalities!)
In particular,
|
s
n
s
|
= ln n s 1 = ϵ
We conclude that (s
n
) is divergent.
14
In the next section we’ll have a definition of what it means for a sequence to diverge to : this is what’s happening for
s
n
= ln n, but it’s not (yet) what we’re trying to demonstrate.
26
A Little Abstraction
Working explicitly with the limit definition is tedious. In the next section we’ll develop and summa-
rize the limit laws so we can combine limits of sequences without providing new ϵ-N proofs. To start
working towards this, here are three general results.
Lemma 8.7. If lim s
n
= s, then lim s
2
n
= s
2
.
The challenge is that we want to bound
s
2
n
s
2
=
|
s
n
s
||
s
n
+ s
|
, which means we need some
control over
|
s
n
+ s
|
. One way uses the triangle-inequality,
|
s
n
+ s
|
=
|
s
n
s + 2s
|
|
s
n
s
|
+ 2
|
s
|
Assuming
|
s
n
s
|
1 gives us a fixed bound
|
s + s
n
|
1 + 2
|
s
|
. We may now begin a proof.
Proof. Suppose s
n
s. Let ϵ > 0 be given, and let δ = min{1,
ϵ
1+2
|
s
|
}. Since s
n
s, N such that
n > N =
|
s
n
s
|
< δ
But then
n > N =
s
2
n
s
2
=
|
s
n
s
||
s
n
+ s
|
|
s
n
s
|
(
|
s
n
s
|
+ 2
|
s
|
)
(-inequality)
< δ(1 + 2
|
s
|
) (since
|
s
n
s
|
< δ 1)
ϵ
Theorem 8.8. Suppose lim s
n
= s.
1. If s
n
m for all except finitely many n, then s m.
2. If s
n
M for all except finitely many n, then s M.
Proof. We prove part 1 by contradiction—part 2 is similar.
Suppose s
n
s < m and let ϵ =
ms
2
> 0. Then N such that
n > N =
|
s
n
s
|
<
m s
2
= s
n
s <
m s
2
= s
n
m <
s m
2
< 0 (add s m to both sides)
By assumption, s
n
< m holds for at most finitely many n. Contradiction.
The expression for all but finitely many n can be added to many abstract limit theorems; other com-
mon variants are for all large n, and for some tail of the sequence. To avoid cumbersome language, the
expression is often omitted. Remember that convergence/divergence is concerned with what hap-
pens when n is large: we can change or delete the first million terms of (s
n
) without altering it’s
convergence status!
27
Theorem 8.9 (Squeeze Theorem). Suppose three sequences satisfy a
n
s
n
b
n
(for all large n) and
that (a
n
) and (b
n
) both converge to s. Then lim s
n
= s.
Proof. By subtracting s from our assumed inequality, we see that
a
n
s s
n
s b
n
s =
|
s
n
s
|
max{
|
a
n
s
|
,
|
b
n
s
|
}
It remains to bound the right hand side by ϵ. Let ϵ > 0 be given, then there exist N
a
, N
b
such that
n > N
a
=
|
a
n
s
|
< ϵ and n > N
b
=
|
b
n
s
|
< ϵ
Finally let N = max{N
a
, N
b
} to see that
n > N =
|
s
n
s
|
max{
|
a
n
s
|
,
|
b
n
s
|
} < ϵ
Example 8.10. If s
n
=
1+sin n
n
, then 0 s
n
2
n
. The squeeze theorem quickly forces lim s
n
= 0.
Exercises 8. 1. For each sequence, determine the limit and prove your claim.
(a) a
n
=
n
n
2
+1
(b) b
n
=
7n19
3n+7
(c) c
n
=
4n+3
7n5
(d) d
n
=
2n+4
5n+2
(e) e
n
=
1
n
sin n (f) f
n
=
n
2
+n1
3n
2
10
2. Let (t
n
) be a bounded sequence (there exists M such that
|
t
n
|
M for all n), and let (s
n
) be a
sequence such that lim s
n
= 0. Prove that lim(s
n
t
n
) = 0.
(Hint: given ϵ, note that
ϵ
|
M
|
is also a small number. . . )
3. Prove the following
(a) lim[
n
2
+ 1 n] = 0 (b) lim[
n
2
+ n n] =
1
2
(c) lim[
4n
2
+ n 2n] =
1
4
4. Let (s
n
) be a convergent sequence, and suppose lim s
n
> a. Prove that there exists N such that
n > N = s
n
> a.
5. (a) Show that n 2 = 2n
2
+ 9n + 5 9n
2
.
(b) (Recall Example 8.4.4) Provide another argument that lim
2n
4
3n+1
3n
4
+n
2
+4
=
2
3
by choosing N =
max{2,
1
ϵ
}.
6. (a) Prove that the sequence with n
th
term s
n
=
2
n
2
does not converge to 1.
(b) Prove that ( s
n
) does not converge to 1.
7. Prove that the sequence defined by t
n
= n
2
diverges.
8. Provide a contradiction argument to justify Example 8.6.3: (ln n) diverges.
9. (Recall Theorem 8.8) Suppose lim s
n
= s where every s
n
> m. Can we conclude that s > m?
Explain your answer.
10. (a) If
|
s
n
s
|
< 1, explain why
s
2
n
+ ss
n
+ s
2
< 1 + 3
|
s
|
+ 3
|
s
|
2
(b) Suppose s
n
s. Prove that s
3
n
s
3
.
28
9 Limit Theorems for Sequences
We’d like to develop some rules for working with limits so that we don’t have to resort to an ϵN
proof every time. The rough idea is that limits respect the basic rules of algebra. For instance. . .
Example 9.1. If lim s
n
= s, it seems natural that a new sequence (5s
n
) obtained by multiplying the
original terms by 5 should have limit lim 5s
n
= 5s. Consider what we have to prove to confirm this:
ϵ > 0, N such that n > N =
|
5s
n
5s
|
< ϵ
This last amounts to observing that
|
s
n
s
|
<
ϵ
5
. The challenge here is to see that we’re essentially
done: this is just the statement lim s
n
= s in disguise! Here is a more formal argument.
Let ϵ > 0 be given. Since lim s
n
= s, we know that
N such that n > N =
|
s
n
s
|
<
ϵ
5
=
|
5s
n
5s
|
< ϵ
The trick in the example will be used repeatedly in the proofs that follow. What’s critical is that
you read the limit definition correctly: given any small number (ϵ,
ϵ
5
, etc.) there is some tail of the
sequence which remains closer to the limit than this.
Theorem 9.2 (Limit laws). Limit calculations respect algebraic operations: ±, ×, ÷ and roots.
More specifically, if (s
n
) converges to s and (t
n
) to t, then,
1. lim(s
n
±t
n
) = s ±t
2. lim(s
n
t
n
) = st; as a special case, if k is constant, then lim ks
n
= ks
3. If t = 0, then lim
s
n
t
n
=
s
t
4. If k N, then lim
k
s
n
=
k
s, provided the roots exist (s
n
, s 0 if k even)
Our first example was the special case of part 2 with k = 5. Note also how parts 2 and 4 extend
Lemma 8.7: by induction we now have s
q
n
s
q
for any q Q.
Proving the limit laws takes a little work, including a small lemma. To advertise their benefit, we
repeated apply them to a limit calculation as you might have seen it in elementary calculus.
Examples 9.3. 1. We evaluate lim
3n
2
+2
n1
5n
2
2
using the limit laws.
lim
3n
2
+ 2
n 1
5n
2
2
= lim
3 +
2
n
3/2
1
n
2
5
2
n
2
=
lim
3 +
2
n
3/2
1
n
2
lim
5
2
n
2
(part 3)
=
lim 3 + lim
2
n
3/2
lim
1
n
2
lim 5 lim
2
n
2
(part 1)
=
3 + 0 0
5 0
=
3
5
(parts 2, 4 and lim
1
n
= 0 (Example 8.4.1))
This calculation involves some generally accepted sleight of hand; formally we’re working from
the bottom up since lim
3n
2
+2
n1
5n
2
2
shouldn’t really be written until you know it exists!
29
2. Suppose (s
n
) is defined inductively via s
1
= 2 and s
n+1
=
1
2
( s
n
+
2
s
n
):
( s
n
) =
2,
3
2
,
17
12
,
577
408
, . . .
This sequence in fact converges, though we’ll need to wait until the next section to see why.
Given this fact, the limit laws allow us to compute the limit s:
s = lim s
n+1
=
1
2
lim s
n
+
2
lim s
n
=
1
2
s +
2
s
=
1
2
s =
1
s
= s
2
= 2
Since s
n
is plainly always positive, we conclude that lim s
n
=
2.
We now commence our assault on the limit laws. The strategy for the first is simple: control both
sequences so that both
|
s
n
s
|
,
|
t
n
t
|
<
ϵ
2
, then add. The only challenge is writing it formally.
Proof of Theorem 9.2, part 1. Let ϵ > 0 be given. Since s
n
s and t
n
t, we see that N
1
, N
2
such that
N
1
such that n > N
1
=
|
s
n
s
|
<
ϵ
2
and,
N
2
such that n > N
2
=
|
t
n
t
|
<
ϵ
2
Let N = max{N
1
, N
2
}, then
n > N =
|
s
n
+ t
n
( s + t)
|
|
s
n
s
|
+
|
t
n
t
|
<
ϵ
2
+
ϵ
2
= ϵ
The argument for s
n
t
n
is almost identical.
The multiplicative limit law requires a preparatory result.
Lemma 9.4. (s
n
) convergent = (s
n
) bounded (M such that n,
|
s
n
|
M).
The converse to this statement is false: why?
The picture shows the strategy: taking ϵ = 1 in the limit
definition bounds an infinite tail of the sequence; the finitely
many terms that come before are a non-issue.
Proof. Suppose lim s
n
= s and let ϵ = 1 in the definition of
limit. Then N such that
n > N =
|
s
n
s
|
< 1 = s 1 < s
n
< s + 1
=
|
s
n
|
< max{
|
s 1
|
,
|
s + 1
|
}
It follows that every term of the sequence is bounded by
M := max
|
s 1
|
,
|
s + 1
|
,
|
s
n
|
: n N
s
n
s + 1
s 1
s
N
n
bounded tail
30
The approach to part 2 is similar to part 1, we just need to be a bit cleverer to break up
|
s
n
t
n
st
|
.
Proof of Theorem 9.2, part 2. Exercise 8.2 deals with (and extends) the case when s = 0. Instead sup-
pose s = 0, and let ϵ > 0 be given. Since s
n
s and t
n
t,
( t
n
) is bounded (Lemma) : M such that n,
|
t
n
|
M
N
1
, N
2
such that n > N
1
=
|
s
n
s
|
<
ϵ
2M
and n > N
2
=
|
t
n
t
|
<
ϵ
2
|
s
|
Again let N = max{N
1
, N
2
}, then
|
s
n
t
n
st
|
=
|
s
n
t
n
st
n
+ st
n
st
|
|
s
n
s
||
t
n
|
+
|
s
||
t
n
t
|
<
ϵ
2M
M +
|
s
|
ϵ
2
|
s
|
= ϵ
The proofs of parts 3 and 4 are in Exercise 6.
More basic examples With a few simple general examples, the limit laws allow us to rapidly com-
pute the limits of a great variety of sequences.
Theorem 9.5. 1. If k > 0 then lim
1
n
k
= 0
2. If
|
a
|
< 1 then lim a
n
= 0
3. If a > 0 then lim a
1/n
= 1
4. lim n
1/n
= 1
Examples 9.6. 1. lim(3n)
2/n
= (lim 3
1/n
)
2
(lim n
1/n
)
2
= 1.
2. lim
n
2/n
+
3 n
1
sin n
1/5
4n
3/2
+ 7
=
(lim n
1/n
)
2
3 lim
sin n
n
1/5
4 lim
1
n
3/2
+ 7
=
1
5
3
7
Note that lim
sin n
n
= 0 follows from the squeeze theorem:
sin n
n
1
n
0.
Proof. 1. This is Example 8.4.1.
2. The a = 0 case is trivial. Otherwise, given ϵ > 0, let N = log
|
a
|
ϵ and observe that
n > N =
|
a
n
|
<
a
N
=
|
a
|
N
= ϵ
3. Suppose a 1, and let s
n
:= a
1/n
1. Since s
n
> 0, the binomial theorem
15
shows that
a = (1 + s
n
)
n
1 + ns
n
= 0 < s
n
a 1
n
The squeeze theorem (8.9) shows that s
n
0, whence lim a
1/n
= 1.
We leave the a < 1 case and part 4 to Exercise 7.
15
(1 + x)
n
=
n
k=0
(
n
k
)
x
k
= 1 + nx +
n(n1)
2
x
2
+
n(n1)(n2)
2·3
x
3
+ ··· + x
n
.
31
Divergence to ± and the ‘divergence laws’
We now consider unbounded sequences and provide a positive definition of a type of divergence.
Definition 9.7. We say that (s
n
) diverges to if,
M > 0, N such that n > N = s
n
> M
We write s
n
or lim s
n
= . The definition for s
n
is similar.
If (s
n
) neither converges nor diverges to ±, we say that it diverges by oscillation.
16
Consider how M is trying to describe “closeness” to infinity similarly to how ϵ measures closeness
to s in the original definition of limit (8.1).
Examples 9.8. As with convergence proofs, it is a good idea to try some scratch work first!
1. We show that lim( n
2
+ 4n) = .
Let M > 0 be given, and let N =
M. Then
n > N = n
2
+ 4n > n
2
> N
2
= M
2. Prove that s
n
= n
5
n
4
2n + 1 .
The negative terms cause some trouble, though our solution should be familiar from previous
calculations:
s
n
>
1
2
n
5
n
5
> 2(n
4
+ 2n 1) n > 2 +
4
n
3
1
n
4
Certainly this holds if n > 6. We can now complete the proof.
Let M > 0 be given, and let N = max{6,
5
2M}. Then
n > N = s
n
>
1
2
n
5
>
1
2
(2M) = M
3. Prove that the sequence defined by s
n
= n
2
n
3
diverges to .
First observe that
s
n
= n
2
(1 n) <
1
2
n
3
1 n <
1
2
n n 2
Now let M > 0 be given,
17
and define N = max{2,
3
2M}. Then
n > N = n > 2 = s
n
<
1
2
n
3
<
1
2
N
3
M
16
In such cases lim s
n
is meaningless; you likely wrote lim s
n
= DNE (“does not exist”) in elementary calculus.
17
The notion that s
n
can be phrased in multiple ways: some prefer
m < 0, N such that n > N = s
n
< m (in our argument M = m)
32
Several of the limit laws can be adapted to sequences which diverge to ±.
Theorem 9.9. Suppose lim s
n
= .
1. If t
n
s
n
for all (large) n, then lim t
n
=
2. If lim t
n
exists and is finite, then lim s
n
+ t
n
= .
3. If lim t
n
> 0 then lim s
n
t
n
= .
4. lim
1
s
n
= 0
5. If lim t
n
= 0 and t
n
> 0 for all (large) n, then lim
1
t
n
=
Similar statements when s
n
should be clear.
Proof. We prove two of the results: try the rest yourself.
2. Since (t
n
) converges, it is bounded (below): m such that n, t
n
m. Let M be given. Since
lim s
n
= , N such that
n > N = s
n
> M m = s
n
+ t
n
> M m + m = M
4. Let ϵ > 0 be given, and let M =
1
ϵ
. Then N such that
n > N = s
n
> M =
1
ϵ
=
1
s
n
< ϵ
Rational Sequences We can now find the limit of any rational sequence:
p
n
q
n
where (p
n
), (q
n
) are
polynomials in n.
Example 9.10. By applying Theorem 9.9 (part 3) to
s
n
:= 3n + 4n
2
and t
n
=
1
2 n
2
1
2
we see that
lim
3n
3
+ 4
2n
2
1
= lim
3n + 4n
2
2 n
2
= lim(3n + 4n
2
) ·lim
1
2 n
2
=
Indeed, you should be able to confirm the familiar result from elementary calculus:
Corollary 9.11. If p
n
, q
n
are polynomials in n with leading coefficients p, q respectively then
lim
p
n
q
n
=
0 if deg(p
n
) < deg(q
n
)
p
q
if deg(p
n
) = deg(q
n
)
sgn(
p
q
) if deg(p
n
) > deg(q
n
)
33
Exercises 9. 1. Suppose lim x
n
= 3, lim y
n
= 7 and that all y
n
are non-zero. Determine the following:
(a) lim(x
n
+ y
n
) (b) lim
3y
n
x
n
y
2
n
(c) lim
p
x
n
y
n
+ 4
2. Consider s
n
= (100n)
100
n
. Describe s
1
(1 followed by how many zeros?). Repeat for s
10
. Now
compute the limit lim s
n
.
