mth303_notes_2024
mth303_notes_2024
Semester 1, 2024-2025
Prahlad Vaidyanathan
Contents
I. The Real Numbers √ 4
1. Introduction: Irrationality of 2. . . . . . . . . . . . . . . . . . . . . . . 4
2. Ordered Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3. The Completeness Axiom . . . . . . . . . . . . . . . . . . . . . . . . . . 7
4. Properties of R . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
III. Topology of R 24
1. Open Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2. Closed Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3. The Closure of a Set . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4. Compact Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
IV.Continuous Functions 32
1. Limits of Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2. Continuous Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3. Continuous Functions on Compact Sets . . . . . . . . . . . . . . . . . . . 39
4. The Intermediate Value Theorem . . . . . . . . . . . . . . . . . . . . . . 40
5. Uniformly Continuous Functions . . . . . . . . . . . . . . . . . . . . . . . 43
V. Differentiation 47
1. Definition and Basic Properties . . . . . . . . . . . . . . . . . . . . . . . 47
2. The Mean Value Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3. Extrema and Curve Sketching . . . . . . . . . . . . . . . . . . . . . . . . 52
4. Taylor’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
VI.Integration 59
1. Lower and Upper Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . 59
2. Integrable Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
3. Properties of the Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4. The Fundamental Theorem of Calculus . . . . . . . . . . . . . . . . . . . 75
5. The Logarithm and Exponential Functions . . . . . . . . . . . . . . . . . 80
2
VII.Sequences of Functions 85
1. Pointwise and Uniform Convergence . . . . . . . . . . . . . . . . . . . . . 85
2. Distances between Functions . . . . . . . . . . . . . . . . . . . . . . . . . 87
3. Weierstrass’ Approximation Theorem . . . . . . . . . . . . . . . . . . . . 89
VIII.
Series of Functions 95
1. Series of Real Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
2. Tests of Convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
3. Power Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
4. Taylor’s Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
5. Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
3
I. The Real Numbers
√
1. Introduction: Irrationality of 2
Notation:
• N = {1, 2, 3, . . .}
• N0 = {0, 1, 2, . . .}
• Q = {p/q : p, q ∈ Z, q 6= 0}
• R =? The set of real numbers does not admit such an easy definition.
p2 = 2q 2 (I.1)
So 2 | p2 , and so 2 | p (this is Euclid’s Lemma - you will see a proof in Lemma 2.6 of
[MTH301]). So we write p = 2m for some m ∈ Z. Then, Equation I.1 reduces to
2m2 = q 2
Once again, 2 | q. But this contradicts the assumption that p and q are relatively
prime.
√
Remark 1.2. Notice that 2 shows up naturally when measuring √ lengths (the length
of the diagonal of a unit square). But what kind of number is 2 then?
Before we answer this, we first isolate the important properties of Q that matter to us
(i.e. to measure length, time, etc. and do basic arithmetic).
(i) Q contains Z.
4
(iv) Given x, y ∈ Q, then either x ≤ y or y ≤ x.
What we want from R is that
√
• R should contain Q, but also contain elements like 2. i.e. R should fill the ‘holes’
in Q.
2. Ordered Fields
Definition 2.1. A field is a set F with two binary operations + : F × F → F and
· : F × F → F satisfying the following axioms:
(i) (F, +) is an abelian group. The identity element is denoted by 0, and the additive
inverse of x is denoted (−x).
5
We will not prove these properties as you will prove them in MTH301 (please see [Rudin,
Proposition 1.14-1.16]).
Definition 2.4.
(ii) If (S, <) is an ordered set, we write x ≤ y to mean x < y or x = y. We write x > y
for the negation of x ≤ y.
Definition 2.5. An ordered field is a field F which is an ordered set such that
In other words, the field operations (+ and ·) respect the order structure.
Example 2.6.
(ii) Z/2Z = {[0], [1]} can be ordered naturally by declaring [0] < [1]. However, it is
not an ordered field because [0] < [1] but
(End of Day 1)
Remark 2.7. If F is an ordered field, one can prove a number of properties such as
As before, we will omit the proofs here. See [Rudin, Proposition 1.18].
R, which we will soon describe, is going to be an ordered field. We now need to explain
what we mean by saying that R will fill the ‘holes’ in Q (i.e. it will include numbers like
√
2)
6
3. The Completeness Axiom
Definition 3.1. Let (S, <) be an ordered set.
x≤z
(ii) Suppose E is bounded above. An element α ∈ S is called the least upper bound
or supremum of E if
(a) α is an upper bound for E.
(b) If β is an upper bound for E, then α ≤ β
(b’) Equivalently, if γ < α, then γ is not an upper bound for E, so there exists
x ∈ E such that γ < x.
Note that if α and α0 are two elements of S that satisfy both properties, then it
follows that
α = α0 .
Therefore, the supremum is unique, and we write α = sup(E).
y≤x
(iv) Suppose E is bounded below. Then an element α ∈ S is called the greatest lower bound
or infimum of E if
(a) α is a lower bound for E.
(b) If β is a lower bound for E, then β ≤ α.
(b’) Equivalently, if α < γ, then γ is not a lower bound for E, so there exists
x ∈ E such that x < γ.
Again, we write inf(E) for the infimum of E.
Example 3.2.
(ii) If
E = {1/n : n ∈ N} ⊂ Q
then E is bounded below (by 0). Indeed, 0 = inf(E) (we will prove this later), but
0∈/ E.
7
(iii) Let
E := {r ∈ Q : r2 < 2}.
Then, E is bounded above (by 2, say). We claim that E does not have a least
upper bound in Q.
Proof. Suppose p = sup(E) is a rational number. We know that p > 0 since 1 ∈ E.
Define
p2 − 2 2p + 2
q := p − = .
p+2 p+2
Then q is rational and
2(p2 − 2)
q2 − 2 = .
(p + 2)2
By Theorem 1.1, we have two possible cases:
(a) Suppose p2 < 2: Then, q > p and q 2 < 2. Hence, q ∈ E and so p is not an
upper bound for E.
(b) Suppose p2 > 2: Then, 0 < q < p and q 2 > 2. So if r ∈ E, then
r2 < 2 < q 2 ⇒ r < q.
Hence, q is an upper bound for E, which contradicts the assumption that
p = sup(E).
In either case, we have a contradition, so sup(E) does not exist in Q.
Definition 3.3. We say that an ordered set S has the least upper bound property if
every set that is non-empty and bounded above in S has a supremum in S.
Note that by Part (iii) of Example 3.2, Q does not have the least upper bound property.
Note: We will often omit the phrase ‘non-empty’ in statements like the next one
and it will be assumed implicitly.
Lemma 3.4. Suppose S is an ordered set that has the least upper bound property, then
whenever E ⊂ S is non-empty and bounded below, then E has an infimum in S.
Proof. Suppose E is bounded below, define A := {x ∈ S : x is a lower bound for E}.
Then, A 6= ∅. Moreover, if y ∈ E, then y is an upper bound for A. By hypothesis,
α := sup(A)
exists in S. We claim that α = inf(E) by verifying the two conditions of Part (ii) of
Definition 3.1.
(i) If γ < α, then γ is not an upper bound for A. Hence, γ ∈
/ E since every element
of E is an upper bound for A. Therefore,
α≤x
for all x ∈ E. Hence, α is a lower bound for E.
8
(ii) Suppose α < γ, then γ ∈
/ A since α is an upper bound for A. Hence, γ is not a
lower bound for E.
Therefore, α = inf(E) holds.
Theorem 3.5 (Cantor, Dedekind (1872)). There exists an ordered field R which has the
least upper bound property and contains Q as a subfield.
Remark 3.6.
(i) To say that R contains Q as a subfield means that Q ⊂ R, and
• Addition/multiplication in R, when applied to elements of Q, is the usual
operation in Q.
• Positive rational numbers are positive in R.
(ii) The ‘Completeness Axiom’ is simply the fact that R satisfies the Least Upper
Bound property, which states that R is ‘order complete’. There are other notions
of completeness you will encounter later in the course. Don’t confuse them.
(End of Day 2)
4. Properties of R
Definition 4.1. For a, b ∈ R with a < b, define
I1 ⊃ I2 ⊃ . . . .
Then,
∞
\
In 6= ∅.
n=1
a1 ≤ a2 ≤ . . . ≤ an ≤ . . . ≤ b n ≤ . . . ≤ b 2 ≤ b 1
9
for all n ∈ N. Moreover, if m ∈ N, then
an ≤ b m
α ≤ bm
(ii) Given y ∈ R with y > 0, there exists n ∈ N such that 1/n < y.