3. Define (s
n
) inductively via s
1
= 1 and s
n+1
=
s
n
+ 1 for n 1.
(a) List the first four terms of (s
n
).
(b) It turns out that (s
n
) converges. Assume this and prove that lim s
n
=
1
2
(1 +
5) .
4. Prove the following:
(a) lim(n
3
98n) = (b) lim
n n +
4
n
=
5. Let x
1
= 1 and x
n+1
= 3x
2
n
for n 1.
(a) Show that if (x
n
) converges with limit a, then a =
1
3
or a = 0.
(b) What is lim x
n
? Prove your assertion and explain what is going on.
6. We prove parts 3 and 4 of the limit laws (Theorem 9.2). Assume lim s
n
= s and lim t
n
= t.
(a) Suppose t = 0. Explain why N
1
such that n > N
1
=
|
t
n
|
>
1
2
|
t
|
.
(b) Let ϵ > 0 be given. Since t
n
t, N
2
such that n > N
2
=
|
t
n
t
|
<
1
2
|
t
|
2
ϵ. Combine
N
1
and N
2
to provide a proof that lim
1
t
n
=
1
t
.
(c) Explain how to conclude part 3: lim
s
n
t
n
=
s
t
.
(d) Use the following inequality (valid when s
n
, s > 0) to help construct a proof for part 4
s
1/k
n
s
1/k
=
|
s
n
s
|
s
k1
k
n
+ s
k2
k
n
s
1
k
+ ··· + s
k1
k
|
s
n
s
|
s
k1
k
7. We finish the proof of Theorem 9.5.
(a) Suppose 0 < a < 1. Prove that lim a
1/n
= 1.
(Hint: consider b =
1
a
. . . )
(b) Let s
n
= n
1/n
1. Apply the binomial theorem to n = (1 + s
n
)
n
to prove that s
n
<
q
2
n1
.
Hence conclude that lim n
1/n
= 1.
8. Prove the remaining parts of Theorem 9.9.
9. Assume s
n
= 0 for all n, and that the limit L = lim
s
n+1
s
n
exists.
(a) Show that if L < 1, then lim s
n
= 0.
(Hint: if L < a < 1, obtain N so that n > N =
|
s
n
|
< a
nN
|
s
N
|
)
(b) Show that if L > 1, then lim
|
s
n
|
= +.
(Hint: apply (a) to the sequence t
n
=
1
|
s
n
|
)
(c) Let p > 0 and a R be given. How does lim
n
a
n
n
p
depend on the value of a?
34
10 Monotone and Cauchy Sequences
The definition of limit (Definition 8.1) exhibits a major weakness; to demonstrate the convergence of
a sequence, we must already know its limit! What we’d like is a method for determining whether a
sequence converges without first guessing a suitable limit.
18
In this section we consider two important
classes of sequence for which this can be done.
Definition 10.1. A sequence (s
n
) is:
Monotone-up
19
if s
n+1
s
n
for all n.
Monotone-down if s
n+1
s
n
for all n.
Monotone if either of the above is true.
Examples 10.2. 1. The sequence with n
th
term s
n
=
7
n
+ 4 is (strictly) monotone-down:
s
n+1
=
7
n + 1
<
7
n
= s
n
2. A constant sequence (s
n
) = (s, s, s, s, . . .) is both monotone-up and monotone-down.
Theorem 10.3 (Monotone Convergence).
Every bounded monotone sequence is convergent.
Specifically:
If (s
n
) is bounded above and monotone-up,
then lim s
n
exists and equals sup{s
n
}.
If (s
n
) is bounded below and monotone-down,
then lim s
n
exists and equals inf{s
n
}.
n
s
n
sup{s
n
}
In fact the conclusion lim s
n
= sup{s
n
} holds for all monotone-up sequences: if unbounded above,
then the result is (see Exercise 5). The statement is lim s
n
= inf{s
n
} for monotone-down sequences.
Proof. If (s
n
) is bounded above, then s := sup{s
n
} exists by the completeness axiom (s is finite!).
Let ϵ > 0 be given. By Lemma 4.8, there exists some s
N
> s ϵ. Since (s
n
) is monotone-up, we have
n > N = s
n
s
N
> s ϵ = 0 s s
n
< ϵ =
|
s s
n
|
< ϵ
The monotone-down case is similar.
18
This gets at the typical role of sequences in analysis: to demonstrate the existence of and define a new object (the limit)
and, more broadly, to transfer useful properties from the sequence to the limit. For instance, if ( f
n
) is a sequence of differ-
entiable functions, we’d like to know if lim f
n
(x) exists and is itself differentiable with derivative lim f
n
(x): discussions of
this ilk will dominate Math 140B.
19
Some authors describe such as sequence as either non-decreasing or increasing. We prefer monotone-up/down since this
directly describes the direction of any potential movement in the sequence and prevents confusion over whether the in-
equality is strict. A sequence with s
n+1
> s
n
may be described as strictly increasing or strictly monotone-up.
35
Examples 10.4. 1. Define (s
n
) via s
n
= 1 and s
n+1
=
1
5
( s
n
+ 8):
( s
n
) =
1, 1.8, 1.96, 1.992, 1.9984, 1.99968, . . .
The sequence certainly appears to be monotone-up and converging to 2. We prove this:
Bounded above: s
n
< 2 = s
n+1
<
1
5
[
2 + 8
]
= 2. By induction, (s
n
) is bounded above by 2.
Monotone-up: s
n+1
s
n
=
4
5
[
2 s
n
]
> 0 since s
n
< 2.
Convergence: By monotone convergence, s = lim s
n
exists. Now use the limit laws to find s:
s = lim s
n+1
=
1
5
(
lim s
n
+ 8
)
=
1
5
( s + 8) = s = 2
2. (Example 9.3.2, cont.) Let s
1
= 2 and s
n+1
=
1
2
s
n
+
2
s
n
.
Bounded below: The sequence is plainly always positive and thus bounded below by zero.
Monotone-down: We first obtain an improved lower bound:
s
2
n+1
=
1
4
s
n
+
2
s
n
2
= 2 +
1
4
s
n
2
s
n
2
2
shows
20
that s
2
n
2 for all n. It follows that
s
n+1
s
n
=
1
2
1 +
2
s
2
n
1 = s
n+1
s
n
Convergence: By monotone convergence, s = lim s
n
exists. Example 9.3.2 provides the limit:
s =
1
2
s +
2
s
= s =
2
This shows the necessity of completeness: (s
n
) is a monotone, bounded sequence of rational
numbers, but its limit is irrational.
3. A decimal 0.d
1
d
2
d
3
. . . may be viewed as the limit of a monotone-up sequence of rational numbers:
0.d
1
d
2
d
3
. . . = lim
n
n
k=0
d
k
10
k
This is bounded above by 1 and so converges. Compare this with Example 4.4.
4. The sequence with s
n
=
1 +
1
n
n
is particularly famous. In Exercise 10 we show that ( s
n
) is
monotone-up and bounded above. The limit provides, arguably, the oldest definition of e:
e := lim
1 +
1
n
n
20
In case you’ve seen it before, this is the famous AM–GM inequality
x+y
2
xy with x = s
n
and y =
2
s
n
.
36
Limits Superior and Inferior
One interpretation of lim s
n
is that it approximately describes s
n
for large n. Even when a sequence
does not have a limit, it remains useful to be able to describe its long-term behavior.
Definition 10.5. Let (s
n
) be a sequence and define two related sequences (v
N
) and (u
N
):
v
N
:= sup{s
n
: n > N}, u
N
:= inf{s
n
: n > N}
1. The limit superior of (s
n
) is
lim sup s
n
=
(
lim
N
v
N
if (s
n
) bounded above
if (s
n
) unbounded above
2. The limit inferior of (s
n
) is
lim inf s
n
=
(
lim
N
u
N
if (s
n
) bounded below
if (s
n
) unbounded below
The original sequence (s
n
) is almost wedged between
21
( v
n
) and (u
n
) in a situation reminiscent of
the squeeze theorem (except lim sup and lim inf need not be equal). The next result summarizes the
situation more formally; we omit the proof since these claims should be clear from the definition and
previous results, particularly the monotone convergence theorem.
Lemma 10.6. 1. (v
N
) is monotone-down, (u
N
) is monotone-up, and u
N
s
N+1
v
N
.
2. lim sup s
n
and lim inf s
n
exist for any sequence (they might be infinite).
3. lim inf s
n
lim sup s
n
.
Examples 10.7. 1. The picture shows sequences ( s
n
), (u
N
) and (v
N
) when s
n
= 6 + (1)
n
1 +
5
n
We won’t compute everything precisely, but the
picture suggests (s
n
) has two “sub”sequences:
the odd terms increase while the even terms de-
crease towards, respectively
lim inf s
n
= 5, lim sup s
n
= 7
Here is one value from each derived sequence:
u
3
= inf{s
n
: n > 3} = s
5
= 4
v
7
= sup{s
n
: n > 7} = s
8
= 7.625
0
0 10
7
5
s
n
u
N
v
N
n/N
3 7
u
3
= inf{s
n
: n > 3}
2. If s
n
=
1
n
, then v
N
= s
N+1
and u
N
= 0 for all N, whence lim sup s
n
= lim inf s
n
= 0.
21
A minor redefinition would remove the ‘almost,’ but at the cost of making some subsequent arguments a little messier.
It is still reasonable to think of (u
N
) and (v
N
) as providing a long-term envelope for the original sequence.
37
3. Let s
n
= (1)
n
. This time the calculation is easy:
u
N
= inf{s
n
: n > N} = 1 and v
N
= sup{s
n
: n > N} = 1
Therefore lim sup s
n
= 1 and lim inf s
n
= 1.
Lemma 10.8. For any sequence (s
n
),
lim inf s
n
= lim sup s
n
= lim s
n
exists
(the limit can be infinite!).
In such a case all three values are equal.
0
s
n
0
lim s
n
n
In fact the converse to this is also true: we could prove it now, but it will come for free a little later. . .
Proof. (s finite) Since u
n1
s
n
v
n1
for all n, the squeeze theorem tells us that lim s
n
= s.
(s = ) Since u
n1
s
n
for all n and lim u
n1
= , it follows (Theorem 9.9.1) that lim s
n
= .
(s = ) This time s
n
v
n1
= lim s
n
= .
Cauchy Sequences
We now come to a class of sequences whose analogues will dominate your study of analysis.
Definition 10.9. A sequence (s
n
) is Cauchy
22
if
ϵ > 0, N such that m, n > N =
|
s
n
s
m
|
< ϵ
A sequence is Cauchy when terms in the tails of the sequence are constrained to stay close to one
another. As we’ll see shortly, this will provide an alternative way to detect and describe convergence.
Examples 10.10. 1. Let s
n
=
1
n
. Let ϵ > 0 be given and let N =
1
ϵ
. Then
23
m > n > N =
|
s
m
s
n
|
=
1
n
1
m
<
1
n
<
1
N
= ϵ
Thus (s
n
) is Cauchy. A similar argument works for any s
n
=
1
n
k
for positive k.
2. Suppose s
1
= 5 and s
n+1
= s
n
+
1
n(n+1)
. As before, let ϵ > 0 be given and let N =
1
ϵ
. Then,
|
s
n+1
s
n
|
=
1
n( n + 1)
=
1
n
1
n + 1
=
|
s
m
s
n
|
|
s
n+1
s
n
|
+ ··· +
|
s
m
s
m1
|
=
1
n
1
m
<
1
n
<
1
N
= ϵ
Again we have a Cauchy sequence.
22
Augustin-Louis Cauchy (1789–1857) was a French mathematician, responsible (in part) for the ϵ-N definition of limit.
38
3. Define (s
n
)
n=0
inductively:
s
0
= 1, s
n+1
=
(
s
n
+ 3
n
if n even
s
n
2
n
if n odd
( s
n
) =
1, 2,
3
2
,
29
18
,
107
72
, . . .
Since
|
s
n+1
s
n
|
2
n
, we see that
0
1
2
0 2 4 6 8 10 12 14
n
s
n
m > n =
|
s
m
s
n
|
|
s
n+1
s
n
|
+ ··· +
|
s
m
s
m1
|
=
m1
k=n
|
s
k+1
s
k
|
m1
k=n
2
k
=
2
n
2
m
1 2
1
< 2
1n
where we used the familiar geometric sum formula from calculus:
b1
k=a
r
k
=
r
a
r
b
1r
.
Suppose ϵ > 0 is given, and let N = 1 log
2
ϵ = log
2
2
ϵ
. Then
m > n > N =
|
s
m
s
n
|
< 2
1n
< 2
1N
= ϵ
We conclude that (s
n
) is Cauchy.
The picture in the last example illustrates the essential point regarding Cauchy sequences: (s
n
) ap-
pears very much to converge. . .
Theorem 10.11 (Cauchy Completeness). A sequence of real numbers is convergent if and only if it
is Cauchy.
Proof. ( ) Suppose lim s
n
= s (finite). Given ϵ > 0 we may choose N such that
m, n > N =
|
s
n
s
|
<
ϵ
2
and
|
s
m
s
|
<
ϵ
2
=
|
s
n
s
m
|
=
|
s
n
s + s s
m
|
|
s
n
s
|
+
|
s s
m
|
< ϵ
whence (s
n
) is Cauchy.
( ) To discuss the convergence of (s
n
) we first need a potential limit! In view of Lemma 10.8, the
obvious candidates are lim sup s
n
and lim inf s
n
. We have two goals: show that (s
n
) is bounded
whence the limits superior and inferior are finite; then show that these are equal.
(Boundedness of (s
n
)) Take ϵ = 1 in Definition 10.9:
N such that m, n > N =
|
s
n
s
m
|
< 1
It follows that
n > N =
|
s
n
s
N+1
|
< 1 = s
N+1
1 < s
n
< s
N+1
+ 1
whence (s
n
) is bounded; it follows that lim sup s
n
and lim inf s
n
are finite.
23
Since m, n are arbitrary, WLOG we may assume m > n; equality is never interesting in these situations. This assump-
tion is very common and we’ll use it repeatedly without comment.
39
(lim sup s
n
= lim inf s
n
) Since (s
n
) is Cauchy, given ϵ > 0,
N such that m, n > N =
|
s
n
s
m
|
< ϵ = s
n
< s
m
+ ϵ
Taking v
N
= sup{s
n
: n > N}, we see that
m > N = v
N
s
m
+ ϵ
Taking the infimum of the right hand side yields
v
N
u
N
+ ϵ (since u
N
= inf{s
m
: m > N})
Since (v
N
) is monotone-down and (u
N
) monotone-up, we see that
lim sup s
n
v
N
u
N
+ ϵ lim inf s
n
+ ϵ = lim sup s
n
lim inf s
n
+ ϵ
This last holds for all ϵ > 0, whence lim sup s
n
lim inf s
n
. By Lemma 10.6 we have
equality.
By Lemma 10.8, we conclude that (s
n
) converges to lim sup s
n
= lim inf s
n
.
In view of the Theorem, the previous examples converge. All three limits can be found precisely
(for instance, see Exercise 7). With a small modification to the second example, however, we obtain
something genuinely new:
Example (10.10.2 cont). Let s
1
= 5 and , for each n, define s
n+1
= s
n
+
sin n
n(n+1)
. Since
|
sin n
|
1, the
computation proceeds almost the same as before:
|
s
n+1
s
n
|
=
|
sin n
|
n( n + 1)
1
n( n + 1)
= ···
The new sequence is Cauchy and therefore convergent; good luck explicitly finding its limit though!
The main point is easy to miss: Cauchy Completeness provides a powerful tool for determining
whether a sequence converges without first guessing a limit. While the result depends on monotone
convergence (via limit superior/inferior), it is more powerful in that it applies even to non-monotone
sequences. We finish with an application of this idea.
An Alternative Definition of R Cauchy sequences suggest a definition of the real numbers which
does not rely on Dedekind cuts (Section 6).
Define an equivalence relation on the collection C of all Cauchy sequences of rational numbers:
24
( s
n
) (t
n
) lim(s
n
t
n
) = 0
We then define R :=
C
. All this is done without reference to Cauchy Completeness, though it
certainly informs our intuition that (s
n
) and (t
n
) have the same limit. Some work is still required to
define +, ·, , etc., and to verify the axioms of a complete ordered field—we won’t pursue this.
24
We don’t need real numbers to define the limit of the rational sequence (s
n
t
n
): ϵ Q
+
is enough. . .
40
Exercises 10. 1. Use the definition to show that the sequence with n
th
term s
n
=
1
n
2
is Cauchy.
Repeat for t
n
=
1
n(n2)
.
2. Let s
1
= 1 and s
n+1
=
n
n+1
s
2
n
for n 1.
(a) Find s
2
, s
3
and s
4
.
(b) Show that lim s
n
exists and hence prove that lim s
n
= 0.
3. Let s
1
= 1 and s
n+1
=
1
3
( s
n
+ 1) for n 1.
(a) Find s
2
, s
3
and s
4
.
(b) Use induction to show that s
n
>
1
2
for all n, and conclude that (s
n
) is monotone-down.
(c) Show that lim s
n
exists and find lim s
n
.
4. (a) Let (s
n
) be a sequence such that n,
|
s
n+1
s
n
|
3
n
. Prove that (s
n
) is Cauchy.
(b) Let s
1
= 10 and, for each n, let s
n+1
= s
n
+
cos n
3
n
. Explain why (s
n
) is convergent.
(c) Is the result in (a) true if we only assume that
|
s
n+1
s
n
|
1
n
for all n N?
5. Suppose (s
n
) is unbounded and monotone-up. Prove that lim s
n
= .
(Thus lim s
n
= sup{s
n
} for any monotone-up sequence)
6. Let s
n
=
(1)
n
n
. Find the sequences (u
N
), (v
N
) and explicitly compute lim sup s
n
and lim inf s
n
.
7. Consider the sequence in Example 10.10.3. Explain why s
2n
= s
2n2
2
4
n
+
9
9
n
.
Now use the geometric sum formula to evaluate lim s
2n
.
(Since (s
n
) converges, this means the original sequence has the same limit)
8. Let S be a bounded nonempty set for which sup S / S. Prove that there exists a monotone-up
sequence (s
n
) of points in S such that lim s
n
= sup S.
(Hint: for each n, use sup S
1
n
to build s
n
)
9. Let (s
n
) be a monotone-up sequence of positive numbers and define σ
n
=
1
n
( s
1
+ s
2
+ ···+ s
n
).
Prove that (σ
n
) is monotone-up.
10. (Hard!) We prove that the sequence defined by s
n
=
1 +
1
n
n
is convergent.
(a) Show that
1 +
1
n+1
1 +
1
n
= 1
1
( n + 1)
2
and
1 +
1
n
1 +
1
n+1
= 1 +
1
n( n + 2)
(b) Prove Bernoulli’s inequality by induction.
For all real x > 1 and n N
0
we have ( 1 + x)
n
1 + nx.
(c) Use parts (a) and (b) to prove that ( s
n
) is monotone-up.
(Hint: consider
s
n+1
s
n
)
(d) Similarly, show that t
n
:=
1 +
1
n
n+1
=
1 +
1
n
s
n
defines a monotone-down sequence.
(e) Prove that ( s
n
) and (t
n
) converge, and to the same limit (this limit is e).
(Hint: compute t
n
s
n
)
41
11 Subsequences
The general behavior of a sequence is often hard to ascertain, but if we delete some of its terms we
might obtain a subsequence with interesting behavior.
Definition 11.1. Let (s
n
) be a sequence. A subsequence (s
n
k
) is a subset (s
n
k
) (s
n
), where
n
1
< n
2
< n
3
< ···
A subsequence is simply an infinite subset, with order inherited from the original sequence.
Example 11.2. Take s
n
= (1)
n
(recall Example 8.6.2) and let
n
k
= 2k. Then s
n
k
= 1 for all k. Note two important facts:
The subsequence (s
n
k
)
k=0
is indexed by k, not n.
The subsequence is constant and thus convergent.
1
0
1
s
n
n
Our main goal in this section is to prove the result illustrated in the example, that every bounded
sequence has a convergent subsequence (the famous Bolzano–Weierstraß theorem).
Lemma 11.3. If lim
n
s
n
= s, then every subsequence (s
n
k
) also satisfies lim
k
s
n
k
= s.
Proof. Suppose s is finite and let ϵ > 0 be given. Then N such that n > N =
|
s
n
s
|
< ϵ. Since
n
k
k for all k, we see that
k > N = n
k
> N =
|
s
n
k
s
|
< ϵ
The case where s = ± is an exercise.
Lemma 11.4. Every sequence has a monotonic subsequence.
Proof. Given (s
n
), we call the term s
n
‘dominant’ if m > n = s
m
< s
n
. There are two cases:
1. If there are infinitely many dominant terms, then the subsequence of such is monotone-down.
2. If there are finitely many dominant terms, choose s
n
1
after all such. Since s
n
1
is not dominant,
n
2
> n
1
such that s
n
2
s
n
1
. Induct to obtain a monotone-up subsequence.
n
s
n
s
n
k
dominant terms
n
s
n
s
n
k
Case 1: monotone-down subsequence Case 2: monotone-up subsequence
42
Theorem 11.5. Given a sequence (s
n
), there exist subsequences (s
n
k
) and (s
n
l
) such that
lim s
n
k
= lim sup s
n
and lim s
n
l
= lim inf s
n
Combining with the lemmas, we may assume these subsequences are monotonic.
Example 11.6. The picture shows the sequence with
n
th
term
s
n
=
(
4
n
( 1)
n
2
+1
when n is even
1
1
n
when n is odd
Monotonic subsequences with limits lim sup s
n
= 1
and lim inf s
n
= 0 are indicated.
1
0
1
2
s
n
10 20
n
Proof. We prove only the lim sup claim since the other is similar. There are three cases to consider;
visualizing the third is particularly difficult and may take several readings.
(lim sup s
n
= ) Since (s
n
) is unbounded above, for any
k > 0 there exist infinitely many terms s
n
> k. We may
therefore inductively choose a subsequence (s
n
k
) via
n
1
= min{n N : s
n
1
> 1}
n
k
= min{n N : n
k
> n
k1
, s
n
k
> k}
Choosing the minimum isn’t necessary here, but it at
least keeps the subsequence explicit. Clearly
s
n
k
> k = lim
k
s
n
k
= = lim sup s
n
0
1
2
3
4
5
s
n
1
s
n
2
s
n
3
s
n
4
s
n
5
n
s
n
Example: lim sup
n
2
(1 + sin n) =
(lim sup s
n
= ) Since lim inf s
n
lim sup s
n
= , Lemma 10.8 says that lim s
n
= , whence
( s
n
) itself is a suitable subsequence.
(lim sup s
n
= v finite) We follow an inductive construction: let n
1
= 1 and define s
n
k
for k 2 via,
Since (v
N
) is monotone-down and converges to v, take ϵ =
1
2k
to see that
25
N
k
n
k1
such that v v
N
k
< v +
1
2k
Since v
N
k
= sup{s
n
: n > N
k
}, Lemma 4.8 says
n
k
> N
k
such that s
n
k
> v
N
k
1
2k
But then
|
v s
n
k
|
|
v v
N
k
|
+
|
v
N
k
s
n
k
|
<
1
k
. The squeeze theorem says that lim
k
s
n
k
= v.
25
(v
N
) being monotone-down is crucial: if N satisfies v
N
v <
1
2k
, so does N
k
:= max{N, n
k1
}.
43
Example (11.6 cont.). The example shows why the two-step construction is necessary. It may seem
that we should simply be able to modify subsequences of (u
N
) and (v
N
). Indeed,
( u
N
) =
1, 1, 1, 1,
1
2
,
1
2
,
1
2
,
1
2
,
1
3
,
1
3
,
1
3
,
1
3
, . . .
contains a monotonic subsequence of (s
n
) converging to lim inf s
n
= 0. Unfortunately, the same isn’t
true for (v
N
) = (2, 1, 1, 1, 1 . . .), where v
N
k
= 1 for all k 2; taking n
k
= 2k 1 results in
s
n
k
= 1
1
2k 1
> 1
1
2k
= v
N
k
1
2k
The above discussion rapidly provides two results, the first of which is Exercise 3.
Theorem 11.7 (Lemma 10.8 with converse). For any sequence,
lim sup s
n
= lim inf s
n
lim s
n
exists
Theorem 11.8 (Bolzano–Weierstraß). Every bounded sequence has a convergent subsequence.
Proof 1. Lemma 11.4 says there exists a monotone subsequence. This is bounded and thus converges
by the monotone convergence theorem.
Proof 2. By Theorem 11.5, there exists a subsequence converging to the finite value lim sup s
n
.
For a third proof(!) we present the classic ‘shrinking-interval’ argument which has the benefit of
generalizing to higher dimensions (rather than intervals, take boxes. . . ).
Proof 3. Suppose (s
n
) is bounded by M. One of the intervals [M, 0] or [0, M] must contain infinitely
many terms of the sequence (perhaps both!). Call this interval E
0
and choose any n
0
E
0
.
Split E
0
into left- and right half-intervals, one of which must contain infinitely many terms of the
sequence for which n > n
0
;
26
call this half-interval E
1
and choose any s
n
1
E
1
for which n
1
> n
0
.
Repeat this ad infinitum to obtain a subsequence (s
n
k
) and a family of nested intervals
[M, M] E
0
E
1
E
2
··· of width
|
E
k
|
=
M
2
k
with s
n
k
E
k
It remains only to see that (s
n
k
) converges; we leave this to Exercise 5.
Example 11.9. (s
n
) = (sin n) is bounded and therefore has a
convergent subsequence! Its limit s must lie in the interval [1, 1].
The picture shows the first 1000 terms—remember that n is mea-
sured in radians. It is not at all clear from the picture what s or
our mystery subsequence should be! There is a reason for this, as
we’ll see momentarily. . .
1
0
1
n
s
n
26
Only finitely many terms in (s
n
) come before s
n
0
. . .
44
Subsequential Limits, Divergence by Oscillation & Closed Sets
Recall Definition 9.7, where we stated that a sequence (s
n
) diverges by oscillation if it neither converges
nor diverges to ±. We can now give a positive statement of this idea.
( s
n
) diverges by oscillation
Thm 11.7
lim inf s
n
= lim sup s
n
Thm 11.5
( s
n
) has subsequences tending to different limits
The word oscillation comes from the third interpretation: if s
1
= s
2
are limits of two subsequences,
then any tail of the sequence {s
n
: n > N} contains infinitely many terms arbitrarily close to s
1
and
infinitely many (other) terms arbitrarily close to s
2
. The original sequence (s
n
) therefore oscillates
between neighborhoods of s
1
and s
2
. Of course there could be many other subsequential limits. . .
Definition 11.10. We call s R } a subsequential limit of a sequence (s
n
) if there exists a
subsequence (s
n
k
) such that lim
k
s
n
k
= s.
Examples 11.11. 1. The sequence defined by s
n
=
1
n
has only one subsequential limit, namely zero.
Recall Lemma 11.3: lim s
n
= 0 implies that every subsequence also converges to 0.
2. If s
n
= (1)
n
, then the subsequential limits are ±1.
3. The sequence s
n
= n
2
(1 + ( 1)
n
) has subsequential limits 0 and .
4. All positive even integers are subsequential limits of (s
n
) = (2, 4, 2, 4, 6, 2, 4, 6, 8, 2, 4, 6, 8, 10, . . .).
5. (Hard!) Recall the countability of Q from a previous class: the standard argument enumerates
the rationals by constructing a sequence
(r
n
) =
0
1
,
1
1
,
1
1
|{z}
|
p
|
+q=2
,
1
2
,
1
2
,
2
1
,
2
1
| {z }
|
p
|
+q=3
,
1
3
,
1
3
,
3
1
,
3
1
| {z }
|
p
|
+q=4
,
1
4
,
1
4
,
2
3
,
2
3
,
3
2
,
3
2
,
4
1
,
4
1
| {z }
|
p
|
+q=5
, . . .
We claim that the set of subsequential limits of (r
n
) is in fact the full set of R {±}!
To see this, let a R be given and choose a subsequence (r
n
k
) inductively:
By the density of Q in R (Corollary 4.12), the set S
n
= Q (a
1
n
, a +
1
n
) contains infinitely
many rational numbers and thus infinitely many terms of the sequence (r
n
).
Choose any r
n
1
S
1
and, for each k 2, choose any
27
r
n
k
S
k
such that n
k
> n
k1
Since
|
r
n
k
a
|
<
1
k
, we conclude that lim
k
r
n
k
= a.
An argument for the subsequential limits ± is in the Exercises. Somewhat amazingly, the
specific sequence (r
n
) is irrelevant: the conclusion is the same for any sequence enumerating Q!
6. (Even harder—Example 11.9, cont.) We won’t prove it, but the set of subsequential limits of
( s
n
) = (sin n) is the entire interval [1, 1]! Otherwise said, for any s [1, 1] there exists a
subsequence ( sin n
k
) such that lim
k
sin n
k
= s.
45
Theorem 11.12. Let (s
n
) be a sequence in R and let S be its set of subsequential limits. Then
1. S is non-empty (as a subset of R {±}).
2. sup S = lim sup s
n
and inf S = lim inf s
n
.
3. lim s
n
exists iff S has only one element: namely lim s
n
.
Proof. 1. By Theorem 11.5, lim sup s
n
S.
2. By part 1, lim sup s
n
sup S. For any convergent subsequence ( s
n
k
), we have n
k
k, whence
N, {s
n
k
: k > N} {s
n
: n > N} = lim s
n
k
= lim sup s
n
k
lim sup s
n
Since this holds for every convergent subsequence, we have sup S lim sup s
n
, and therefore
equality. The result for inf S is similar.
3. Applying Theorem 11.7, we see that lim s
n
exists if and only if
lim sup s
n
= lim inf s
n
sup S = inf S S has only one element
Closed Sets You should be comfortable with the notion of a closed interval (e.g. [0, 1]) from elemen-
tary calculus. Using sequences, we can make a formal definition.
Definition 11.13. Let A R.
We say that s R is a limit point of A if there exists a sequence (s
n
) A converging to s.
The closure A is the set of limit points of A.
A is closed if it equals its closure: A = A.
Examples 11.14. 1. The interval [0, 1] is closed. If (s
n
) [0, 1] has lim s
n
= s, then
0 s
n
1
Thm 8.8
= s [0, 1]
More generally, every ‘closed interval’ [a, b] is closed, as are finite unions of closed intervals, for
instance [1, 5] [7, 11].
2. The interval (0, 1] is not closed since its closure is (0, 1] = [0, 1]. In particular, the sequence
s
n
=
1
n
lies in the original interval but has limit 0. Indeed this example shows that an infinite
union of closed intervals need not be closed.
Theorem 11.15. If (s
n
) is a sequence, then its set of (finite) subsequential limits is closed.
We omit the proof since it is hard to read, involving unpleasantly many subscripts (subsequences of
subsequences. . . ).
27
As in the proof of Theorem 11.5, we could make this more explicit by choosing minimums, but there is no need: if
there are infinitely many r
n
in S
k
, then only finitely many of them can come before r
n
k1
.
46
Exercises 11. 1. Consider the sequences with the following n
th
terms:
a
n
= (1)
n
b
n
=
1
n
c
n
= n
2
d
n
=
6n + 4
7n 3
(a) For each sequence, give an example of a monotone subsequence.
(b) For each sequence, state its set of subsequential limits.
(c) For each sequence, state its lim sup and lim inf.
(d) Which of the sequences converge? diverge to +? diverge to ?
(e) Which of the sequences are bounded?
2. Prove the case of Lemma 11.3 when lim s
n
=
3. Suppose that lim s
n
= s (could be ±). Use Theorem 11.5 and Lemma 11.3 to prove that
lim sup s
n
= s = lim inf s
n
.
(This completes the proof of Theorem 11.7)
4. Suppose that L = lim s
2
n
exists and is finite.
(a) Given an example of such a sequence where ( s
n
) is divergent.