Proof.
(i) Let A = {nx : n ∈ N}, then we WTS that y is not an upper bound for A. Suppose
y is an upper bound for A, then by the Completeness Axiom,
α := sup(A)
α − x < mx.
(ii) Since y > 0, Part (i) implies that there exists n ∈ N such that ny > 1, so that
y > 1/n.
Remark 4.4. The Well-Ordering Principle states that every non-empty subset of posi-
tive integers contains a smallest member. This is an axiom, and we will use it below.
10
Proof. We wish to find m, n ∈ Z such that
m
a< < b.
n
(i) To find the denominator n: We choose n so that, if we take steps of length 1/n,
then the steps are too small to ‘step over’ the interval [x, y]. To do this, we note
that (y − x) > 0 so by Theorem 4.3, there exists n ∈ N so that
1
y−x> ⇒ ny > nx + 1
n
(ii) To find the numerator: Note that nx ∈ R so by Theorem 4.3, there exists k0 ∈ N
such that
k0 > nx.
Consider the set
S := {k ∈ N : k > nx}
Then, S is non-empty since k0 ∈ S. By the well-ordering principle, S has a smallest
member, say m. Then,
m > nx ≥ (m − 1)
(i) Uniqueness: Suppose y1 and y2 are two distinct positive roots of a, then either
y1 < y2 or y2 < y1 must hold. Assume WLOG that 0 < y1 < y2 . Then, it follows
(from Definition 2.5 - See Homework 1) that
11
(ii) Existence: Consider
E := {x ∈ R : xn ≤ a}
Then, E 6= ∅ because 0 ∈ E. Moreover,
n n n−1 n n−2
(a + 1) = a + na + a + . . . + na + 1 ≥ na > a
2
So E is bounded above by (a + 1). Hence,
α := sup(E)
exists in R.
(a) Suppose αn < a, then we WTS that there exists M ∈ N such that
n
1
α+ <a
M
By the Binomial theorem,
n n
1 n
X n n−k 1
α+ =α + α
M k=1
k Mk
" n #
1 X n
< αn + αn−k
M k=1 k
Pn n
Therefore, if B := k=1 k
αn−k , we wish to find M ∈ N so that
B B
αn +<a⇔M >
M a − αn
However, a − αn > 0 by hypothesis, so by the Archimedean property, such an
M ∈ N exists. So we conclude that
1
α+ ∈E
M
which contradicts the assumption that α is an upper bound for E.
(b) Suppose αn > a, then we WTS that there exists M ∈ N such that
n
1
α− > a.
M
By the Binomial theorem,
n n
1 n
X
k n 1
α− =α + (−1) αn−k k
M k=1
k M
n
n
X n n−k 1
>α − α
k=1
k Mk
" n #
1 X n
> αn − αn−k
M k=1 k
12
Pn n
n−k
Again, set B := k=1 k
α and choose M ∈ N so that
B
M>
αn − a
works. Then, for any x ∈ E,
n
n 1 1
x ≤a< α− ⇒x<α− .
M M
13
II. Sequences of Real Numbers
1. Limits of Sequences
Definition 1.1. The absolute value of a real number x ∈ R is defined as
(
x : if x > 0
|x| =
−x : if x ≤ 0
Remark 1.2.
d(x, y) := |x − y|
(iii) Note that for any x, y ∈ R, x = y if and only if |x − y| < for all > 0.
Proof. If x = y and > 0, then |x − y| = 0 < .
Conversely, suppose that |x − y| < for all > 0. WTS: x = y. Suppose not, then
r := |x − y| > 0
If = r/2, then clearly |x − y| < does not hold. This violates our assumption,
so x = y must hold.
Definition 1.3.
(i) A sequence of real numbers is an ordered list of the form (x1 , x2 , x3 , . . .), where
each xi ∈ R.
(ii) We say that a sequence (xn ) converges to a point x ∈ R if for each > 0, there
exists N ∈ N such that |xn − x| < for all n ≥ N . If this happens, we write
lim xn = x.
n→∞
14
(iii) If no such x exists, then we say that (xn ) diverges (or we say that limn→∞ xn does
not exist).
Example 1.4.
1
(i) limn→∞ n
=0
Proof. For > 0, there exists N ∈ N such that N1 < (by Theorem I.4.3). So for
n ≥ N , we have
1 1
|xn − 0| = ≤ < .
n N
This is true for any > 0 so limn→∞ xn = 0.
So for = 1/2, there is no N ∈ N such that |n2 − x| < for all n ≥ N . This
contradicts the assumption that limn→∞ n2 exists.
(End of Day 4)
Also have a look at [Abbot, Examples 2.2.5, 2.2.6 and 2.2.7] for a good explanation of
convergence and divergence.
Lemma 1.5. If a sequence is convergent, then its limit is unique. i.e. If limn→∞ xn = x1
and limn→∞ xn = x2 , then x1 = x2 .
Proof. Suppose x1 6= x2 , then r := |x1 − x2 | > 0. So if = r/2, then there exists N1 ∈ N
such that
|xn − x1 | <
for all n ≥ N1 . Similarly, there exists N2 ∈ N such that
|xn − x2 | <
15
Note: A set S ⊂ R is bounded if and only if it is bounded above and bounded be-
low. (Why?)
|xn − x| < 1
Proof.
(i) Homework.
16
(iii) Since (xn ) is convergent, it is bounded by Lemma 1.7. So there exists M > 0 so
that
|xn | ≤ M
for all n ∈ N. Fix > 0 and consider two cases:
(a) Suppose first that y = 0, then there exists N ∈ N so that |yn | < /M for all
n ≥ N . Then,
|xn yn | ≤ M |yn | <
for all n ≥ N . Thus, limn→∞ (xn yn ) = 0.
(b) Suppose that y 6= 0, then there exists N1 , N2 ∈ N so that
|xn − x| < /2|y| for all n ≥ N1
|yn − y| < /2M for all n ≥ N2
If N := max{N1 , N2 }, then both inequalities hold. Therefore, for all n ≥ N ,
|xn yn − xy| ≤ |xn yn − xn y| + |xn y − xy|
≤ M |yn − y| + |y||xn − x|
< /2 + /2 = .
Thus, limn→∞ (xn yn ) = xy.
(iv) Suppose y 6= 0, so that r := |y| > 0. Then, there exists N ∈ N so that
r
|yn − y| <
2
for all n ≥ N . Therefore,
r r
|yn | > |y| − = > 0.
2 2
1
for all n ≥ N . Moreover, the ratio yn makes sense for all n ≥ N . We assume
WLOG that yn 6= 0 for all n ∈ N.
By part (iii), it now suffices to show that
1 1
lim = .
n→∞ yn y
Note that
1 1 |yn − y| |yn − y|
− = ≤
yn y |yn y| r2 /2
So if > 0, there exists N0 ∈ N so that for all n ≥ N ,
1 1
|yn − y| < (r2 /2) ⇒ − < .
yn y
1
Hence, limn→∞ yn
= y1 .
(End of Day 5)
17
2. Limits and Order
Proposition 2.1. Suppose limn→∞ xn = x and limn→∞ yn = y.
Proof.
(i) Suppose x < 0, then there exists > 0 such that x + < 0. Now, there exists
N ∈ N so that |xn − x| < for all n ≥ N . Then, xN ∈ (x − , x + ) so
xN < 0.
lim zn = y − x
n→∞
y − x ≥ 0 ⇒ y ≥ x.
(iii) Again, take yn = c for all n ∈ N. Then, limn→∞ yn = c, so we may apply part (ii).
Remark 2.2.
(i) Suppose limn→∞ xn = x and suppose that there exists N ∈ N so that xn ≥ 0 for
all n ≥ N (some initial terms may be negative). Then, the proof of part (i) of
Proposition 2.1 shows that x ≥ 0.
(ii) Indeed, if a property (P) holds for all members of a sequence (xn ) except possibly
for the first finitely many terms, then we say that (xn ) eventually has property (P).
Therefore, part (i) Proposition 2.1 may be restated to say that if (xn ) is eventually
non-negative, then its limit is non-negative.
(iii) The same is true for the assumptions in part (ii), (iii) and (iv). Similar results will
hold in future theorems as well.