(b) Prove that (s
n
) contains a convergent subsequence. What are the possible limits of this
subsequence? Why?
(Hint: use Bolzano–Weierstraß)
5. Complete the third proof of Bolzano–Weierstraß (Theorem 11.8) by proving that the constructed
subsequence (s
n
k
) is Cauchy.
6. (a) Show that the closed interval [a, b] is a closed set in the sense of Definition 11.13.
(b) Is there a sequence ( s
n
) such that (0, 1) is its set of subsequential limits?
7. Let (r
n
) be any sequence enumerating of the set Q of rational numbers. Show that there exists
a subsequence (r
n
k
) such that lim
k
r
n
k
= +.
(Hint: modify the argument in Example 11.11.5)
8. (Hard) Let (s
n
) be the sequence of numbers defined
in the figure, listed in the indicated order.
(a) Find the set S of subsequential limits of (s
n
).
(b) Determine lim sup s
n
and lim inf s
n
.
1
1
2
1
3
1
4
1
5
···
1
1
2
1
3
1
4
1
5
···
1
1
2
1
3
1
4
1
5
···
1
1
2
1
3
1
4
1
5
···
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
47
12 Lim sup and Lim inf
In this section we collect a couple of useful results, mostly for later use. First, we observe that the
limit laws do not work as tightly for limits superior and inferior.
Theorem 12.1. Let (s
n
), (t
n
) be bounded sequences.
1. lim sup(s
n
+ t
n
) lim sup s
n
+ lim sup t
n
2. If, in addition, (s
n
) is convergent to s, then we have equality
lim sup(s
n
+ t
n
) = s + lim sup t
n
Modifications can be made infima and products of sequences (Exercise 3).
Example 12.2. To see that equality is unlikely, take s
n
= (1)
n
= t
n
, then
lim sup(s
n
+ t
n
) = 0 < 2 = lim sup s
n
+ lim sup t
n
Proof. 1. For each N, the set {s
n
+ t
n
: n > N} is bounded above by
sup{s
n
: n > N}+ sup{t
n
: n > N}
from which
sup{s
n
+ t
n
: n > N} sup{s
n
: n > N}+ sup{t
n
: n > N}
Simply take limits as N for the first result.
2. By part 1, we already know that
lim sup(s
n
+ t
n
) s + lim sup t
n
For the other direction, let a
n
= s
n
+ t
n
and apply part 1 again:
lim sup t
n
= lim sup
( s
n
+ t
n
) s
n
lim sup(s
n
+ t
n
) + lim sup(s
n
)
= lim sup(s
n
+ t
n
) s
The next result will be critical when we study infinite series, particularly the ratio and root tests.
Theorem 12.3. Let (s
n
) be a non-zero sequence. Then
lim inf
s
n+1
s
n
lim inf
|
s
n
|
1/n
lim sup
|
s
n
|
1/n
lim sup
s
n+1
s
n
In particular, lim
s
n+1
s
n
= L = lim
|
s
n
|
1/n
= L ( )
48
Examples 12.4. 1. Here is a quick proof that lim n
1/n
= 1 (recall Theorem 9.5): let s
n
= n, then
lim
s
n+1
s
n
= lim
n + 1
n
= 1 = lim n
1/n
= lim
|
s
n
|
1/n
= 1
2. Let s
n
= n! and apply the corollary to see that
lim(n!)
1/n
= lim
s
n+1
s
n
= lim(n + 1) =
Proof. Assume lim sup
s
n+1
s
n
= L = (otherwise the third inequality is trivial) and let ϵ > 0. Then
lim
N
sup
s
n+1
s
n
: n > N
< L + ϵ = N such that sup
s
n+1
s
n
: n > N
< L + ϵ
For brevity, denote a = L + ϵ and b = a
N1
|
s
N+1
|
. For any n > N, we therefore have
s
n+1
s
n
< a =
|
s
n
|
< a
nN1
|
s
N+1
|
=
|
s
n
|
1/n
< a
a
N1
|
s
N+1
|
1/n
= ab
1/n
= lim sup
|
s
n
|
1/n
a lim b
1/n
= a = L + ϵ
Since this holds for all ϵ > 0, we conclude the third inequality: lim sup
|
s
n
|
1/n
L.
The second inequality is trivial and the first is similar to the third.
Exercises 12. 1. Compute lim
1
n
( n!)
1/n
(Hint: let s
n
=
n!
n
n
in Theorem 12.3 and recall that lim
1 +
1
n
n
= e)
2. Evaluate lim
(2n)!
(n!)
2
1/n
3. Let (s
n
) and (t
n
) be non-negative, bounded sequences.
(a) Prove that lim sup(s
n
t
n
)
(
lim sup s
n
) (
lim sup t
n
)
(b) Give an example which shows that we do not expect equality in part (a).
(c) If, in addition, lim s
n
= s, prove that lim sup(s
n
t
n
) = s lim sup t
n
.
4. Consider the sequence with s
2m
= s
2m+1
= 2
m
:
( s
n
)
n=0
=
1, 1,
1
2
,
1
2
,
1
4
,
1
4
,
1
8
,
1
8
, . . .
Compute
|
s
n
|
1/n
and
s
n+1
s
n
when n is even and then when it is odd. Thus find all expressions
in Theorem 12.3 and hence conclude that the converse of () is false.
49
14 Series
What should be meant by the following expression?
n=1
1
n
2
= 1 +
1
4
+
1
9
+
1
16
+ ···
The -symbol in the summation should lead you to suspect a role for limits. . .
Definition 14.1. Define the n
th
partial sum s
n
of a sequence (a
n
)
n=m
via
s
n
:=
n
k=m
a
k
= a
m
+ a
m+1
+ ··· + a
n
The (infinite) series
28
n=m
a
n
is the limit lim s
n
of the sequence (s
n
) of partial sums.
A series converges, s to ± or diverges by oscillation as does the sequence (s
n
).
a
n
converges absolutely if
|
a
n
|
converges.
a
n
converges conditionally if it converges but not absolutely (
|
a
n
|
diverges to ).
To return to our motivating example,
n=1
1
n
2
= lim s
n
where s
n
=
n
k=1
1
k
2
= 1 +
1
4
+ ··· +
1
n
2
We don’t (yet) know whether the series converges or diverges to . . .
Theorem 14.2 (Basic Series Laws). Infinite series behave nicely with respect to addition and scalar
multiplication. For instance:
1. If
a
n
is convergent and k is constant, then
ka
n
= k
a
n
is convergent.
2. If
a
n
and
b
n
are convergent, then
(a
n
±b
n
) =
a
n
±
b
n
are also convergent.
3. If
a
n
= and k > 0, then
ka
n
= .
4. If
a
n
= and
b
n
converges, then
(a
n
+ b
n
) = .
Proof. Simply apply the limit/divergence laws to the sequence of partial sums. E.g. for 1,
ka
n
= lim
n
n
j=m
ka
j
finite
sum
= lim
n
k
n
j=m
a
j
limit
laws
= k lim
n
n
j=m
a
j
= k
a
n
The others may be proved similarly.
Note that series do not behave nicely with respect to multiplication (see also Exercise 3):
a
1
b
1
+ a
2
b
2
+ ··· =
a
n
b
n
=
a
n
b
n
=
a
1
+ a
2
+ ···
b
1
+ b
2
+ ···
28
It is common to denote a series
a
n
if the initial term is understood (typically a
0
or a
1
), or is irrelevant to the situation.
Series which may be evaluated exactly
Our major goal is to develop techniques for answering the binary question, “Does
a
n
converge?”
Even when the answer is yes, a precise computation of the limit is usually beyond us. However,
our techniques (the upcoming series tests) will typically rely on comparing
a
n
to a ‘standard’ series
whose properties are completely understood. You have met two such series in elementary calculus.
Definition 14.3 (Geometric series). A sequence (a
n
) is geometric if the ratio of successive terms is
constant: a
n
= ba
n
for some constants a, b. A geometric series is the sum of a geometric sequence.
The computation of the sequence of partial sums should be familiar (for simplicity assume b = 1)
(1 a)s
n
=
a
m
+ a
m+1
+ ··· + a
n
a
m+1
+ a
m+2
+ ··· + a
n
+ a
n+1
= a
m
a
n+1
from which we quickly conclude:
Theorem 14.4. If a is constant, then
s
n
=
n
k=m
a
k
=
a
m
a
n+1
1 a
if a = 1
n + 1 m if a = 1
=
n=m
a
n
converges to
a
m
1 a
if
|
a
|
< 1
diverges to if a 1
diverges by oscillation if a 1
In particular,
a
n
converges absolutely if
|
a
|
< 1 and diverges otherwise.
Examples 14.5. 1.
n=1
2
4
5
n
= 2
4
5
1
1 +
4
5
=
5
2
·
5
9
=
25
18
2. Consider the series
a
n
=
n=3
2
5
n
+ 2
n
. If this were convergent, then
2
n
=
a
n
2
5
n
would converge (Theorem 14.2); a contradiction.
Telescoping series A rarer type of series can be evaluated using the algebra of partial fractions.
Example 14.6. To compute
n=1
1
n(n+1)
, first observe that
s
n
=
n
k=1
1
k(k + 1)
=
n
k=1
1
k
1
k + 1
=
1
1
1
2
+
1
2
1
3
+ ··· +
1
n
1
n + 1
= 1
1
n + 1
It follows that
n=1
1
n( n + 1)
= lim
1
1
n + 1
= 1
Similar arguments can be made for other series such as
1
n(n+2)
.
51
The Cauchy Criterion
The starting point for general series convergence uses Cauchy completeness.
Example 14.7. Consider again the series
1
n
2
. We show that the sequence of partial sums (s
n
) is
Cauchy. Let ϵ > 0 be given and let N =
1
ϵ
. Then,
m > n > N =
|
s
m
s
n
|
=
m
k=n+1
1
k
2
<
m
k=n+1
1
k(k 1)
=
m
k=n+1
1
k 1
1
k
=
1
n
1
m
<
1
N
= ϵ
where we follow the telescoping series approach to cancel most terms. By Cauchy completeness
(Theorem 10.11), (s
n
) converges and we conclude
The series
1
n
2
is convergent
Computing the value of this series rigorously is significantly harder, though a sketch argument for
why
1
n
2
=
π
2
6
is in Exercise 10.
Theorem 14.8 (Cauchy Criterion for Series). A series
a
n
converges if and only if
ϵ > 0, N such that m n > N =
|
s
m
s
n1
|
=
m
k=n
a
k
< ϵ
In the previous example we essentially verified the Cauchy criterion for the series
1
n
2
.
Proof. Let (s
n
) be the sequence of partial sums. Then
a
n
converges (s
n
) converges (s
n
) is Cauchy (Thm 10.11)
ϵ > 0, N such that m > n > N =
|
s
m
s
n
|
< ϵ
To finish, simply replace n with n 1.
Example 14.9. Assume, for contradiction, that the harmonic series
1
n
converges. Now take ϵ =
1
2
in
the Cauchy criterion:
N such that m n > N =
m
k=n
1
k
<
1
2
However, taking m = 2(n 1) n (true since n > N 1) results in a contradiction:
1
2
>
m
k=n
1
k
=
1
n
+ ··· +
1
m
m (n 1)
m
= 1
n 1
m
=
1
2
We conclude that the harmonic series diverges to .
52
The Series Tests
For the remainder of this section we develop several standard tests for the convergence/divergence
of an infinite series: the divergence, comparison, root and ratio tests. The first of these follow quickly
from the Cauchy criterion.
Theorem 14.10 (Divergence/n
th
-term test). If lim a
n
= 0 then
a
n
is divergent.
Proof. We prove the contrapositive. Suppose
a
n
is convergent, let ϵ > 0 be given and take m = n in
the Cauchy criterion. Then
N such that n > N =
|
a
n
|
< ϵ
Otherwise said, lim a
n
= 0.
Examples 14.11. 1. The series
sin(
nπ
9
) diverges.
2. The divergence test tells us that the geometric series
a
n
diverges whenever
|
a
|
1. We still
need our earlier analysis for when
|
a
|
< 1.
3. The converse of the n
th
-term test is false! For the canonical example, consider the divergent har-
monic series
1
n
(Example 14.9, even though lim
1
n
= 0.
Theorem 14.12 (Comparison test). 1. Let
b
n
be a convergent series of non-negative terms and
assume
|
a
n
|
b
n
for all (large n). Then both
b
n
and
|
b
n
|
are convergent.
2. If
a
n
= and a
n
b
n
for all (large) n, then
b
n
= .
Proof. Suppose “large n means n > M.
1. Let ϵ > 0 be given. Then N M such that
m n > N =
m
k=n
a
k
m
k=n
|
a
k
|
m
k=n
b
k
< ϵ
2. The n
th
partial sum of
b
n
is
n
k=M
b
k
n
k=M
a
k
+
Corollary 14.13. 1. Take
|
a
n
|
= b
n
in part 1 to see that
|
a
n
|
converges =
a
n
converges. Thus
absolute convergence implies convergence.
2. If
b
n
is a convergent series of non-negative terms and
|
a
n
|
b
n
for all n, then
a
n
|
a
n
|
b
n
53
Examples 14.14. 1. Since
2n+1
(n+2)3
n
2 · 3
n
and the geometric series
2 ·3
n
converges, we see that
the resulting series converges (absolutely), to some value
n=0
2n + 1
( n + 2)3
n
2
n=0
3
n
=
2
1
1
3
= 3
2. One can usually find a sensible series to compare with just by thinking about how a
n
behaves
when n is very large. For instance, a
n
=
(n
2
+1)
1/2
(1+
n)
4
behaves like
n
n
2
=
1
n
and we see that
a
n
>
n
(1 +
n)
4
>
n
(2
n)
4
=
1
16n
Comparison with
1
16
1
n
shows that
a
n
diverges to .
3. Since ln n < n =
1
ln n
>
1
n
, we see that
1
ln n
diverges to by comparison with
1
n
.
4.
sin n
n
2
converges absolutely by comparison to
1
n
2
. Corollary 14.13 gives an estimation for the
value of the series, though it is not accurate! The n
th
partial sums satisfy
n=1
sin n
n
2
n=1
|
sin n
|
n
2
n=1
1
n
2
=
π
2
6
(approximately 1.014 1.280 1.645)
5. The alternating harmonic series
n=1
(1)
n+1
n
converges via a sneaky comparison.
Consider the series t =
n=1
1
2n(2n1)
which converges by comparison with
1
4(n1)
2
. Its n
th
par-
tial sum is
t
n
=
n
k=1
1
2k(2k 1)
=
n
k=1
1
2k 1
1
2k
which is the even partial sum of the alternating harmonic series s
2n
=
2n
m=1
(1)
m+1
m
.
Take limits of s
2n+1
= s
2n
+
1
2n+1
, we see that lim s
2n+1
= t from which lim s
n
= t. Since the
harmonic series
1
n
diverges (Example 14.9), we conclude that the alternating harmonic series
converges conditionally. We will revisit this discussion in the next section.
6.
n
n+1
n
2
converges by comparison with the geometric series
2
n
. To see this, note that
n
n + 1
n
=
n + 1
n
1
1
n + 1
n+1
n
e
1
<
1
2
from which we see that, for all large n,
n
n + 1
n
<
1
2
=
n
n + 1
n
2
< 2
n
In fact (compare Exercise 10.10),
n
n+1
n
is monotone-down, whence e
1
n
n+1
n
1
2
and
0.58198
1
e 1
=
e
1
1 e
1
=
n=1
e
n
n=1
n
n + 1
n
2
n=1
2
n
=
1/2
1 1/2
= 1
A computer estimate yields
n=1
n
n+1
n
2
0.8174.
54
Our final two tests in this section are less powerful, but have the advantage of being easier to use.
Theorem 14.15 (Root test). Let lim sup
|
a
n
|
1/n
= L.
1. If L < 1, then
a
n
converges absolutely.
2. If L > 1, then
a
n
diverges.
If L = 1, then no conclusion can be drawn.
We defer the proof until after seeing some examples.
Corollary 14.16 (Ratio test). Suppose (a
n
) is a sequence of non-zero terms.
1. If lim sup
a
n+1
a
n
< 1, then
a
n
converges absolutely
2. If lim inf
a
n+1
a
n
> 1, then
a
n
diverges
Proof. This follows directly from the root test and Theorem 12.3:
lim inf
a
n+1
a
n
lim inf
|
a
n
|
1/n
lim sup
|
a
n
|
1/n
lim sup
a
n+1
a
n
The versions of these tests familiar from elementary calculus are the special cases when
L = lim
|
a
n
|
1/n
= lim
a
n+1
a
n
Our versions are more general since these limits are not guaranteed to exist.
Examples 14.17. 1. The ratio test is particularly useful for series involving factorials and exponentials.
(a)
n
4
2
n
converges: just observe that lim
a
n+1
a
n
= lim
(n+1)
4
2n
4
=
1
2
< 1.
(b)
n!
2
n
diverges: in this case lim
a
n+1
a
n
= lim
(n+1)!
2n!
= lim
n+1
2
= .
2. Both tests are inconclusive for rational sequences: if a
n
=
b
n
c
n
where b
n
, c
n
are polynomials, then
lim
a
n+1
a
n
= 1 = lim
|
a
n
|
1/n
For example,
n + 5
n
2
lim
a
n+1
a
n
= lim
( n + 6)n
2
( n + 5)(n + 1)
2
= 1
In fact this example is divergent by comparison with the harmonic series
1
n
.
3. Recall Example 14.14.6: our use of the comparison test was really the root test in disguise
a
n
=
n
n + 1
n
2
= lim
|
a
n
|
1/n
= lim
n
n + 1
n
= e
1
< 1 =
a
n
converges
In this case the root test was much easier to apply!
55
4. The ratio test is the weakest test thus far; certainly it does not apply if any of the terms a
n
are
zero! For a more subtle example, consider:
a
n
=
(
2
n
if n is even
3
n
if n is odd
First we try applying the ratio test:
a
n+1
a
n
=
(
1
3
2
3
n
if n is even
1
2
3
2
n
if n is odd
= lim inf
a
n+1
a
n
= 0, lim sup
a
n+1
a
n
=
The ratio test is therefore inconclusive. However
|
a
n
|
1/n
=
(
1
2
if n is even
1
3
if n is odd
= lim sup
|
a
n
|
1/n
=
1
2
< 1
By the root test, the series
a
n
converges. We need not even have used the root test:
a
n
converges by comparison with
2
n
!
Proof of the Root Test. 1. Suppose lim sup
|
a
n
|
1/n
= L < 1. Recall that v
N
= sup{
|
a
n
|
1/n
: n > N}
defines a monotone-down sequence converging to L. Choose any ϵ > 0 such that L + ϵ < 1 (say
ϵ =
1L
2
) to see that
N such that v
N
L < ϵ
But then
n > N =
|
a
n
|
1/n
L < ϵ =
|
a
n
|
< (L + ϵ)
n
|
a
n
|
therefore converges by comparison with the geometric series
(L + ϵ)
n
.
2. If L > 1 then there exists some subsequence (a
n
k
) such that
|
a
n
k
|
1/n
k
L > 1. In particular,
infinitely many terms of this subsequence must be greater than 1. Clearly a
n
does not converge
to zero whence the series diverges by the n
th
term test.
Summary The logical flow of the tests in this section is as follows:
(divergence tests) n
th
term
+3
Root
+3
Ratio
(testing both) Definition of
a
n
ks +3
Cauchy Criterion
KS
+3
Comparison
(convergence tests) Root
+3
Ratio
The ratio test is typically the easiest to use, but the least powerful. Every series which converges by
the ratio test can be seen to converge by the root and comparison tests, etc. If a series diverges by the
ratio test, then it in fact diverges by the n
th
term test.
56
Exercises 14. 1. Determine which of the following sequences converge. Justify your answers.
(a)
n1
n
2
(b)
( 1)
n
(c)
3
n
n
3
(d)
n
3
3
n
(e)
n
2
n!
(f)
1
n
n
(g)
n
2
n
(h)
n!
n
n
(i)
n=2
1
n+(1)
n
2
(j)
n + 1
n
2. Let
a
n
and
b
n
be convergent series of non-negative terms. Prove that
a
n
b
n
converges.
(Hint: start by showing that
a
n
b
n
a
n
+ b
n
)
3. (a) If
a
n
converges absolutely, prove that
a
2
n
converges.
(b) More generally, if
|
a
n
|
converges and (b
n
) is a bounded sequence, prove that
a
n
b
n
converges absolutely.
4. Find a series
a
n
which diverges by the root test but for which the ratio test is inconclusive.
5. (Hard) Let (a
n
) be a sequence such that lim inf
|
a
n
|
= 0. Prove that there is a subsequence (a
n
k
)
such that
a
n
k
converges.
(Hint: Try to construct a subsequence which converges to zero faster than
1
k
2
.
6. Prove that the harmonic series
1
n
diverges by comparing with the series
a
n
, where
(a
n
) =
1,
1
2
,
1
4
,
1
4
,
1
8
,
1
8
,
1
8
,
1
8
,
1
16
,
1
16
,
1
16
,
1
16
,
1
16
,
1
16
,
1
16
,
1
16
,
1
32
,
1
32
, . . .
7. Suppose b
n
a
n
for all n and that
b
n
and
a
n
converge. Prove that
b
n
a
n
.
(This also proves part 2 of Corollary 14.13)
8. Use the basic series laws to find the values of
1
(2n)
2
,
1
(2n+1)
2
and
(1)
n+1
n
2
.
9. The limit comparison test states:
Suppose
a
n
,
b
n
are series of positive terms and that L = lim
a
n
b
n
(0, ). Then the
series have the same convergence status (both converge or both diverge to ).
(a) Use the limit comparison test with b
n
=
1
n
2
to show that the series
1
n
ln
1 +
1
n
converges.
(Hint: Recall that e = lim
1 +
1
n
n
)
(b) Prove the limit comparison test.
(Hint: first show that
L
2
<
a
n
b
n
<
3L
2
for large n)
(c) What can you say about the series
a
n
and
b
n
if L = 0 or L = ? Explain.
10. Euler asserted that the sine function, written as an infinite polynomial in the form of a Maclau-
rin series, could also be expressed as an infinite product,
sin x =
n=0
( 1)
n
(2n + 1)!
x
2n+1
= x
1
x
2
π
2
1
x
2
4π
2
1
x
2
9π
2
···
By considering the solutions to sin x = 0, give some weight to Euler’s claim. By comparing
coefficients in these expressions, deduce the fact
1
n
2
=
π
2
6
.
(As we’ve presented it, this argument is non-rigorous!)
57
15 The Integral and Alternating Series Tests
In this section we develop two further standalone series tests, both with narrower applications than
our previous tests.
The first a little out of place given that it requires (improper) integration.
29
Theorem 15.1 (Integral test). Let a
n
= f (n), where f is non-
negative, non-increasing and integrable on [1, ). Then
n=1
a
n
converges
Z
1
f (x) dx converges
in which case
Z
1
f (x) dx
n=1
a
n
a
1
+
Z
1
f (x) dx
The statement is easily modified if the initial term is a
m
.
0
f (x)
0 1 2 3 4 5 6
x
n
a
1
a
2
a
3
a
4
a
5
Proof. We need only interpret the picture:
Z
n+1
1
f (x) dx
n
k=1
a
k
= s
n
= a
1
+
n
k=2
a
k
a
1
+
Z
n
1
f (x) dx ()
Taking limits gives the result.
Even for divergent sums, () allows us to estimate s
n
and analyze its rate of growth. For greater accu-
racy, explicitly evaluate the first few terms and use a modified integral test to estimate the remainder.
The big application of the integral test is a complete description of the convergence status of p-series,
another useful collection of series to which others may be compared.
Corollary 15.2 (p-series). Let p > 0. The series
1
n
p
converges if and only if p > 1.
Examples 15.3. 1.
1
n
3
converges (it is a p-series with p > 1). For a simple estimate, observe that
Z
1
1
x
3
dx = lim
b
1
2
x
2
b
1
=
1
2
=
1
2
n=1
1
n
3
3
2
This is a poor estimate, especially the lower bound. For a quick improvement, we could explic-
itly evaluate the first term and re-run the test starting at n = 2:
1 +
Z
2
1
x
3
dx
n=1
1
n
3
1 +
1
8
+
Z
2
1
x
3
dx = 1 +
1
8
n=1
1
n
3
1 +
1
4
If greater accuracy is required, more terms can be explicitly evaluated.
29
Which in turn requires limits of functions:
R
1
f (x) dx := lim
b
R
b
1
f (x) dx. Even though we haven’t developed these
concepts, the relevant computations should be familiar from elementary calculus.
58
2. In Example 14.9, we used the Cauchy criterion to show that the harmonic series diverges to .
The integral test makes this much easier! The integral test also allows us to estimate how many
terms are required for the partial sum to s
n
to reach a certain threshold, say 10. Since
ln(n + 1) =
Z
n+1
1
1
x
dx
n
k=1
1
k
1 +
Z
n
1
1
x
dx = 1 + ln n
we see that
s
n
10 = ln(n + 1) 10 1 + ln n = e
9
n e
10
1 = 8104 n 22025
Somewhere between 8 and 22 thousand terms are required! The harmonic series diverges to
infinity, but it does so very slowly.
3. The integral test shows that
n=2
1
n ln n
= and moreover that, to exceed 10, somewhere be-
tween 10
3223
and 10
6631
terms are required!
4. The series
2n+1
4n
3
1
diverges to by comparison with the p-series
1
n
.
Alternating Series and Conditional Convergence
Our final test is the only one capable of detecting conditional
convergence, the canonical example of which is the alternat-
ing harmonic series (recall Example 14.14.5). With an eye on
generalization, we re-index so that the first term is a
0
= 1:
s =
n=0
( 1)
n
n + 1
= 1
1
2
+
1
3
1
4
+ ···
=
n=0
( 1)
n
a
n
= a
0
a
1
+ a
2
a
3
+ ···
0
10 20
s
n
(1)
n
a
n
s
+
n
s
n
The alternating ±-signs give the series its name. For us, however, the behavior of the sequence (s
n
)
of partial sums is more interesting. Consider two subsequences (s
+
n
) = (s
2n
) and (s
n
) = (s
2n1
):
s
+
n
=
2n
k=0
( 1)
k
a
k
= 1
1
2
1
3
|{z}
a
1
a
2
1
2
1
3
|{z}
a
3
a
4
···
1
2n
1
2n + 1
| {z }
a
2n1
a
2n
(n 0)
s
n
=
2n1
k=0
( 1)
k
a
k
=
1
1
2
|{z}
a
0
a
1
+
1
3
1
4
|{z}
a
2
a
3
+ ··· +
1
2n 1
1
2n
| {z }
a
2n2
a
2n1
(n 1)
Since the brackets are non-negative, (s
+
n
) is monotone-down and (s
n
) monotone-up. Moreover,
1
2
= s
1
s
n
s
n
+ a
2n
= s
+
n
s
+
0
= 1 (†)
from which both subsequences are bounded and thus convergent. Not only this, but
lim
s
+
n
s
n
= lim a
2n
= 0
shows that the limits of both subsequences are identical (of course both are s).
59
The above discussion depended only on two simple properties of the sequence (a
n
); we’ve therefore
proved a general statement.
Theorem 15.4 (Alternating series test). Let (a
n
) be monotone-down with lim a
n
= 0.
1. The series
( 1)
n
a
n
converges.
2. If (s
n
) is the sequence of partial sums converging to s =
( 1)
n
a
n
, then
|
s s
n
|
a
n+1
.
Think about where our assumptions on (a
n
) are used in the proof. It can, in fact, be shown that
the alternating harmonic series converges to ln 2, although the estimates provided by the alternating
series test make this a terrible method of approximation. Even summing the first 100 terms only
results in 2 decimal places of accuracy!
Examples 15.5. 1. Since a
n
=
1
n!
converges monotone-down to zero, the alternating series
(1)
n+1
n!
converges. By taking the first 9 and 10 terms of this series, we see that
0.9010498898 . . .
n=1
( 1)
n+1
n!
0.90105016538 . . .
which at least yields the estimate 0.9015 to 5 decimal places. The alternating series test is not
required for this example, since it in fact converges absolutely.
2. The series
n=2
sin
π
2
n
ln n
can be viewed as an alternating series since every even term is zero. Writ-
ing m = 2n + 1, we obtain
n=2
sin
π
2
n
ln n
=
m=1
sin(πm +
π
2
)
ln( 2m + 1)
=
m=1
( 1)
m
ln( 2m + 1)
Since
1
ln(2m+1)
decreases to zero, the alternating series test shows convergence.
Rearranging Infinite Series
A rearrangement of an infinite series
a
n
arises when we change the order of the terms of the sequence
(a
n
) before computing the sequence of partial sums. We still have to use every term a
n
in the new
series. Since the resulting sequence of partials sums is likely completely different, we shouldn’t
assume that the new series has the same convergence properties as the old.
Example 15.6. We rearrange the alternating harmonic series by summing two positive terms before
each negative term:
1 +
1
3
1
2
+
1
5
+
1
7
1
4
+
1
9
+
1
11
1
6
+ ··· +
1
4n 3
+
1
4n 1
1
2n
+ ···
Every term in the original sequence is used here, so this is a genuine rearrangement. It is perhaps
surprising to discover that the new series converges, though its limit is not the same as the original
alternating harmonic series! We leave the details to Exercise 9.
60
This behavior is quite different to that of finite sums, where the order of summation makes no differ-
ence at all. The situation can be summarized in a famous result of Riemann.
Theorem 15.7 (Riemann rearrangement). 1. If
a
n
is conditionally convergent and s R {±},
then there exists a rearrangement of the series which tends
30
to s.
2. If
a
n
converges absolutely, then all rearrangements converge to the same limit.
The second part says that absolutely convergent series behave just like finite sums! We omit the
proofs since they are lengthy and require nasty notation. Instead we illustrate the rough idea of part
1 via an example.
Example 15.8. We show how to construct a rearrangement of the alternating harmonic series which
converges to s =
2 1 = 0.41421 . . .
First we convince ourselves that the sum of the positive terms
a
+
n
diverges to infinity. In this case
the comparison test comes to our rescue:
1
2n 1
>
1
2n
=
a
+
n
=
1
2n 1
>
1
2
1
n
=
The negative terms also diverge
a
n
= . Construction of the rearrangement is inductive.
1. Sum just enough positive terms until the partial sum exceeds s: plainly S
1
= 1 will do.
2. Sum negative terms starting at the beginning of the sequence until the sum is less than s:
S
2
= 1
1
2
1
4
= 0.25 < s
3. Repeat: add positive terms until the sum just exceeds s, then add negative terms, etc.,
S
3
= S
2
+
1
3
= 0.583 . . . > s, S
4
= S
3
1
6
1
8
= 0.291 . . . < s
Continuing the process ad infinitum, we claim that
s = 1
1
2
1
4
+
1
3
1
6
1
8
+
1
5
1
10
+
1
7
1
12
1
14
+
1
9
1
16
1
18
+
1
11
1
20
+
1
13
···
To see why, observe:
Since
a
+
n
= and
a
n
= , at each stage we only add/subtract finitely many terms.
All terms of the original sequence (a
n
) are eventually used since we add the positive (negative)
terms in order. E.g., a
495
=
1
495
appears, at the latest, during the 495
th
positive-addition phase.
|
S
n
s
|
|
a
m
n
|
, where a
m
n
is the last term used at the n
th
stage. This plainly converges to zero,
whence lim S
n
= s.
30
Riemann’s result is in fact even stronger. Rearrangements also exist which diverge by oscillation between any given
lim inf and lim sup!
61
Exercises 15. 1. Use the integral test to determine whether the series
n=1
1
n
2
+1
converges or diverges.
2. Prove Corollary 15.2 regarding the convergence/divergence of p-series.
3. Let s
n
=
n
k=1
1
k
. Estimate how many terms are required before s
n
100.
4. (Example 15.3.3) Verify the claim that
n=2
1
n ln n
= . If you want a challenge, verify the estimate
claim also.
5. (a) Give an example of a series
a
n
which converges, but for which
a
2
n
diverges.
(Exercise 14.3 really requires that
a
n
be absolutely convergent!)
(b) Give an example of a divergent series
b
n
for which
b
2
n
converges.
6. Suppose (a
n
) satisfies the hypotheses of the alternating series test except that lim a
n
= a > 0.
What can you say about the sequences (s
+
n
) and (s
n
) and the series
( 1)
n
a
n
?