18
Theorem 2.3 (Squeeze Theorem). If limn→∞ xn = x, limn→∞ yn = x and xn ≤ zn ≤ yn
for all n ∈ N, then (zn ) is convergent and limn→∞ zn = x as well.
Proof. Homework.
Definition 2.4.
(i) A sequence (xn ) is said to be increasing if xn ≤ xn+1 for all n ∈ N.
α − < xN ≤ α.
It then follows that |yj − x| < for all j ≥ j0 , so that limn→∞ yj = x as well.
19
Remark 3.3.
(i) We know that limn→∞ n1 = 0. The sequence (1/2n ) is clearly a subsequence of the
original sequence. Therefore,
1
lim n = 0.
n→∞ 2
(ii) Consider the sequence (1, −1, 1, −1, . . .). It has a subsequence (1, 1, 1, . . .) and
another subsequence (−1, −1, −1, . . .). The first subsequence converges to 1, while
the second converges to −1. This does not violate Lemma 3.2 because the original
sequence is not convergent!
(iii) More generally, suppose (xn ) is a sequence such that the set {xn : n ∈ N} is
finite (i.e. There are only finitely many numbers that keep repeating). Then, one
number must repeat infinitely often, and so (xn ) has a convergent subsequence.
Note: For an interval I = [a, b], we write `(I) = (b − a) for its length.
−M ≤ xn ≤ M
for all n ∈ N.
(i) Divide the interval [−M, M ] into two halves,
If both J1 and J2 contain only finitely many members of S, then S would have
to be finite. Since this is not the case, either J1 or J2 contains infinitely many
elements of S. Let I1 denote the interval which does.
(ii) Now divide the interval I1 into two equal halves, which we denote by K1 and K2 .
Once again, either K1 ∩ S or K2 ∩ S is infinite. Let I2 denote the interval which
contains infinitely members of S.
(iv) We now proceed inductively. Suppose we have chosen closed intervals {I1 , I2 , . . . , Ik }
such that
(a) Each Ij , 1 ≤ j ≤ k is closed.
(b) Ij ⊂ Ij−1 for all 2 ≤ j ≤ k.
(c) For each 2 ≤ j ≤ k, `(Ij ) = `(Ij−1 )/2.
20
(d) Each Ij contains infinitely many members of S.
To construct Ik+1 , we do exactly what we did in Step (ii).
(v) Thus proceeding, we construct a sequence (Ij ) of nested closed intervals, each of
which contains infinitely members of S. By Proposition I.4.2,
∞
\
In 6= ∅.
n=1
(vii) We claim that limj→∞ xnj = x. To see this, fix > 0 and observe that
`(I1 ) = 2M
`(I2 ) = M
`(I3 ) = M/2
.. ..
.=.
`(Ij ) = 2M/2j−1 = M/2j−2 .
M
<
2j−2
for all j ≥ j0 . Now, if j ≥ j0 , then
(End of Day 6)
21
4. Cauchy Sequences
Definition 4.1. A sequence (xn ) ⊂ R is said to be Cauchy if for each > 0, there exists
N ∈ N such that |xn − xm | < whenever n, m ≥ N .
Proposition 4.2. Every convergent sequence is Cauchy.
Proof. If (xn ) converges to x, then for any > 0, there exists N ∈ N such that |xn − x| <
/2 whenever n, m ≥ N . So by the triangle inequality
|xn − xm | ≤ |xn − x| + |x − xm | <
for all n, m ≥ N .
We wish to prove the converse, which gives us a criterion to determine when a sequence
converges. For that, we need two lemmas, the first of which is analogous to Lemma 1.7.
Lemma 4.3. Every Cauchy sequence is bounded.
Proof. If (xn ) is Cauchy, then for = 1, there exists N ∈ N such that |xn − xm | < 1 for
all n, m ≥ N . In particular
|xn − xN | < 1
whenever n ≥ N , so that
|xn | < 1 + |xN |
for all n ≥ N . So if
M := max{|x1 |, |x2 |, . . . , |xN |, |xN | + 1},
then |xj | ≤ M for all j ∈ N.
Lemma 4.4. If a Cauchy sequence has a convergent subsequence, then the whole se-
quence converges.
Proof. Suppose (xn ) is Cauchy and (xnk ) is a convergent subsequence that converges to
x ∈ R. Then, for each > 0, there exist N1 , N2 ∈ N such that
|xn − xm | < /2 for all n, m ≥ N1 , and
|xnk − x| < /2 for all k ≥ N2 .
Now, (nk ) is a subsequence of (xn ) so n1 < n2 < . . .. Hence, there exists K0 ∈ N such
that
nK0 ≥ N1
Moreover, we may choose K0 so that K0 ≥ N2 as well (Why?). Therefore,
|xn − xnK0 | < /2 for all n ≥ N1 , and
|xnK0 − x| < /2
So if n ≥ N1 , we have
|xn − x| ≤ |xn − xnK0 | + |xnK0 − x| < .
Hence (xn ) converges to x.
22
Theorem 4.5 (Cauchy Criterion). A sequence in R is convergent if and only if it is
Cauchy.
Remark 4.6. The Cauchy Criterion of Theorem 4.5 is also referred to as ‘completeness’
of R. Note that this is different from the ‘order completeness’ of Section 3. Again, don’t
confuse them (see Remark I.3.6).
23
III. Topology of R
1. Open Sets
Lemma 1.1. Consider an open interval U := (a, b) ⊂ R. If x ∈ U , then there exists
δ > 0 such that (x − δ, x + δ) ⊂ U .
Definition 1.2.
B(x, δ) = (x − δ, x + δ) = {y ∈ R : |x − y| < δ}
(ii) A set U ⊂ R is said to be open if for each x ∈ U , there exists δ > 0 such that
B(x, δ) ⊂ U . (Note: The value of δ depends on x, just like in Lemma 1.1).
Example 1.3.
(iii) If I = [0, 1], then I is not an open set because if x = 1, then for no δ > 0 is it true
that B(x, δ) ⊂ I.
Note: You should think of an open set as one where every point has some room around
it.
Theorem 1.4.
24
Proof.
S
(i) Suppose {Uj : j ∈ J} is an arbitrary collection of open sets and U = j∈J Uj .
WTS: U is open, so pick x ∈ U . Then, there exists j ∈ J such that x ∈ Uj . Uj is
open so there exists δ > 0 such that B(x, δ) ⊂ Uj . Clearly, B(x, δ) ⊂ U . This is
true for every x ∈ U so U is open.
(End of Day 7)
Suppose {U1 , U2 , . . . , Uk } is a collection of finitely many open sets and U :=
(ii) T
k
i=1 Ui . WTS: U is open, so pick x ∈ U . Then, for any fixed 1 ≤ i ≤ k,
x ∈ Ui , so there exists δi > 0 such that B(x, δi ) ⊂ Ui . Let
δ := min{δi : 1 ≤ i ≤ k} > 0
Then, we claim that B(x, δ) ⊂ U . To see this, fix 1 ≤ j ≤ k. Then,
B(x, δ) ⊂ B(x, δj ) ⊂ Uj .
This is true for each 1 ≤ j ≤ k, so B(x, δ) ⊂ U . This is true for each x ∈ U , so U
is open.
Remark 1.5.
(i) Note that in Part (iii) of Theorem 1.4, we necessarily needed finitely many open
sets so that the value of δ chosen in the proof is positive. Therefore, this proof
would not work if there were infinitely many sets.
(ii) Moreover, the statement would be false for infinitely many sets: Set Un := (−1/n, 1/n),
then each Un is open but
\∞
Un = {0}
n=1
which is not open.
2. Closed Sets
Definition 2.1. Let A ⊂ R be a set. A point x ∈ R is called a limit point of A if for
each δ > 0, there is a point y ∈ A such that
y 6= x and |y − x| < δ.
In other words, (B(x, δ) ∩ A) \ {x} =
6 ∅ for each δ > 0.
Lemma 2.2. x ∈ R is a limit point of A if and only if there is a sequence (xn ) ⊂ A
such that xn 6= x for all n ∈ N and
lim xn = x.
n→∞
25
Proof. If x is a limit point of A, then for each n ∈ N, with δ = 1/n, there is xn ∈ A
such that xn 6= x and |xn − x| < 1/n. Now if > 0, there is N ∈ N so that 1/N < , so
if n ≥ N , we have
|xn − x| < 1/n ≤ 1/N <
for all n ≥ N . Therefore, limn→∞ xn = x.