7. We know that the harmonic series has a growth rate comparable to ln n. Let a
n
=
1
n
and define
a new sequence (t
n
) by
t
n
= s
n
ln n = 1 +
1
2
+ ··· +
1
n
ln n
where s
n
=
n
k=1
a
n
is the n
th
partial sum. Prove that (t
n
) is a positive, monotone-down sequence,
which therefore converges.
31
(Hint: you’ll need the mean value theorem from elementary calculus)
8. (a) Show that the series
n=1
(1)
n
n
n
2
+1
is conditionally convergent to some real number s.
(b) How many terms are required for the partial sum s
n
to approximate s to within 0.01.
(c) Following Example 15.8, use a calculator to state the first 15 terms in a rearrangement of
the series in part (a) which converges to 0.
9. In Example 15.6 we rearranged the terms of the alternating harmonic series by taking two pos-
itive terms before each negative term.
(a) Verify, for each n N, that
b
n
:=
1
4n 3
+
1
4n 1
1
2n
> 0
whence the subsequence of partial sums (s
3n
) is monotone-up.
(b) Use the comparison test to show that
b
n
converges.
(c) Prove that the rearranged series converges, to some value s >
5
6
.
(Thus s > ln 2 0.69, the limit of the original alternating harmonic series)
31
The limit γ := lim t
n
0.5772 is the Euler–Mascheroni constant. It appears in many mathematical identities, and yet
very little about it is known; it is not even known whether γ is irrational!
62
17 Continuous Functions
For the rest of the course, we discuss continuous functions. Functions themselves should be familiar.
For reference, we begin with a review of some basic concepts and conventions.
We are concerned with functions f : U V where both U, V are subsets of the real numbers R and
f is some rule assigning to each real number x U a real number f (x) V. For instance
f (x) =
x
2
(x 7)
(x 2)(x
2
9)
assigns to x = 1 the value f (1) =
1(6)
( 1)(8)
=
3
4
Domain dom f = U is the set of inputs to f . When f is defined by a formula, its implied domain is the
largest set on which the formula is defined: for the above example, dom f = R \ {2, 3, 3}. In
examples, the domain is typically a union of intervals of positive length.
Codomain codom f = V is the set of possible outputs. In real analysis, we often take V = R by default.
Range range f = f (U) = {f (x) : x U}; is the set of realized outputs. It is a subset of V.
Injectivity f is injective/one-to-one if distinct inputs produce distinct outputs. This is usually stated in
the contrapositive: f (x) = f (u) = x = u.
Surjectivity f is surjective/onto if every possible output is realized: otherwise said, f (U) = V.
Inverses f is bijective/invertible if it is both injective and surjective. Equivalently, f has an inverse
function f
1
: V U defined as follows:
Given y V, f surjective = x U such that f (x) = y.
Since f is injective, f (x) = f (u) = x = u, so x is unique. We define f
1
( y) = x.
Example 17.1. The function defined by f (x) =
1
x(x2)
has implied
dom f = R \{0, 2} = (, 0) (0, 2) (2, )
range f = ( , 1] (0, )
The function is neither injective (e.g., f (3) = f (1)) nor surjective
(e.g., 0 range f ).
We can remedy both issues by restricting the domain and codomain.
For instance, the same rule/formula but with
dom
ˆ
f = [1, 2) (2, )
codom
ˆ
f = ( , 1] (0, )
defines a bijection with inverse function
ˆ
f
1
( y) =
(
1 + y
1
p
y + 1 if y > 0
1 y
1
p
y + 1 if y 1
Observe that dom
ˆ
f
1
= codom
ˆ
f and codom
ˆ
f
1
= dom
ˆ
f .
2
1
1
2
f (x)
1 1 2 3
x
2 1 0 1 2
1
2
3
ˆ
f
1
(y)
y
63
To introduce continuity, we consider two common na
¨
ıve notions.
The graph of f can be drawn without removing one’s pen from the page This is intuitive but un-
usable: drawn is poorly defined, so how could we calculate or prove anything with this concept?
Moreover, it cannot reasonably be extended to other situations (e.g., multivariable functions)
where drawing a graph is meaningless.
If x is close to a, then f (x) is close to f (a) This is better and can be generalized to other situations.
The major issue is the unclear meaning of close. Our formal definition of continuity addresses
this using sequences and limits.
Definition 17.2 (Sequential continuity). Let f : U V R be a function. We say that f is
continuous at u U if,
(x
n
) U, we have lim x
n
= a = lim f (x
n
) = f (a)
f is continuous (on U) if it is continuous at every point a U.
A discontinuity of f is a value a U at which f is discontinuous (not continuous),
(x
n
) U, such that lim x
n
= a and
f (x
n
)
does not converge to f (a)
x
1
f (x
1
)
x
2
f (x
2
)
(a, f (a))
···
.
.
.
lim f (x
n
)
a
Continuity at a: every sequence
with lim x
n
= a has lim f (x
n
) = f (a)
(a, f (a))
x
1
f (x
1
)
x
2
f (x
2
)
···
.
.
.
lim f (x
n
)
f (a)
a
Discontinuity at a: at least one sequence
with lim x
n
= a has lim f (x
n
) = f (a)
Examples 17.3. 1. f : R R : x 7 x
2
is continuous (at every a R). To see this, suppose (x
n
)
converges to a, then, by the limit laws,
lim f (x
n
) = lim x
2
n
= (lim x
n
)
2
= a
2
= f (a)
2. The function with g(x) = 1 +
4
x
2
is continuous. Choose any a dom g = R \ {0} and any
(x
n
) dom g with lim x
n
= a. Again, by the limit laws,
lim g(x
n
) = lim
1 +
4
x
2
n
= 1 +
4
(lim x
n
)
2
= 1 +
4
a
2
= f (a)
This example (with a = 1 and x
n
= 1 +
2
n
) is the first picture in the definition.
64
3. h : [0, ) R : x 7 3x
1/4
is continuous. Again, everything follows from the limit laws. If
x
n
a where x
n
0 and a 0, then
lim h(x
n
) = lim 3x
1/4
n
= 3(lim x
n
)
1/4
= 3a
1/4
= h(a)
4. The function defined by
k(x) =
(
1 + 2x
2
if x < 1
2 x if x 1
is discontinuous at a = 1. This seems obvious from the picture,
but we need to use the definition. The sequence with x
n
= 1
1
n
converges to 1 from below, whence
lim k(x
n
) = lim
1 + 2
1
2
n
+
1
n
2
= 3 = 1 = k(1)
0
2
k(x)
0 2
x
1
k(1)
3
x
n
Basic examples and combinations of continuous functions
By appealing to the limit laws for sequences (Theorem 9.2), we can combine continuous functions in
natural ways. For instance, if f , g are continuous at a, then
lim x
n
= a = lim f (x
n
) + g(x
n
) = lim f (x
n
) + lim g(x
n
) = f (a) + g(a)
whence f + g is continuous at a. Here is a general summary.
Theorem 17.4. 1. Suppose f , g are continuous and that k is constant. Then the following functions
are continuous (on their domains)
k f ,
|
f
|
, f + g, f g, f g,
f
g
, max( f , g), min( f , g)
2. If n N then the function f : x 7 x
1/n
is continuous on its domain.
3. Compositions of continuous functions are continuous. Specifically, if g is continuous at a and f
is continuous at g(a), then f g is continuous at a.
4. Algebraic functions are continuous (this includes all polynomials and rational functions).
Proof. Parts 1, 2 are the limit laws; for the maximum and minimum, see Exercise 2. For part 3:
lim x
n
= a
g cont
= lim g(x
n
) = g(a)
f cont
= lim f
g(x
n
)
= f
g(a)
Part 4 follows from parts 1, 2 and 3.
Example 17.5. Part 3 of the theorem says that the following algebraic function is continuous
f : (7, ) R : x 7
s
3x
5/2
+ 7x
2
+ 4
(x 7)
1/3
65
Theorem 17.6 (Squeeze theorem). Suppose f (x) g(x) h(x) for all x = a, that f , h are continu-
ous at a, and that f (a) = g(a) = h(a). Then g is continuous at a.
Proof. This is simply the squeeze theorem (8.9) for sequences: if lim x
n
= a, then
f (x
n
) g(x
n
) h(x
n
) = lim g(x
n
) = g(a)
To provide more interesting examples, we state the following without proof.
Theorem 17.7. The common trigonometric, exponential and logarithmic functions are continuous.
It is possible, though slow and ugly to address some of this now, though it is not very profitable. It is
better to define these functions later using power series
32
when their continuity (and differentiabil-
ity/integrability!) come for free.
Examples 17.8. 1. f (x) =
x
sin e
x
is continuous on its domain R \{ln(nπ) : n N
0
}.
2. The function defined by g(x) = x sin
1
x
if x = 0 and g(0) = 0 is
continuous on R. When x = 0, this follows from Theorems 17.4
and 17.7, while at a = 0 we rely on the squeeze theorem:
x = 0 = x x sin
1
x
x
g(x)
x
The ϵδ Definition of Continuity
The sequential definition of continuity uses limits twice. By stating each of these using the ϵ-definition
of limit, we can reformulate continuity without ever mentioning sequences.
To motivate this, consider f (x) = x
2
at a = 2. By continuity, if (x
n
) is a sequence with lim x
n
= 2,
then lim f (x
n
) = 4. We restate each of these using the definition of limit:
(a) (lim x
n
= 2) δ > 0, M such that n > M =
|
x
n
2
|
< δ
(b) (lim x
2
n
= 4) ϵ > 0, N such that n > N =
x
2
n
4
< ϵ
Here is a short argument that shows how (a) (b) (we’ll revisit this formally in a moment).
Assume (a), let ϵ > 0 be given, and define δ = min( 1,
ϵ
5
). Since lim x
n
= 2, M such that
n > M =
x
2
n
4
=
|
x
n
2
||
x
n
+ 2
|
< δ
|
(x
n
2) + 4
|
(by (a))
δ
|
x
n
2
|
+ 4
(-inequality)
< δ(δ + 4) 5δ ϵ ((a) again)
Let N = M to conclude (b).
It turns out not to be very important that (x
n
) be a sequence. In fact we can dispense with it entirely. . .
32
For instance via the Maclaurin series e
x
=
n=0
x
n
n!
, sin x =
n=0
(1)
n
(2n+1)!
x
2n+1
and cos x =
n=0
(1)
n
(2n)!
x
2n
66
Definition 17.9 (ϵδ continuity). A function f : U V R is continuous at a if
33
ϵ > 0, δ > 0 such that (x U)
|
x a
|
< δ =
|
f (x) f (a)
|
< ϵ ()
A discontinuity of f is a value a U for which,
ϵ > 0 such that δ > 0, x U with
|
x a
|
< δ and
|
f (x) f (a)
|
ϵ (†)
f (a) + ϵ
f (a) ϵ
a δ a + δa
x
f (x)
To force f (x)
to live here. . .
. . . it is enough for x
to live here
Continuity at a
a + δa δ
f (a) + ϵ
f (a) ϵ
f (a)
a
x
f (x)
Regardless of δ, there is
always some x here. . .
. . . so that f (x)
does not live here
Discontinuity at a
This is the intuitive interpretation of continuity: if x is close to a, then f (x) is close to f (a); ϵ and δ
are merely measures of closeness. Most mathematicians consider the ϵδ version to be the definition
of continuity. Thankfully, it doesn’t matter which you prefer. . .
Theorem 17.10. The sequential and ϵδ definitions of continuity (17.2 & 17.9) are equivalent.
Examples (17.3, cont). Before seeing the proof, we repeat our earlier examples using the ϵδ defini-
tion. As with ϵN arguments for limits, it is often useful to do some scratch work first.
1. Suppose f (x) = x
2
and a R. Our goal is to control the size of
x
2
a
2
whenever
|
x a
|
is
small. To keep things simple, assume
|
x a
|
< 1, then,
x
2
a
2
=
|
x a
||
x + a
|
=
|
x a
|
(x a) + 2a
|
x a
|
|
x a
|
+ 2
|
a
|
=
|
x a
|
(1 + 2
|
a
|
)
Now let ϵ > 0 be given and define δ = min( 1,
ϵ
1+2
|
a
|
). Then
|
x a
|
< δ =
|
f (x) f (a)
|
=
x
2
a
2
< δ(1 + 2
|
a
|
) ϵ
Thus f is continuous at a. This is simply a general version of the argument on page 66 with all
mention of sequences removed!
33
The bracketed statement x U is often omitted in (), since the implication requires x to be universally quantified. It
is important that x U = dom f rather than merely x R! By contrast, the expression x U in (†) is always written.
67
2. Let g(x) = 1 +
4
x
2
and a = 0. The first challenge is to control
1
x
by staying away from zero: to
do this, we start by insisting that δ
|
a
|
2
. But now,
|
x a
|
< δ =
|
x
|
>
|
a
|
2
=
1
|
x
|
<
2
|
a
|
()
Now consider the required difference; if
|
x a
|
< δ, then
|
g(x) g(a)
|
=
1 +
4
x
2
1
4
a
2
=
4
a
2
x
2
a
2
x
2
=
4
|
a + x
|
a
2
x
2
|
x a
|
4
1
|
a
|
x
2
+
1
a
2
|
x
|
|
x a
|
()
< 4
4
|
a
|
3
+
2
|
a
|
3
!
|
x a
|
=
24
|
a
|
3
δ
Given ϵ > 0, it suffices to let δ = min(
1
2
|
a
|
,
1
24
|
a
|
3
ϵ). Then
|
x a
|
< δ =
|
f (x) f (a)
|
< ϵ.
3. For h(x) = 3x
1/4
, there are two cases. Suppose ϵ > 0 is given.
If a = 0, let δ =
ϵ
3
4
, then
34
|
x a
|
< δ = 0 x < δ =
|
h(x) h(a)
|
= 3x
1/4
< 3δ
1/4
= ϵ
If a > 0, let δ =
1
3
a
3/4
ϵ. Then, if
|
x a
|
< δ, we have
|
h(x) h(a)
|
= 3
x
1/4
a
1/4
=
3
|
x a
|
x
3
4
+ a
1
4
x
2
4
+ a
2
4
x
1
4
+ a
3
4
3
|
x a
|
a
3/4
<
3δ
a
3/4
= ϵ
4. Suppose k is continuous at 1 and let ϵ = 1. Then δ > 0 for which
|
x 1
|
< δ =
|
k(x) k(1)
|
=
|
k(x) 1
|
< 1
= 0 < k(x) < 2
However, if we choose x = max(
1
2
, 1
δ
2
), then
|
x 1
|
δ
2
< δ
and k(x) k(
1
2
) = 1 +
2
2
= 2. Contradiction.
0
1
2
3
0 1 2x
k(x)
2δ
2ϵ
The basic rules for combining continuous functions may also be proved using ϵδ arguments. E.g.,
ϵδ proof of the squeeze theorem. Given ϵ > 0, we know there exist δ
1
, δ
2
> 0 for which
|
x a
|
< δ
1
=
|
f (x) f (a)
|
< ϵ and
|
x a
|
< δ
2
=
|
h(x) h(a)
|
< ϵ
Let δ = min(δ
1
, δ
2
), then
|
x a
|
< δ =
|
g(x) g(a)
|
max
|
f (x) f (a)
|
,
|
h(x) h(a)
|
< ϵ
whence g is continuous at 0.
34
Remember the hidden quantifier:
|
x a
|
< δ for all x dom f = [0, ), thus x 0 for the duration of this example.
68
Several other arguments are in the exercises. Finally, here is the promise proof of equivalence.
Proof of Theorem 17.10. (sequential ϵδ) We prove the contrapositive. Suppose a is an ϵδ discon-
tinuity (†) and let δ =
1
n
. Then there exists x
n
U such that
|
x
n
a
|
<
1
n
and
|
f (x
n
) f (a)
|
ϵ
Repeating for all n N results in a sequence (x
n
) for which lim x
n
= a and lim f (x
n
) = f (a):
otherwise said, a is a sequential discontinuity.
(ϵδ sequential) Assume (), let (x
n
) U and suppose lim x
n
= a; we must prove that
lim f (x
n
) = f (a). Let ϵ > 0 be given so that a suitable δ satisfying () exists. Since lim x
n
= a,
N such that n > N =
|
x
n
a
|
< δ (since x
n
a and δ > 0 is given)
=
|
f (x
n
) f (a)
|
< ϵ (by ())
We conclude that lim f (x
n
) = f (a), as required.
Examples 17.11. We finish with a couple of esoteric examples on the same theme.
1. Let f : R R be the indicator function for the rational numbers:
f (x) =
(
1 x Q
0 x / Q
Suppose f is continuous at a and let ϵ = 1. Then δ such that
|
x a
|
< δ =
|
f (x) f (a)
|
< 1 ()
There are two cases; these rely on the fact that any interval contains both rational and irrational
numbers (Corollary 4.12, etc.).
(a) If a Q, then f (a) = 1. There exists an irrational number x (a δ, a + δ), whence
|
f (x) f (a)
|
=
|
0 1
|
= 1 < 1.
(b) If a / Q, then f (a) = 0. There exists a rational number x (a δ, a + δ) , whence
|
f (x) f (a)
|
=
|
1 0
|
= 1 < 1.
Either way, we have contradicted (). We conclude that f is nowhere continuous.
2. Let g : R R be defined by
g : R R : x 7→
(
x x Q
0 x / Q
Since 0
|
g(x)
|
|
x
|
, the squeeze theorem tells us that g is continuous at x = 0.
Now suppose g is continuous at a = 0 and let ϵ =
|
a
|
. Then δ such that
|
x a
|
< δ =
|
f (x) f (a)
|
<
|
a
|
The same two cases as in the previous example provide contradictions. We conclude that g is
continuous at precisely one point!
69
Exercises 17. 1. Consider the function with f (x) =
1
x
2
+2x3
.
(a) The implied domain of f has the form dom f = (, a) (b, ). Find a and b.
(b) What is the range of f ?
(c) Show that f : (b, ) range f is bijective and compute its inverse function.
(d) Find the inverse function when we instead restrict the domain to (, a).
(e) Briefly explain why f is continuous on its domain.
2. Let f and g be continuous functions at a.
(a) Show that max( f , g) =
1
2
( f + g) +
1
2
|
f g
|
and deduce that max( f , g) is continuous at a.
(b) How might you show continuity of min( f , g)?
3. Use ϵδ arguments to prove the following.
(a) f (x) = x
2
3x is continuous at x = 1
(b) g(x) = x
3
is continuous at x = a.
(c) h : [0, ) R : x 7
x is continuous.
(d) j(x) = 3x
1
is continuous on R \{0}.
4. Prove that x = 0 is a discontinuity of each function: use both definitions of continuity.
(a) f (x) = 1 for x < 0 and f (x) = 0 for x 0.
(b) g(x) = sin(1/x) for x = 0 and g(0) = 0.
5. Suppose f and g are continuous at a. Prove the following using ϵδ arguments.
(a) f g is continuous at a.
(b) If h is continuous at f (a), then h f is continuous at a.
6. Suppose f : U V R is a function whose domain U contains an isolated point a: i.e. r > 0
such that (a r, a + r) U = {a}. Prove that f is continuous at a.
7. In Example 17.11.2, provide the details of the required contradiction.
8. (a) Suppose f : R R is a continuous function for which f (x) = 0 whenever x Q. Prove
that f (x) = 0 for all x R.
(b) Suppose f , g : R R are continuous functions such that f (x) = g(x) for all rational x.
Prove that f = g.
9. (Hard) Consider f : R R where
f (x) =
(
1
q
whenever x =
p
q
Q with q > 0 and gcd(p, q) = 1
0 if x Q
For example, f (1) = f (2) = f (7) = 1, and f (
1
2
) = f (
1
2
) = f (
3
2
) = ··· =
1
2
, etc. Prove that f
is continuous at each irrational number, and discontinuous at each rational number.
70
18 Properties of Continuous Functions
In this section consider how continuous functions transform intervals.
Example 18.1. f (x) = x
2
maps [3, 2] onto [0, 9]. In particular:
f transforms one interval into another.
f transforms one closed bounded set into another.
The purpose of this section is to see that these are general properties
exhibited by any continuous function.
2
4
6
8
f (x)
3 2 1 0 1 2
x
Before stating our first result, recall a couple of definitions.
Definition 18.2. Let U, V R and f : U V.
1. (a) U is bounded if M such that x U,
|
x
|
M.
(b) f is bounded if its range is a bounded set: M such that x U,
|
f (x)
|
M.
2. (Definition 11.13) U is closed if every convergent sequence in U has its limit in U:
(x
n
) U, lim x
n
= s ( R) = s U
Theorem 18.3 (Extreme Value Theorem). Suppose f : U V is continuous where U is closed and
bounded. Then f is bounded and attains its bounds:
s, i U such that f ( s) = sup
f (U)
and f (i) = inf
f (U)
In fact f (U) is also closed and bounded.
In Example 18.1, if U = [3, 2], then s = 3 and i = 0.
Examples 18.4. Before seeing the proof, here are three examples where we weaken one of the hy-
potheses and see that the result fails.
0
1
2
3
0 1 2
0
1
2
3
0 1 2
0
2
4
6
0 1 2
1. f discontinuous 2. U not closed 3. U not bounded
1. sup(range f ) = 3 is not attained by f (x) =
(
3x if 0 x < 1
1 if 1 x 2
2. If f (x) =
1
2x
and U = [0, 2), then range f = [
1
2
, ) is unbounded.
3. If f (x) = x
2
and U = [0, ), then range f = [0, ) is unbounded.
71
The goal of the proof is to show that every limit point of f (U) = range f lies in f (U). The proof is
broken into simple steps: observe where each hypothesis is used.
Proof. 1. Suppose M is a limit point of f (U): that is, M = lim
f (x
n
)
for some sequence (x
n
) U.
A priori M need not be finite, but it is possible
35
that M = sup
f (U)
or inf
f (U)
.
2. Since (x
n
) U is bounded, Bolzano–Weierstraß (Theorem 11.8) says it has a convergent subse-
quence, lim
k
x
n
k
= x.
3. Since U is closed, we have x U. This means f (x) makes sense (it is a real number!).
4. Since f is continuous, we have lim f (x
n
k
) = f (x).
5. Finally, M = f (x) since all subsequences of a convergent (or divergent to ±) sequence tend
to the same limit (Lemma 11.3). It follows that all limit points M are finite and lie in f (U):
otherwise said, f (U) is closed and bounded.
In particular, choosing M = sup
f (U)
yields x = s U as in the Theorem.
Example 18.5. It is worth thinking about why we needed to use a subsequence in the proof. The
reason is that it is possible for the bounds of f to be attained multiple times. For example, consider
f : [0, 4π] R : x 7→ sin x
This satisfies the hypotheses of the extreme value theorem: [0, 4π] is closed and bounded and f is
continuous. Indeed max(range f ) = 1 is attained at both x =
π
2
and
5π
2
. The sequence defined by
x
n
=
(
π
2
+
1
n
if n odd
5π
2
+
1
n
if n even
has f (x
n
) = sin
π
2
+
1
n
n
1 = sup(range f )
and therefore satisfies step 1 of the proof. However, (x
n
) itself is divergent by oscillation. Bolzano–
Weierstraß is used to force the existence of a convergent subsequence; in this case the subsequence of
odd terms (x
n
k
) = (x
2k1
) satisfies the remaining steps of the proof.
1
0
f (x)
x
x
1
f (x
1
)
x
2
π
2
π
3π
2
2π
5π
2
3π
7π
2
4π
1
35
If M = sup
f (U)
, then a suitable (x
n
) might be constructed as follows:
If M R, then for each n N, x
n
U such that M
1
n
< f (x
n
) M (Lemma 4.8).
If M = , then for each n N, x
n
U such that f (x
n
) n.
72
The Intermediate Value Theorem and its Consequences
This result should be familiar from elementary calculus, even if its proof is not! It is also intuitive: if
you climb a hill, then at some point you must be half-way up the hill. . .
Theorem 18.6 (Intermediate Value Theorem (IVT)). Let f be continuous on the interval [a, b] and
let y lie strictly between f (a) and f (b). Then ξ (a, b) such that f (ξ) = y.
Proof. WLOG assume f (a) < y < f (b). Now let S = {x [a, b] : f (x) < y} and define ξ := sup S.
Since S is non-empty (a S) and bounded above
(by b), we see that ξ exists and is finite. It remains
to prove that a < ξ < b and f (ξ) = y.
First choose any (s
n
) S such that
36
lim s
n
= ξ.
Continuity forces lim f (s
n
) = f (ξ). Moreover
f (s
n
) y = f (ξ) y
This also shows that ξ < b.
y
a b
f (a)
f (b)
s
n
x
n
ξ
S
We now play a similar game from the other side: define x
n
:= min(ξ +
1
n
, b), then lim x
n
= ξ and
x
n
> ξ = sup S = x
n
S, whence
f (x
n
) y = f (ξ) = lim f (x
n
) y
This also shows that ξ > a. Putting it all together, we conclude that f (ξ) = y and ξ (a, b).
Note how the value of ξ in the proof is always the largest of potentially several choices.
Examples 18.7. The intermediate value theorem was typically used in elementary calculus to show
the existence of solutions to equations. Here are a couple of examples of this process.
1. We show that the equation x
7
+ 3x = 1 + 4 cos(πx) has a solution.
The trick is to express the equation in the form f (x) = y where f is continuous, then choose
suitable a, b to fit the theorem. In this case,
f (x) = x
7
+ 3x 4 cos(πx) and y = 1
are suitable choices. Now observe that
f (0) = 4 < y and f (1) = 1 + 3 + 4 = 8 > y (i.e., a = 0 and b = 1)
whence ξ (0, 1) such that f (ξ) = 01. Otherwise said, ξ is a solution to the original equation.
The function f is in fact continuous on R, a much larger interval that [a, b], but no matter!
36
Similarly to step 1 of the proof of the Extreme Value Theorem.
73
2. The existence of a root ξ of the (continuous) polynomial
f (x) = x
5
5x
4
+ 150
follows from the intermediate value theorem by observing that
f (0) = 150 > 0 and f (4) = 256 + 150 = 106 < 0
We conclude that such a root exists and that ξ (0, 4).
As the graph suggests, there are other roots (η, ζ), the existence of
which may be shown by also evaluating, say,
f (3) = 798 < 0 and f (5) = 150 > 0
100
100
200
2 2 4
ξ
η
ζ
With an eye on generalizing, consider a slightly different approach. Define two sequences (s
n
)
and (t
n
) via
s
n
:=
f (n)
n
5
= 1
5
n
+
150
n
5
t
n
:=
f (n)
n
5
= 1
5
n
+
150
n
5
Since lim s
n
= 1 and lim t
n
= 1, we see that
a such that s
a
<
1
2
= f (a) = a
5
s
a
<
1
2
a
5
< 0
b such that t
b
>
1
2
= f (b) = b
5
t
b
>
1
2
b
5
> 0
Applying the intermediate value theorem on [a, b] shows the existence of a root.
The second approach in Example 18.5.2 may be applied to prove the general result.
Corollary 18.8. A polynomial function of odd degree has at least one real root.
The proof is an exercise. An even simpler exercise shows the existence of a fixed point for a particular
type of continuous function.
Corollary 18.9 (Fixed Point Theorem). Suppose a and b are finite and that f : [a, b] [a, b] is
continuous. Then f has a fixed point:
ξ [a, b] such that f (ξ) = ξ
As the picture shows, a function could have several fixed points.
This is the first of several fixed-point theorems you’ll meet if your
studies of analysis continue. Many important consequences flow from
these, including a common fractal construction and the standard exis-
tence/uniqueness result for differential equations.
ξ
1
ξ
2
ξ
3
a b
a
b
74
For our final corollary, we first note a straightforward characterization: a set I R is an interval if
a, b I and a < y < b = y I ()
Corollary 18.10 (Preservation of Intervals). Suppose f : U V is continuous where U = dom f is
an interval (of positive length) and V = range f .
1. V is an interval or a point.
2. If f is strictly increasing (decreasing), then:
(a) V is an interval (it has positive length).
(b) f is injective (and thus bijective).
(c) The inverse function f
1
: V U is also continuous and strictly increasing (decreasing).
Example 18.11. In part 1, note that the interval V need not be of the
same type as U. For instance, if f (x) = 10x x
2
, then f maps the open
interval U = (2, 9) to the half-open interval V = (9, 25].
The extreme value theorem, however, guarantees that if U is a closed
bounded interval, then V is also, for instance,
f
[2, 9]
= [9, 25]
0
5
10
15
20
25
0 2 4 6 8
U
V
Proof. 1. If V is not a point, then a, b U such that f (a) < f (b). Let y
f (a), f (b)
; IVT says
ξ between a and b such that y = f (ξ). That is, y range f . By (), V = range f is an interval.
2. (a,b) If f is strictly increasing, then a, b U, a < b = f (a) < f (b). It follows that f is
injective and that V contains at least 2 points; by part 1 it has positive length.
(c) Let y
1
< y
2
where both lie in V, and define x
i
= f
1
( y
i
) for i = 1, 2. Since f is increasing,
x
2
x
1
= y
2
= f (x
2
) f (x
1
) = y
1
is a contradiction. Thus x
1
< x
2
and f
1
is also strictly increasing.
It remains to show that f
1
is continuous at b = f
1
(a). Assume first that a is not an
endpoint of U. Given ϵ > 0 for which [a ϵ, a + ϵ] U, let
δ := min
b f (a ϵ), f (a + ϵ) b
and observe that
|
y b
|
< δ = f (a ϵ) b < y b < f (a + ϵ) b = f (a ϵ) < y < f (a + ϵ)
= a ϵ < y < a + ϵ ( f strictly increasing)
=
f
1
( y) a
< ϵ
If a is an endpoint of U, instead use [a ϵ, a] U or [a, a + ϵ] U and only the corre-
sponding half of the expression δ.
75
Example 18.12. The function f : [0, 2] [0, 4] defined by
f (x) =
(
3
x if 0 x 1
x
2
if 1 < x 2
is continuous and strictly increasing. It therefore has a continuous in-
verse function f
1
: [0, 4] [0, 2].
Compare this with the result from elementary calculus: f
> 0 = f
injective. We cannot apply this here since f is not differentiable!
0
1
2
3
4
f (x)
0 1 2
x
Exercises 18. 1. Give an example of a discontinuous function f : [0, 1] R which is not bounded.
2. Let a < b be given. Give examples of continuous functions g, h : (a, b) R such that:
(a) g is not bounded.
(b) h is bounded but does not attain its bounds.
3. Compute the inverse of the function f in Example 18.12.
4. Let S R and suppose there exists a sequence (x
n
) in S that converges to a number x
0
S.
Show that there exists an unbounded continuous function on S.
5. Prove that x = cos x for some x in (0,
π
2
).
6. Suppose that f is a real-valued continuous function on R and that f (a) f (b) < 0 for some
a, b R. Prove that there exists x between a, b such that f (x) = 0.
7. Suppose that f is continuous on [0, 2] and that f (0) = f (2). Prove that there exist x, y [0, 2]
such that
|
y x
|
= 1 and f (x) = f (y).
(Hint: consider g(x) = f (x + 1) f (x) on [0, 1])
8. (a) Prove the fixed point theorem (Corollary 18.9).
(Hint: If neither a nor b are fixed points, consider g(x) = f (x) x)
(b) Prove Corollary 18.8 for a general odd-degree monic polynomial f (x) = x
2m+1
+
2m
k=0
α
k
x
k
.
9. Consider f : R R where f (x) = x sin
1
x
if x = 0 and f (0) = 0.
(a) Explain why f is continuous on any interval U.
(b) Suppose a < 0 < b and that f (a), f (b) have opposite signs. If y = 0, show that the
intermediate value theorem is satisfied by infinitely many distinct values ξ.
10. (a) Suppose f : U R is continuous and that U =
n
S
k=1
I
k
is the union of a finite sequence (I
k
)
of closed bounded intervals. Prove that f is bounded and attains its bounds.
(b) Let U =
S
n=1
I
n
, where I
n
= [
1
2n
,
1
2n1
] for each n N. Give an example of a continuous
function f : U R which is either unbounded or does not attain its bounds. Explain.