Conversely, suppose there is a sequence (xn ) as above, then for any δ > 0, there is N ∈ N
so that |xN − x| < δ. Therefore,
xN ∈ B(x, δ) ∩ A
Remark 2.3.
(i) A limit point is also sometimes called a cluster point of the set.
(ii) The set of all limit points of A is called the derived set of A, and is denoted by A0 .
Definition 2.4. A set F ⊂ R is said to be closed if it contains all its limit points.
Example 2.5.
a ≤ xn ≤ b
26
(iii) If A = {1/n : n ∈ N}, then A is not closed.
Proof. By Example II.1.4, limn→∞ 1/n = 0. So 0 ∈ A0 . Since 0 ∈
/ A, A is not
closed.
Proof.
B(x, δ) ∩ F 6= ∅.
Since x ∈
/ F , it follows that x is a limit point of F . Since F is closed, x ∈ F . But
this contradicts the assumption that x ∈ U . Hence, U must be open.
(ii) Suppose U is open, we WTS that F is closed. The proof reverses the above
argument: Suppose x ∈ F 0 , then we WTS that x ∈ F . Suppose not, then x ∈ U .
Since U is open, there is δ > 0 so that B(x, δ) ⊂ U . In that case,
B(x, δ) ∩ F = ∅.
Corollary 2.7.
A = A ∪ A0 .
Lemma 3.2. Let A ⊂ R and x ∈ R be fixed. Suppose δ > 0 is such that B(x, δ) ∩ A 6= ∅,
then
B(x, δ) ∩ A 6= ∅.
27
Proof. Suppose B(x, δ) ∩ A 6= ∅. Since A = A ∪ A0 , there are two options: either
B(x, δ) ∩ A 6= ∅ or B(x, δ) ∩ A0 6= ∅. In the first case, there is nothing to prove, so
assume that
B(x, δ) ∩ A0 6= ∅.
In that case, there is a y ∈ A0 such that |y − x| < δ. Now r := δ − |y − x| > 0, so by
definition
B(y, r) ∩ A 6= ∅
Choose z ∈ B(y, r) ∩ A, then
|z − x| ≤ |z − y| + |y − x| < δ − |y − x| + |y − x| = δ.
(End of Day 8)
B(x, δ) ∩ A 6= ∅
Proof.
(i) Suppose x ∈ A and δ > 0. Then there are two cases: either x ∈ A or x ∈ A0 . In
the first case, it is clear that
x ∈ B(x, δ) ∩ A.
(ii) Conversely, suppose that the given condition holds. We WTS that x ∈ A. Suppose
not, then x ∈
/ A and x ∈/ A0 . Since x ∈
/ A0 , there must be a δ > 0 such that
(B(x, δ) ∩ A) \ {x} = ∅.
Since x ∈
/ A, it follows that
B(x, δ) ∩ A = ∅.
This contradicts the hypothesis.
28
(i) A is a closed set.
(ii) If F is a closet set containing A, then A ⊂ F .
In other words, A is the smallest closed set containing A.
Proof.
c
(i) To show that A is closed, it suffices to show that U := A is open. So fix x ∈ U .
Then, x ∈
/ A so by Proposition 3.3, there exists δ > 0 such that
B(x, δ) ∩ A = ∅.
We claim that B(x, δ) ⊂ U . To see this, choose y ∈ B(x, δ), then |y − x| < δ. Set
r := δ − |y − x|, then we claim that
B(y, r) ⊂ B(x, δ).
Indeed, if z ∈ B(y, r), then
|z − x| ≤ |z − y| + |y − x| < δ − |y − x| + |y − x| = δ ⇒ z ∈ B(x, δ).
Therefore,
B(y, r) ∩ A = ∅.
By Proposition 3.3, y ∈
/ A, so that y ∈ U . This is true for each point y ∈ B(x, δ),
so B(x, δ) ⊂ U . Such a δ > 0 exists for each x ∈ U , so U is an open set.
(ii) Suppose F is a closed set that contains A, then any limit point of A must also
either be in F , or be a limit point of F . In either case, we see that A0 ⊂ F , so that
A = A ∪ A0 ⊂ F.
Example 3.5.
(i) If F is a closed set, then F = F . Hence,
(a) {1} = {1}.
(b) [0, 1] = [0, 1].
(ii) If A = (0, 1), then A = [0, 1].
Proof. Since [0, 1] is closed, we know from Theorem 3.4 that A ⊂ [0, 1]. Therefore,
it suffices to show that [0, 1] ⊂ A. Since A ⊂ A, it suffices to show that
{0, 1} ⊂ A.
(a) If x = 0: Then (1/n) ⊂ A that converges to x, so 0 ∈ A0 .
(b) If x = 1: Then, (1 − 1/n) ⊂ A that converges to x, so 1 ∈ A0 .
(iii) Q = R.
Proof. Homework.
29
4. Compact Sets
Definition 4.1. A set K ⊂ R is said to be compact if every sequence in K has a
subsequence that converges to a point in K.
Lemma 4.2. A compact set must be both closed and bounded.
Proof.
(i) If K is compact and not closed, then K 6= K, so K 0 is not contained in K. So
there must be a sequence (xn ) in K that converges to a point x ∈
/ K. Then, every
subsequence if (xn ) converges to x (by Lemma II.3.2), so no subsequence of (xn )
can converge to a point in K.
(ii) Suppose K is compact and not bounded. Then, for each n ∈ N, there exists
xn ∈ K such that |xn | > n. The sequence (xn ) is now unbounded. Indeed,
every subsequence of (xn ) is also unbounded, so no subsequence can converge (by
Lemma II.1.7).
Example 4.3.
(i) Every closed and bounded interval is compact.
Proof. If K = [a, b], then any sequence in K is bounded, so it has a convergent
subsequence by the Bolzano-Weierstrass theorem. Since K is closed, that limit
point also belongs to K.
(iv) More generally, the union of finitely many compact sets is compact.
(v) The union of infinitely many compact sets may not be compact: Take Kn =
[0, 1 − 1/n], then each Kn is compact, but
∞
[
K= Kn = [0, 1)
n=1
30
Theorem 4.4 (Heine-Borel Theorem). A subset K ⊂ R is compact if and only if it is
closed and bounded.
Proposition 4.5. If K is a compact set, then sup(K) and inf(K) both exist, and are
elements of K.
Proof. Since K is bounded, both α := sup(K) and β := inf(K) exist. We prove that
α ∈ K (the proof for β is analogous). For each > 0, α − is not an upper bound for
K, so there is a point x ∈ K such that
α−<x≤α
Since K is compact, (xn ) has a convergent subsequence (xnj ) so that x := limj→∞ xnj ∈
K. Now,
α − 1/nj < xnj ≤ α
holds for all j ∈ N, so by the Squeeze theorem, x = α, so α ∈ K.
31
IV. Continuous Functions
1. Limits of Functions
Definition 1.1. Let A ⊂ R and f : A → R be a function, and c ∈ A0 be a limit point
of A. For a real number L ∈ R, we write
“ lim f (x) = L”
x→c
Example 1.2.
This is because, if (xn ) is a sequence with limn→∞ xn = 1/2, then limn→∞ x2n = 1/4
by the Algebra of Limits (Theorem II.1.8).
lim = 1/2n .
x→1/2
(iii) Suppose f : [0, 1] → R is the constant function f (x) = 3 for all x ∈ [0, 1], then for
any c ∈ [0, 1],
lim f (x) = 3 = f (c).
x→c
32
(v) Suppose f : [0, 2] → R is given by
(
x : if 0 ≤ x ≤ 1
f (x) =
x2 : otherwise.
33
Remark 1.3. In Definition 1.1, we required that c ∈ A0 . This is to ensure that there are
(non-constant) sequences in A converging to c. We do not necessarily need c to belong
to A (the domain of f ).
Proof. Follows from the definition and the Algebra of Limits (Theorem II.1.8).
Theorem 1.5. Given f : A → R and c ∈ A0 , limx→c f (x) = L if and only if for each
> 0, there is a δ > 0 such that
Proof.
(i) Suppose limx→c f (x) = L, and > 0 is given. We WTS that there is a δ > 0
satisfying this condition. Suppose not, then δ = 1/n does not satsify this condition.
So there is a point xn ∈ A with 0 < |xn − c| < 1/n but |f (xn ) − L| ≥ . Now, the
sequence (xn ) converges to c, xn 6= c for all n ∈ N, but (f (xn )) does not converge
to L. This contradicts the definition.