(This is related to the idea that finite unions of closed sets remain closed, but infinite unions need not)
76
19 Uniform Continuity
Suppose f : U V is continuous. By the ϵ -δ definition (17.9),
a U, ϵ > 0, δ(a, ϵ) > 0 such that (x U)
|
x a
|
< δ =
|
f (x) f (a)
|
< ϵ ()
We write δ(a, ϵ) to stress the fact that δ can depend both on the location a and the distance ϵ. The goal
of this section is to understand if/when it is possible to choose δ independently of the location a.
Example 19.1. We start by considering how this desire might be impossible to satisfy.
Consider f (x) = x
2
with domain U = [0, ). Given ϵ > 0
and a
1
U, we can certainly find some
37
δ such that
|
x a
1
|
< δ =
|
f (x) f (a
1
)
|
=
x
2
a
2
1
< ϵ
Visualize what happens if we try to use the same δ for dif-
ferent a
i
: imagine sliding the fixed-width δ-interval along
the x-axis while simultaneously sliding the ϵ-interval verti-
cally. As a
i
increases, the image of the δ-interval eventually
becomes too large for the ϵ-interval to contain:
length
f (a
i
δ, a
i
+ δ)
= (a
i
+ δ)
2
(a
i
δ)
2
= 4a
i
δ
increases unboundedly with a
i
.
0
0
a
2
1
a
1
a
2
2
a
2
a
2
3
a
3
For fixed ϵ, as a increases, the increasing gradient of f means that we need to choose a smaller δ .
By contrast, if we consider the same formula f (x) = x
2
but on a restricted finite domain [0, b], then
any δ that suffices to demonstrate continuity at x = b will also do so everywhere else on [0, b]. We’ll
check this explicitly in a moment.
To formalize things, consider rewriting (), where we additionally assume that δ may be chosen
independently of the location a; this amounts simply to moving the quantifier a U after δ.
Definition 19.2. A function f : U V R is uniformly continuous if
ϵ > 0, δ > 0 such that (x, y U)
|
x y
|
< δ =
|
f (x) f (y)
|
< ϵ (†)
For reasons of symmetry we use y instead of a. Note how δ now depends only on ϵ since it is
quantified before x and y; as previously, the quantifiers for x, y are usually hidden. Note also how
uniform continuity is only relevant on the entire domain U; it makes no sense to speak of uniform
continuity at a point a.
For the sake of tidiness, we make one more observation before seeing some examples.
Lemma 19.3. If f is uniformly continuous on U, then it is continuous on U.
This should be trivial: () is the ϵ-δ continuity of f at y U, for all y simultaneously! The special
feature of the definition is that the same δ works for all y.
37
For instance δ = min(1,
ϵ
1+2a
1
), as we saw on page 67.
77
Examples 19.4. 1. We re-analyze f (x) = x
2
in view of the definition. Recall first that
|
f (x) f (y)
|
=
x
2
y
2
=
|
x y
||
x + y
|
where
|
x y
|
is easily controlled by δ. We consider the behavior of
|
x + y
|
in two cases.
Bounded domain If U = dom f [T, T] for some T > 0, we show that f is uniformly
continuous. This follows because
|
x + y
|
2T is also easily controlled.
Let ϵ > 0 be given and define δ =
ϵ
2T
, then
|
x y
|
< δ =
|
f (x) f (y)
|
< δ · 2T = ϵ
Compare with Example 19.1. Our approach works for this function because the gradient
(and therefore potential discrepancy between x
2
y
2
and x y) is greatest at the endpoints
of the interval. The same approach may not work for other functions!
Unbounded domain We show that f is not uniformly continuous when dom f = [0, ).
For contradiction, assume f is uniformly continuous: let ϵ = 1 and suppose δ > 0 satisfies
the definition. Supposing x y =
δ
2
, we see that
|
x + y
|
= 2y +
δ
2
=
|
f (x) f (y)
|
=
δ
2
2y +
δ
2
= δ
y +
δ
4
> δx
Letting y =
1
δ
(x =
1
δ
+
δ
2
) yields the contradiction
|
f (x) f (y)
|
> 1 = ϵ.
2. Let g(x) =
1
x
; we again consider two domains.
Uniform continuity on [a, b) whenever 0 < a < b .
Let ϵ > 0 be given and let δ = a
2
ϵ. Then,
|
x y
|
< δ =
|
g(x) g(y)
|
=
y x
xy
<
δ
xy
δ
a
2
= ϵ
where the last inequality follows because x, y a.
Non-uniform continuity on (0, b) whenever 0 < b .
As before, let ϵ = 1 and suppose δ > 0 is given. Let
x = min
δ, 1,
b
2
and y =
x
2
Certainly x, y (0, b) and
|
x y
|
=
x
2
δ
2
< δ. However,
|
f (x) f (y)
|
=
1
x
1 = ϵ
g(x)
x
a b
2δ
2ϵ
Think about how ϵ and δ must relate as one slides the intervals in the picture up/down and
left/right. In this case, large values of x, y are not the problem, it’s the vertical asymptote at
zero that causes trouble.
78
General Conditions for Uniform Continuity
For the remainder of this section, we develop a few general ideas related to uniform continuity. The
first is a little out of order since it depends on differentiation and the mean value theorem.
Theorem 19.5. Suppose f is continuous on an interval U (finite or infinite) and differentiable except
perhaps at its endpoints. If f
is bounded, then f is uniformly continuous on U.
Proof. Suppose
|
f
(x)
|
M. Let ϵ > 0 and δ =
ϵ
M
. Then
|
x y
|
< δ =
|
f (x) f (y)
|
=
f
( ξ)
|
x y
|
< Mδ = ϵ
where the existence of ξ between x, y follows from the Mean Value Theorem.
38
Examples 19.6. 1. Compare the arguments in the previous exercise. For instance, if dom f [T, T],
f (x) = x
2
= f
(x) = 2x =
f
(x)
2T
The derivative is bounded, whence f is uniformly continuous on [T, T].
2. Any polynomial is uniformly continuous on any bounded interval.
3. The function f (x) = sin x is uniformly continuous on R since f
(x) = cos x is bounded (by 1).
4. Consider f (x) =
1
x
5
x
2
on ( 1, ). We have
f
(x) =
1
x
2
+
10
x
3
=
f
(x)
11
We conclude that f is uniformly continuous on (1, ).
The approach is often useful when you are asked to show using the definition that a function is
uniformly continuous; provided f
is bounded by M, you may always choose δ =
ϵ
M
to obtain
an argument. For instance, with our function:
Given ϵ > 0, let δ =
ϵ
11
. If x, y (1, ) and
|
x y
|
< δ, then
|
f (x) f (y)
|
=
1
x
1
y
+
5
y
2
5
x
2
=
|
x y
|
5(x + y)
x
2
y
2
1
xy
=
|
x y
|
5
xy
2
+
5
x
2
y
1
xy
< 11
|
x y
|
(-inequality, since x, y > 1)
< 11δ = ϵ
As we’ll see very shortly, the above result isn’t a biconditional: non-differentiable functions and
functions with unbounded derivatives can be uniformly continuous.
38
If x < y then ξ (x, y) such that f
(ξ) =
f (x) f (y)
x y
.
79
Our remaining conditions are variations on a theme: uniform continuity on a bounded interval U is
roughly the same thing as continuity on its closure U (recall Definition 11.13).
Theorem 19.7. A continuous function on a closed bounded domain is uniformly continuous.
Proof. Assume f is continuous, but not uniformly so, on a closed bounded domain U. Then
ϵ > 0 such that δ > 0, x, y U with
|
x y
|
< δ and
|
f (x) f (y)
|
ϵ ( )
Let δ =
1
n
for each n N to obtain sequences (x
n
), (y
n
) U satisfying ().
39
Since (x
n
) U is bounded, Bolzano–Weierstraß says there exists a convergent subsequence (x
n
k
)
which, since U is closed, converges to some x
0
U.
Since
|
x
n
k
y
n
k
|
<
1
n
k
1
k
, we see that lim
k
y
n
k
= x
0
. Finally, the continuity of f contradicts ():
ϵ lim
|
f (x
n
k
) f (y
n
k
)
|
=
|
f (x
0
) f (x
0
)
|
= 0
Both hypotheses are crucial: Examples 19.4 provide counter-examples if either is weakened.
Example 19.8. f (x) =
x is uniformly continuous on [0, 1]. This cannot be concluded from Theorem
19.5, since its derivative f
(x) =
1
2
x
1/2
is unbounded on ( 0, 1).
Our next goal is to develop a partial converse, for which we first need a lemma.
Lemma 19.9. If f is uniformly continuous on U and (x
n
) U is Cauchy, then
f (x
n
)
is also Cauchy.
To apply the result, consider a convergent (Cauchy) sequence in U whose limit is not itself in U.
Example 19.10. Let f (x) =
1
x
be defined on U = (0, ) and consider the sequence defined by x
n
=
1
n
.
This is plainly Cauchy since it converges; note crucially that its limit 0 does not lie in U. Moreover,
lim f (x
n
) = lim n =
f (x
n
)
is not Cauchy, whence f is not uniformly continuous. This is a far simpler argument than
that presented previously!
Proof. Let ϵ > 0 be given. Since f is uniformly continuous,
δ > 0 such that x, y U,
|
x y
|
< δ =
|
f (x) f (y)
|
< ϵ
Now use this δ in the definition of (x
n
) being Cauchy:
40
N such that m, n > N =
|
x
n
x
m
|
< δ =
|
f (x
n
) f (x
m
)
|
< ϵ
Otherwise said,
f (x
n
)
is Cauchy.
39
These arguments should feel familiar: compare this line to the proof of Theorem 17.10 and the rest to Theorem 18.3.
40
The Cauchy condition is important here: we cannot apply the uniform continuity condition directly to a convergent
sequence (
|
x
n
x
|
< δ . . .) if we do not already know that its limit (here x) lies in U!
80
We apply the Lemma to show that a continuous function on a bounded interval is uniformly continu-
ous if and only if has a continuous extension.
Theorem 19.11. Suppose f is continuous on a bounded interval (a, b). Define g : [a, b] R via
g(x) :=
(
f (x) if x (a, b)
lim f (x
n
) whenever (x
n
) (a, b) and lim x
n
= a or b
Then f is uniformly continuous if and only if g is well-defined; in such a case g is automatically
continuous.
Examples 19.12. 1. f (x) = x
2
3x + 4 is uniformly continuous on
( 2, 4) since it has a continuous extension
g : [2, 4] R : x 7→ x
2
3x + 4
It should be obvious what is happening from the picture: to cre-
ate the extension g, we simply fill in the holes at the endpoints of
the graph.
2. The function f (x) =
1
5x
is continuous, but not uniformly, on the
interval ( 0, 5). This follows since
lim f
5
1
n
= lim n =
means we cannot define g(5) unambiguously. Again the picture
is helpful; while we can fill in the hole at the left endpoint (a =
0) , the vertical asymptote at b = 5 means that there is no hole to
fill in and thereby extend the function.
4
8
12
2 0 2 4
0
1
2
3
4
0 1 2 3 4 5
Proof. () Suppose g is well-defined; we leave the claim that it is continuous as an exercise, but by
Theorem 19.7 it is uniformly so. Since f = g on a subset (a, b) dom g, the same choice of δ
will work for f as it does for g: f is therefore uniformly continuous.
() Suppose f is uniformly continuous on (a, b). Let (x
n
), (y
n
) (a, b) be sequences converging to
a. To show that g(a) is unambiguously defined, we must prove that
f (x
n
)
and
f (y
n
)
are
convergent, and to the same limit.
Define a sequence
( u
n
) = (x
1
, y
1
, x
2
, y
2
, x
3
, y
3
, . . .)
Plainly lim u
n
= a since (x
n
) and (y
n
) have the same limit. But then (u
n
) is Cauchy; by Lemma
19.9,
f (u
n
)
is also Cauchy and thus convergent. Since
f (x
n
)
and
f (y
n
)
are subsequences
of a convergent sequence, they also converge to the same (finite!) limit.
The case for g(b) is similar.
81
Examples 19.13. We finish with three related examples of continuous functions f : R \ {0} R;
these will appear repeatedly as you continue to study analysis.
1. f (x) = sin
1
x
is continuous but not uniformly so. To see
this, note that x
n
=
1
(n+
1
2
)π
defines a Cauchy sequence
(lim x
n
= 0), and yet
f (x
n
) = sin
n +
1
2
π = (1)
n
is not Cauchy since it diverges by oscillation.
Consequently, there is no way to extend f to a continuous
function on any interval containing x = 0.
1
1
x
2
π
1
π
2
π
1
π
2. f (x) = x sin
1
x
is uniformly continuous. One way to see
this is to extend the function to the origin by defining
g(x) =
(
x sin
1
x
if x = 0
0 if x = 0
By the squeeze theorem, lim x
n
= 0 = lim f (x
n
) = 0,
so g is well-defined and continuous on R. By Theorem
19.11, f is uniformly continuous on any bounded inter-
val. Moreover, the derivative
1
x
2
π
2
π
2
π
f
(x) = sin
1
x
1
x
cos
1
x
is bounded whenever x is large; together with Exercise 6 we could use this to conclude uni-
form continuity of f (x). Note however that f
(x) is unbounded when x small (lim f
1
2πn
=
lim(2πn) = ) so we can’t use Theorem 19.5 to conclude that f is uniformly continuous on
its entire domain.
3. f (x) = x
2
sin
1
x
is also uniformly continuous: again extend
by g(0) = 0. This time however, we could argue that the
derivative is bounded
f
(x)
=
2x sin
1
x
cos
1
x
3
since
|
sin y
|
|
y
|
for all y.
In fact something stranger is going on. As you may ver-
ify (see Exercise 3), the extended function g is everywhere
differentiable with g
(0) = 0, and yet the derivative g
(x)
itself is discontinuous at x = 0!
1
1
f (x)
x
2
π
2
π
f
(x)
82
Exercises 19. 1. Decide whether each function is uniformly continuous on the given interval.
Explain your answers.
(a) f (x) = x
3
on [2, 4] (b) f (x) = x
3
on (2, 4)
(c) f (x) = x
3
on ( 0, 4] (d) f (x) = x
3
on ( 1, 4]
(e) f (x) = e
x
on ( , 100) (f) f (x) = e
x
on R
2. Prove that each function is uniformly continuous on the indicated domain by verifying the ϵ-δ
property.
(a) f (x) = 3x + 11 on R (b) f (x) = x
2
on [0, 3]
(c) f (x) =
1
x
2
on [
1
2
, ) (d) f (x) =
x+2
x+1
on [0, 1]
3. Verify the claim in Example 19.13.3 that the function g(x) is differentiable at zero
41
but that the
derivative g
(x) is discontinuous there.
4. (a) If f is uniformly continuous on a bounded set U, prove that f is bounded on U.
(Hint: for contradiction, assume (x
n
) U for which
|
f (x
n
)
|
. . .)
(b) Use (a) to give another proof that
1
x
2
is not uniformly continuous on ( 0, 1).
(c) Give an example to show that a uniformly continuous function on an unbounded set U
could be unbounded.
5. Suppose g is defined on U and a U. Give very brief (one line!) arguments for the following.
(a) Prove that g is continuous at a provided
ϵ > 0, δ > 0 such that 0 <
|
x a
|
< δ =
|
g(x) g(a)
|
< ϵ
(b) Prove that g is continuous at a provided
(x
n
) U \ {a}, lim x
n
= a = lim g(x
n
) = g(a)
(c) Verify that the function g defined in Theorem 19.11 is indeed continuous whenever it is
well-defined.
6. (a) Suppose f is uniformly continuous on intervals U
1
, U
2
for which U
1
U
2
is non-empty.
Prove that f is uniformly continuous on U
1
U
2
.
(Hint: if x, y do not lie in the same interval U
i
, choose some a U
1
U
2
between x and y)
(b) Prove that f (x) =
x is uniformly continuous on [0, ).
(c) More generally, prove that any root function f (x) = x
1/n
(n N) is uniformly continuous
on its domain (R if n is odd and [0, ) if n is even).
(d) (Hard) Given f (x) = x
1/n
, show that δ = ϵ
n
demonstrates uniform continuity when n is
even and δ =
ϵ
2
n
when n is odd.
(Hint: use the binomial theorem to prove that 0 y < x + δ = y
1/n
< x
1/n
+ δ
1/n
)
41
Use the definition g
(0) = lim
x0
g(x)g(0)
x0
. Limits of functions are covered formally in the next section (course!), but you
should be familiar with the idea from elementary calculus.
83