(ii) Conversely, suppose the Equation IV.1 holds. We WTS: limx→c f (x) = L. So
choose a sequence (xn ) ⊂ A such that limn→∞ xn = c and xn 6= c for all n ∈ N. We
WTS: limn→∞ f (xn ) = L. So fix > 0. By hypothesis, there is a δ > 0 satisfying
Equation IV.1. For this δ > 0, there is N ∈ N such that
34
2. Continuous Functions
Definition 2.1. Given a function f : A → R and a point c ∈ A, we say that f is
continuous at c if
lim f (x) = f (c).
x→c
(i) f is continuous at c.
(iii) For each > 0, there exists δ > 0 such that if x ∈ A and |x − c| < δ, then
|f (x) − f (c)| < .
Proof.
(i) f + g is continuous at c.
(iii) f g is continuous at c.
35
Proof. This is a combination of Definition 1.1, Theorem II.1.8 and Definition 2.1. For
instance, we prove part (i). Suppose (xn ) is a sequence in A with limn→∞ xn = c. Then,
by Definition 2.1,
By Theorem II.1.8,
Hence, (f + g) is continuous at c.
Remark 2.5. The sequential definition (Definition 1.1) is useful to determine when a
function is not continuous at a point. There are two possible reasons:
(i) limx→c f (x) does not exist: If there are two sequences (xn ) and (yn ) in A with
limn→∞ xn = c = limn→∞ yn , but
Example 2.6.
p(x) = a0 + a1 x + . . . + an xn
x2 + 3x + 7
f (x) =
(x − 1)(x − 2)(x − 3)2
36
(v) Let f : [0, 2] → R be given by
(
1 : if 0 ≤ x < 1
f (x) =
2 : otherwise
Here, the limit does not exist because the function oscillates too much.
Then, f is continuous at c = 0.
Proof. Note that | sin(t)| ≤ 1 for all t ∈ R, so
|f (x)| ≤ |x|
37
(ix) Let A := [0, ∞) and fix m ∈ N. Define f : A → R by
f (x) = x1/m .
δ := min{c − a, b − c}
38
Proposition 2.8 (Composition of Continuous Functions). Let f : A → R and g : B →
R as above and let c ∈ A. If f is continuous at c and g is continuous at f (c), then g ◦ f
is continuous at c.
Hence,
lim (g ◦ f )(xn ) = (g ◦ f )(c).
n→∞
So g ◦ f is continuous at c.
(ii) We say that f is bounded if f (A) is a bounded set. In that case, we may define
Remark 3.2.
39
Proof. Let (yn ) be a sequence in f (K), then we may write yn = f (xn ) for some sequence
(xn ) ⊂ K. Since K is compact, there is a subsequence (xnj ) of (xn ) and a point x ∈ K
such that limj→∞ xnj = x. Since f is continuous, limj→∞ f (xnj ) = f (x) ∈ f (K). So
(f (xnj )) is a subsequence of (yn ) that converges to a point in f (K). Hence, f (K) is
compact.
Corollary 3.4 (Extreme Value Theorem). If K is compact and f : K → R is continu-
ous, then f is bounded. Moreover, there exist points x0 , x1 ∈ K such that
Proof. Note that f (K) is compact by Theorem 3.3, so f (K) is both closed and bounded
by Lemma III.4.2. Hence, f (K) is bounded.
Since y0 ∈ f (K), there exist x0 ∈ K such that f (x0 ) = y0 and similarly, there is x1 ∈ K
such that f (x1 ) = y1 . This proves the result.
Remark 3.5. You will have used Corollary 3.4 in Calculus before. Given a function
f : [0, 1] → R, you are often asked to find the maxima/minima of f . This is done in two
steps:
(i) First determine if f (0) or f (1) is a maximum.
(ii) For points in (0, 1), you can use the derivative to test.
This entire process relies on the fact that f has a maximum/minimum on the set,
otherwise there is no guarantee that this process would work.
I1 ⊃ I2 ⊃ I3 ⊃ . . .
and further assume that limn→∞ |bn − an | = 0. Then, there is a unique point x0 ∈ R
such that ∞
\
In = {x0 }.
n=1
40
Suppose A has two distinct points, say y0 < y1 so that r := |y1 − y0 | > 0. Then, for each
n ∈ N,
an ≤ y0 < y1 ≤ bn .
In other words, for each n ∈ N,
|bn − an | ≥ |y1 − y0 | > r.
This contradicts the assumption that limn→∞ |bn − an | = 0. Therefore, the intersection
must contain exactly one point.
Theorem 4.2 (Intermediate Value Theorem). Suppose [a, b] ⊂ R is a closed interval
and f : [a, b] → R is continuous. Suppose that f (a) < f (b), and c ∈ R is any value such
that
f (a) ≤ c ≤ f (b).
Then, there is a point x ∈ [a, b] such that f (x) = c.
Proof. Assume without loss of generality that f (a) < c < f (b).
(i) Set a1 = a, b1 = b and consider x1 := (a1 + b1 )/2. If f (x1 ) = c, then we are done.
If not, then either
f (x1 ) > c or f (x1 ) < c.
Note that
lim an = x = lim bn
n→∞ n→∞
Since f is continuous, we have
f (x) = lim f (an ) = lim f (bn ).
n→∞ n→∞
Since f (an ) ≤ c ≤ f (bn ) and for all n ∈ N, it follows from Theorem II.2.3 that
f (x) = c.
41
Note that in the hypothesis of Theorem 4.2, we assumed that f (a) < f (b). But the
same is true if we assume f (b) > f (a) as well.
Remark 4.3.
(i) Consider the function f (x) = 1/x defined in R \ {0}. Given > 0, there is N ∈ N
so that if |x| > N , one has
|f (x)| <
(Indeed, we may use the Archimedean property (see Theorem I.4.3) to find N ∈ N
such that 1/N < ).
(ii) Similarly, if f (x) = 1/x2 and > 0, there is N ∈ N such that if |x| > N , one has
|f (x)| < .
(iii) In general, if k ≥ 1 and a ∈ R are fixed, and > 0 is given, there exists N ∈ N
such that if |x| > N ,
a
< .
xk
We use this fact below.
(End of Day 13)
Corollary 4.4. Suppose f : R → R is a polynomial of odd degree, then f has at least
one real root. i.e. There exists x ∈ R such that f (x) = 0.
Proof. Write
f (x) = a0 + a1 x + . . . + am xm
where am 6= 0 and m is odd. Then,
m−1
f (x) X
= 1 + ak xk−m
xm k=0
1
Fix := 2(m−1) > 0 (you will see why later). By Remark 4.3, for each 0 ≤ k ≤ m − 1,
there is Nk ∈ N such that if |x| > Nk . then
|ak xk−m | <
If N = max{N0 , N1 , . . . , Nm−1 }, then for any x ∈ R with |x| > N , one has
f (x)
− 1 < (m − 1).
xm
Hence, if |x| > N , one has
f (x) 1
m
> 1 − (m − 1) > .
x 2
1
because = 2(m−1)
. Now consider two cases:
42
(i) If x > N , then
xm
f (x) > > 0.
2
(ii) If x < −N , then
xm
f (x) << 0.
2
Hence, in particular, f (−N − 1) < 0 and f (N + 1) > 0. So by applying the
Intermediate Value Theorem Theorem 4.2 to f on the interval [−N − 1, N + 1],
we see that f must have a root.
Remark 4.5. Note that Corollary 4.4 does not hold for polynomials of even degree
(for instance, x2 + 1 does not have a real root). Try to find where exactly we used the
assumption of odd degree in the proof.
43
So δ = 1/39 works.
Therefore, the value of δ > 0 depends not just on but also on the point c.
Remark 5.3.
(i) Note that the definition of continuity (see Definition 2.1) is local. i.e. It describes
the behaviour of a function around a point. However, the definition of uniform
continuity is global ; it describes the behaviour of the function on the set.
Example 5.4.
1 1 1
− =
n n+1 n(n + 1)
but
1 1
f −f =1
n n+1
44
So if δ > 0 is chosen, there exists n ∈ N so that 1/n(n + 1) < δ, and in that case,
x = 1/n and y = 1/(n + 1) satisfy |x − y| < δ but
1
Moreover, |zk − wk | < njk
, so x = y must hold.
|f (zk ) − f (wk )| ≥ .
45
Example 5.8. Let [a, b] ⊂ R be a closed and bounded (compact) interval. Let f :
[0, 1] → R be a continuous, non-negative function. For a given > 0, there is a δ > 0
such that
|x − y| < δ ⇒ |f (x) − f (y)| <
Choose n ∈ N so that 1/n < δ. Define a partition of [0, 1] by
Then, we have
0≤A−B <
We will see later that Z b
B≤ f (t)dt ≤ A
a
as well. For now, try to visualize these numbers A and B in terms of the area under the
curve f .
46
V. Differentiation
1. Definition and Basic Properties
Definition 1.1. Let A = (a, b) ⊂ R be an open interval and f : A → R be a function,
and c ∈ A be fixed.
f (x) − f (c)
lim
x→c x−c
exists. In that case, we denote this number by f 0 (c).
f 0 (t) = ntn−1
f (x) − f (c)
lim = lim(xn−1 + cxn−2 + . . . + cn−1 ) = ncn−1
x→c x−c x→c
because polynomials are continuous functions (see Part (iii) of Example IV.2.6).
|x|
lim
x→0 x
does not exist.
47
(iii) Let f : R → R is (
x sin(1/x) : if x 6= 0
f (x) =
0 : if x = 0.
Then, f is not differentiable at c = 0.
Proof. Taking xn := 1/(2nπ) and yn := 1/(2nπ + π/2), we see that
f (x)
lim = lim sin(1/x)
x→0 x x→0
(iv) Let f : R → R be (
x2 sin(1/x) : if x 6= 0
f (x) =
0 : if x = 0.
Then, f is differentiable at c = 0.
Proof. Again,
f (x) − f (0)
lim = lim x sin(1/x) = 0
x→0 x−0 x→0
f (xn ) − f (c)
lim (f (xn ) − f (c)) = lim (xn − c) = 0 × f 0 (c) = 0
n→∞ n→∞ xn − c
by Theorem II.1.8. Hence, limx→c f (x) = f (c).
Lemma 1.4. Suppose A ⊂ R is any set and g : A → R is a continuous.
(i) If c ∈ A is such that g(c) > 0, then there exists δ > 0 such that g(y) > 0 for all
y ∈ B(x, δ) ∩ A.
(ii) If c ∈ A is such that g(c) < 0, then there exists δ > 0 such that g(y) < 0 for all
y ∈ B(x, δ) ∩ A.
Proof. Exercise.
Theorem 1.5 (Algebra of Differentiable Functions). Suppose f, g : (a, b) → R are two
functions that are differentiable at c ∈ (a, b), and let k ∈ R. Then
48
(i) (f + g) is differentiable at c and (f + g)0 (c) = f 0 (c) + g 0 (c).
(ii) (kf ) is differentiable at c and (kf )0 (c) = kf 0 (c).
(iii) [Product Rule] (f g) is differentiable at c and (f g)0 (c) = f 0 (c)g(c) + f (c)g 0 (c).
(iv) [Quotient Rule] If g(c) 6= 0, then (f /g) is differentiable at c and
0
f f 0 (c)g(c) − f (c)g 0 (c)
(c) =
g g(c)2
Proof.
(i) Exercise
(ii) Exercise
(iii) Suppose (xn ) ⊂ (a, b) is such that limn→∞ xn = c. Then, write
(f g)(xn ) − (f g)(c) f (xn )g(xn ) − f (c)g(xn ) + f (c)g(xn ) − f (c)g(c)
=
xn − c xn − c
f (xn ) − f (c) g(xn ) − g(c)
= g(xn ) + f (c)
xn − c xn − c
Since g is continuous at c by Proposition 1.3,
lim g(xn ) = g(c)
n→∞
1 g(c) − g(xn ) 1
= lim ·
g(c) n→∞ xn − c g(xn )
1 1
= · (−g 0 (c)) ·
g(c) g(c)
where the last limit follows from the algebra of limits Theorem II.1.8.
49
(End of Day 15)
(i) f is said to have a local maximum at c ∈ A if there exists δ > 0 such that
Proposition 2.3. Suppose f : (a, b) → R has a local extremum at a point c ∈ (a, b). If
f is differentiable at c, then f 0 (c) = 0.
50
Proof. Assume without loss of generality that c is a local maximum. Then, there is
δ > 0 such that
f (y) ≤ f (c) for all y ∈ B(c, δ)
(We may assume that B(c, δ) ⊂ (a, b) since (a, b) is open). Now choose a sequence
xn := c + 1/n so that limn→∞ xn = c. Then, by ignoring the first few terms, we may
assume that xn ∈ B(c, δ) for all n ∈ N. Hence,
f (xn ) − f (c)
f 0 (c) = lim ≤0
n→∞ xn − c
since xn ≥ c for all n ∈ N. Taking yn = c − 1/n, we see that
f (yn ) − f (c)
f 0 (c) = lim ≥0
n→∞ yn − c
Hence, f 0 (c) = 0.
Theorem 2.4 (Rolle’s Theorem). Let f : [a, b] → R be continuous and differentiable on
(a, b). If f (a) = f (b), then there exists c ∈ (a, b) such that f 0 (c) = 0.
Proof. Assume that f 0 (c) 6= 0 for all c ∈ (a, b). By Proposition 2.3, f has no local
extrema in (a, b). By Corollary IV.3.4, f has absolute extrema
M := sup f ([a, b]) and m := inf f ([a, b])
in [a, b]. So these extrema must be attained at a or b. Since f (a) = f (b), it follows that
m = M.
But in that case, f is a constant function, so f 0 (c) = 0 for any c ∈ (a, b).
Theorem 2.5 (Mean Value Theorem). Suppose f : [a, b] → R is continuous and differ-
entiable on (a, b). Then, there is a point c ∈ (a, b) such that
f (b) − f (a)
= f 0 (c).
b−a
(Equivalently, f (b) − f (a) = f 0 (c)(b − a)).
Proof. Let h : [a, b] → R be given by
h(x) := f (x)(b − a) − x(f (b) − f (a)).
Then, h is continuous on [a, b] and differentiable on (a, b). Moreover,
h(a) = f (a)(b − a) − a(f (b) − f (a)) = f (a)b − af (b) and
h(b) = f (b)(b − a) − b(f (b) − f (a)) = −f (b)a + bf (a)
So h(a) = h(b). By Rolle’s Theorem (Theorem 2.4), there exists c ∈ (a, b) such that
h0 (c) = 0. Hence,
f 0 (c)(b − a) − (f (b) − f (a)) = 0
So c is the desired point.
51
3. Extrema and Curve Sketching
Theorem 3.1. Let f : [a, b] → R be continuous and differentiable on (a, b).
(i) If f 0 (x) > 0 for all x ∈ (a, b), then f is strictly increasing on [a, b].
(ii) If f 0 (x) < 0 for all x ∈ (a, b), then f is strictly decreasing on [a, b].
(iii) If f 0 (x) = 0 for all x ∈ (a, b), then f is constant on [a, b].
Proof.
Since f 0 (c) > 0, it follows that f (y) > f (x). Hence, f is strictly increasing.
Since f 0 (c) = 0, f (x) = f (a). This is true for all x ∈ (a, b], so f is constant.
Proposition 3.2 (First Derivative Test for Extrema). Let f : [a, b] → R be continuous
and differentiable on (a, b). Fix c ∈ (a, b) and δ > 0 so that (c − δ, c + δ) ⊂ (a, b).
(i) If f 0 (x) > 0 for all x ∈ (c − δ, c) and f 0 (x) < 0 for all x ∈ (c, c + δ), then f has a
local maximum at c.
(ii) If f 0 (x) < 0 for all x ∈ (c − δ, c) and f 0 (x) > 0 for all x ∈ (c, c + δ), then f has a
local minimum at c.
so c is a local maximum.
52
Proposition 3.4 (Second Derivative Test for Extrema). Suppose f : [a, b] → R is
continuous and twice continuously differentiable on (a, b) (i.e. f 0 is differentiable and f 00
is continuous). Suppose c ∈ (a, b) is a critical point of f .
(i) If f 00 (c) < 0, then f has a local maximum at c.
(ii) If f 00 (c) > 0, then f has a local minimum at c.
Proof. We only prove part (ii) because part (i) is similar.
Since f 00 is continuous, it follows from Lemma 1.4 that there is δ > 0 such that
f 00 (x) > 0
for all x ∈ B(c, δ). By Theorem 3.1, it follows that f 0 is strictly increasing on [c−δ, c+δ].
However,
f 0 (c) = 0
So it follows that
(i) f 0 (x) < 0 for all x ∈ (c − δ, c)
(ii) f 0 (x) > 0 for all x ∈ (c, c + δ)
So by the First Derivative Test for Extrema (Proposition 3.2), we see that f is a local
minimum.
(End of Day 16)
Example 3.5. We (indicate how to) sketch the graph of f : [−5, 5] → R given by
f (x) = x4 − 4x3 + 10
53
4. Taylor’s Theorem
Definition 4.1. Let f : (a, b) → R be a differentiable function, so that f 0 : (a, b) → R
is defined.
(i) If f 0 is differentiable, we say that f is twice differentiable, and we define f (2) :=
f 00 := (f 0 )0 .
f (n) := (f (n−1) )0 .
(iii) Notice that the RHS is the equation of the tangent line at the point (x0 , f (x0 )).
So the Mean Value Theorem implies that the graph of the function ‘near’ x0 looks
like the tangent line.
(iv) Taylor’s theorem is a generalization of this idea for functions that have higher
order derivatives.
Definition 4.3. Let n ∈ N, and let f : (a, b) → R be a n-times differentiable function.
Given x0 ∈ (a, b) fixed, we define the nth order Taylor polynomial of f at x0 to be
54
(i) Note that Pn is a polynomial of degree ≤ n.
(ii) Moreover,
Pn (x0 ) = f (x0 )
Pn0 (x0 ) = f 0 (x0 )
Pn00 (x0 ) = f 00 (x0 )
..
.
Hence,
Pn(i) (x0 ) = f (i) (x0 ) for all 0 ≤ i ≤ n
Example 4.5.
Then, at x0 = 0, we have
P0 (t) = f (0) = 1
P1 (t) = f (0) + f 0 (0)(t − 0) = 1 + 2t
f 00 (0) 2
P2 (t) = f (0) + f 0 (0)t + t = 1 + 2t + 3t2
2!
P3 (t) = 1 + 2t + 3t2 + 4t3 = f (t)
f (x) = a0 + a1 x + . . . + an xn
55
Hence,
P0 (t) = f (0) = 1
P1 (t) = f (0) + f 0 (0)t = 1
t2
P2 (t) = 1 −
2!
P3 (t) = P2 (t)
t2 t4
P4 (t) = 1 + +
2! 4!
P5 (t) = P4 (t)
The graphs of these functions are shown here: https://www.geogebra.org/m/
s9SkCsvC.
(iv) Suppose f : (−1, 1) → R is given by
1
f (x) =
(1 − x)
Then,
1
f 0 (x) =
(1 − x)2
2
f 00 (x) =
(1 − x)3
..
.
n!
f (n) (x) =
(1 − x)n+1
So at x0 = 0, we get
P0 (t) = f (0) = 1
P1 (t) = f (0) + f 0 (0)t = 1 + t
P2 (t) = 1 + t + t2
..
.
Pn (t) = 1 + t + t2 + . . . + tn .
The next lemma is an analogue of Rolle’s Theorem (Theorem 2.4) for functions with
higher order derivatives.
Lemma 4.6. Suppose g : (a, b) → R is a (n + 1)-times differentiable function and
x0 , x ∈ (a, b) are fixed and distinct. Suppose that
g(x) = g(x0 ) = g 0 (x0 ) = g 00 (x0 ) = . . . = g (n) (x0 ) = 0.
Then, there is a point c between x0 and x such that
g (n+1) (c) = 0
56
Proof. We use the Mean Value Theorem (Theorem 2.5) repeatedly. Assume WLOG
that x0 < x.
(i) By the Mean Value Theorem (Theorem 2.5), there exists x1 ∈ (x0 , x) such that
g 0 (x1 ) = 0.
(ii) Now, g 0 (x0 ) = 0 so there exists x2 ∈ (x0 , x1 ) such that g 00 (x2 ) = 0.
(iii) Thus proceeding, we obtain points
x0 < xn < xn−1 < . . . < x1 < x
such that
g (i) (xi ) = 0 for all 0 ≤ i ≤ n.
(iv) Finally, there exists c := xn+1 ∈ (x0 , xn ) such that
g (n+1) (c) = 0
57
(iii) So by Lemma 4.6, there exists c ∈ (x0 , x) such that
g (n+1) (c) = 0.
58
VI. Integration
1. Lower and Upper Integrals
Throughout this section, we will fix a closed and bounded interval [a, b] ⊂ R and and a
bounded function f : [a, b] → R. i.e. There exist m, M ∈ R such that
m ≤ f (x) ≤ M
Definition 1.1.
Lemma 1.2. Let f : [a, b] → R be a bounded function and P be a partition of [a, b], and
let m := inf f ([a, b]) and M := sup(f [a, b]) as above. Then,
m ≤ mi ≤ Mi ≤ M
59
So
m
X
m(b − a) = m∆i
i=1
n
X
≤ mi ∆i = L(f, P)
i=1
Xn
≤ Mi ∆i = U(f, P)
i=1
n
X
≤ M ∆i = M (b − a)
i=1
Note: To understand what Riemann sums measure, see this link: https://www.
geogebra.org/m/Fv6t696j.
In the above setting, the set
Definition 1.3. The upper and lower Riemann integrals of f are defined by
mi = 0 and Mi = 1.
60
Definition 1.5. Let P and P e be two partitions of the interval [a, b]. We say that P
e is
a refinement of P (or P
e is finer than P) if P ⊂ P.
e
In other words, P
e must break the interval into more pieces than P and each subinterval
created by P
e is contained in a subinterval created by P.
Lemma 1.6. If P
e is a refinement of P, then
L(f, P) ≤ L(f, P)
e and U (f, P) ≥ U (f, P).
e
Proof. We prove the first inequality since the second is analogous. Write
and write n
X
L(f, P) := mi (xi − xi−1 )
i=1
L(f, P)
e − L(f, P) = α(y − xi−1 ) + β(xi − y) − mi (xi − xi−1 )
≥ mi (y − xi−1 ) + mi (xi − y) − mi (xi − xi−1 )
= 0.
⇒ L(f, P)
e ≥ L(f, P).
P1 = P ∪ {y1 }
P2 = P1 ∪ {y2 }
..
.
Pk = Pk−1 ∪ {yk } = P.
e
61
Remark 1.7. Given f : [a, b] → R as above, we had two sets
Af := {L(f, P) : P a partition of [a, b]}, and
Bf := {U(f, P) : P a partition of [a, b]}.
If P and P
e are two partitions of [a, b], then
S := P ∪ P
e
Lemma 1.8. Suppose A, B ⊂ R are two non-empty sets such that a ≤ b for any a ∈ A
and b ∈ B. Then, sup(A) and inf(B) both exist, and
sup(A) ≤ inf(B)
Proof. Fix b ∈ B, then by hypothesis, b is an upper bound for A. So sup(A) exists and
sup(A) ≤ b.
Now this holds for all b ∈ B, so B is bounded below by sup(A). Hence, inf(B) exists
and
sup(A) ≤ inf(B).
Proposition 1.9. For any bounded function f : [a, b] → R, write m := inf(f [a, b]) and
M := sup(f [a, b]). Then
m(b − a) ≤ L(f ) ≤ U(f ) ≤ M (b − a).
Proof. We know from Lemma 1.2 that for any partition P of [a, b], we have
m(b − a) ≤ L(f, P) ≤ U(f, P) ≤ M (b − a).
Therefore,
L(f ) = sup{L(f, P) : P a partition of [a, b]} ≥ m(b − a)
and similarly, U(f ) ≤ M (b − a) as well.
62
2. Integrable Functions
Definition 2.1. A bounded function f : [a, b] → R is said to be Riemann integrable if
Lba (f ) = Uab (f ).
Lemma 2.2. Let A, B ⊂ R be two non-empty sets such that a ≤ b for all a ∈ A and
b ∈ B. Then, sup(A) = inf(B) iff, for each > 0, there exist a ∈ A and b ∈ B such that
(b − a) < .
Proof. By Lemma 1.8, sup(A) and inf(B) both exist and sup(A) ≤ inf(B).
⇒: Suppose that sup(A) = inf(B) and fix > 0. Then, there exists a ∈ A such that
⇐: Suppose that the condition given above holds. We WTS that sup(A) = inf(B).
Suppose that sup(A) < inf(B), so that
63
(iii) There is a sequence of partitions Pn of [a, b] such that
In that case, Z b
f = lim U(f, Pn ) = lim L(f, Pn ).
a n→∞ n→∞
Proof.
(i) ⇒ (ii) : Suppose f is integrable and fix > 0. Then, by Lemma 2.2, there exist two
partitions P and P
e such that
U(f, P) − L(f, P)
e < .
Take S := P ∪ P,
e then by Lemma 1.6, we have
(ii) ⇒ (iii) : Suppose (ii) holds, then for each n ∈ N, there is a partition Pn such that
1
U(f, Pn ) − L(f, Pn ) < .
n
So it follows that
lim (U(f, Pn ) − L(f, Pn )) = 0.
n→∞
64
(iii) ⇒ (i) : Suppose (iii) holds. Take
so that L(f, P) = U(f, P). So the Integrability Criterion of Theorem 2.3 holds.
Definition 2.6. A function f : [a, b] → R is said to be
(i) increasing (or non-decreasing) if f (x) ≤ f (y) whenever x ≤ y.
65
Proof. Assume WLOG that f is increasing (the proof is similar for decreasing func-
tions).
(i) First note that f (a) ≤ f (x) ≤ f (b) for all x ∈ [a, b], so f is a bounded function.
(ii) Now we wish to verify the Integrability Criterion (Part (iii) of Theorem 2.3). Fix
n ∈ N, and choose a partition Pn = {x0 , x1 , x2 , . . . , xn } where
xi = a + i(b − a)/n
1
so that ∆i = (xi − xi−1 ) = n
for all 1 ≤ i ≤ n. Then, observe that
Therefore,
n
X
L(f, Pn ) = mi ∆i
i=1
n
1X
= f (xi−1 ), and similarly
n i=1
n
1X
U(f, Pn ) = f (xi ).
n i=1
n
1X
⇒ U(f, Pn ) − L(f, Pn ) = [f (xi ) − f (xi−1 )]
n i=1
f (b) − f (a)
=
n
⇒ lim (U(f, Pn ) − L(f, Pn )) = 0.
n→∞
(i) Since f is uniformly continuous by Theorem IV.5.7, there is δ > 0 such that for
all x, y ∈ [a, b],
|x − y| < δ ⇒ |f (x) − f (y)| < /(b − a).
(ii) Choose n ∈ N such that (b−a)/n < δ and consider the partition P = {x0 , x1 , . . . , xn }
given by
xi = a + i(b − a)/n.
Then, note that ∆i = (xi − xi−1 ) < δ for all i.
66
(iii) Consider
mi = inf(f [xi−1 , xi ]) and Mi = sup(f [xi−1 , xi ]).
Since [xi−1 , xi ] is a compact set, there are points yi , zi ∈ [xi−1 , xi ] such that
mi = f (yi ) and Mi = f (zi ).
This follows from the Extreme Value Theorem Corollary IV.3.4.
(iv) Since |yi − zi | ≤ (xi − xi−1 ) < δ, we have
Mi − mi < .
b−a
for all 1 ≤ i ≤ n. Therefore,
n
X n
X
U(f, P) − L(f, P) = Mi ∆i − mi ∆i
i=1 i=1
Xm
= (Mi − mi )∆i
i=1
n
X
< ∆i
(b − a) i=1
=
So by the Integrability Criterion (Part (ii) of Theorem 2.3), f is integrable.
67
(ii) We prove this by induction on m. Moreover, we may assume WLOG that n ≥ m.
(a) If m = 1, then this is true because n2 > n2 − 1 = (n + 1)(n − 1).
(b) Assume the result is true for (m − 1). i.e. We assume that
nm > (n + 1)m−1 (n − m + 1)
We wish to prove the result for m. Since n ≥ m, we have
nm+1 = n · nm
> n(n + 1)m−1 (n − m + 1)
= (n + 1)m−1 (n2 − nm + n)
≥ (n + 1)m−1 (n2 − nm + n − m)
= (n + 1)m−1 (n + 1)(n − m)
= (n + 1)m (n − m)
68
Example 2.10. Fix m ∈ N, b > 0 and let f : [0, b] → R be given by f (x) = xm . Then,
f is increasing so it is integrable by Proposition 2.7. Now consider the partition
Then
n−1
1X jb
L(f, Pn ) = f
n j=0 n
n−1
bm X m
= m+1 j
n j=0
bm nm+1
<
nm+1 m + 1
bm
=
m+1
Note that the middle inequality holds by Lemma 2.9. Similarly,
bm
U(f, Pn ) >
m+1
Therefore, we have
bm
L(f, Pn ) < < U(f, Pn )
m+1
By the Squeeze theorem (Theorem II.2.3), we see that
b
bm
Z
xm dx = .
0 m+1
Theorem 3.1 (Monotonicity). Suppose f, g ∈ I[a, b] are such that f (x) ≤ g(x) for all
x ∈ [a, b]. Then
Z b Z b
f≤ g
a a
69
Proof. Let P = {x0 , x1 , . . . , xn } be a partition of [a, b]. For each 1 ≤ i ≤ n, set
Hence,
U(f ) ≤ U(g, P)
This is true for every P, so U(f ) is a lower bound for the set {U(g, P) : P a partition of [a, b]}.
Hence,
U(f ) ≤ inf{U(g, P) : P a partition of [a, b]} = U(g)
Since f and g are integrable, it follows that
Z b Z b
f = U(f ) ≤ U(g) = g.
a a
Corollary 3.2. Let f ∈ I[a, b] and suppose M > 0 is such that |f (x)| ≤ M for all
x ∈ [a, b]. Then,
Z b
f (x)dx ≤ M (b − a)
a
Proof. By hypothesis,
−M ≤ f (x) ≤ M
for all x ∈ [a, b], so by applying Theorem 3.1, we see that
Z b
−M (b − a) ≤ f (x)dx ≤ M (b − a)
a
Proof.
(i)
70
⇒: Suppose f ∈ I[a, c], fix > 0 with a view to using Theorem 2.3. Then, there
is a partition P = {x0 , x1 , . . . , xn } of [a, c] such that
U(f, P) − L(f, P) <
If P
e = P ∪ {b}, then by Lemma 1.6,
U(f, P)
e − L(f, P)
e ≤ U(f, P) − L(f, P) < .
= U(f, P)
e − L(f, P)
e < .
71
As above, P := P1 ∪ P2 is a partition of [a, c] and
Z c
f ≤ U(f, P)
a
= U(f, P1 ) + U(f, P2 )
< L(f, P1 ) + L(f, P2 ) +
Z b Z c
≤ f+ f +
a b
Corollary 3.4. If f ∈ I[a, b] and [c, d] ⊂ [a, b], then f |[c,d] ∈ I[c, d].
Proof. Exercise.
Theorem 3.5. If f : [a, b] → R is a bounded function with finitely many discontinuities,
then f ∈ I[a, b].
Proof. Let F ⊂ [a, b] be the finite set so that f is continuous on [a, b] \ F .
(i) Assume first that F = {x} is a singleton. Fix > 0 and choose M > 0 so that
|f (x)| ≤ M for all x ∈ [a, b]. Let δ := /12M , then on the subinterval [x − δ, x + δ],
we have
` := inf(f [x − δ, x + δ]) ≥ −M and L := sup(f [x − δ, x + δ]) ≤ M
Therefore,
(L − `)(2δ) ≤ 4M δ ≤ /3.
Now note that f ∈ I[a, x − δ] ∩ I[x + δ, b] because it is continuous on both subin-
tervals (by Theorem 2.8). So there is a partition P1 of [a, x − δ] and P2 of [x + δ, b]
such that
U(f, P1 ) − L(f, P1 ) < /3
U(f, P2 ) − L(f, P2 ) < /3.
Now consider P = P1 ∪ P2 treated as a partition of [a, b]. Then,
U(f, P) = U(f, P1 ) + U(f, P2 ) + L(2δ), and
L(f, P) = L(f, P1 ) + L(f, P2 ) + `(2δ)
⇒ U(f, P) − L(f, P) < /3 + /3 + /3 =
Hence, f ∈ I[a, b] by Theorem 2.3.
72
(ii) Now suppose F = {x1 , x2 , . . . , xn } is any finite set. Again, assume x1 < x2 < . . . <
xn and consider the subintervals
Theorem 3.6 (Linearity - I). Let f ∈ I[a, b] and α ∈ R. Then, αf ∈ I[a, b] and
Z b Z b
(αf ) = α f
a a
Proof.
(i) First assume α ≥ 0. Let P = {x0 , x1 , . . . , xn } be a partition of [a, b]. For each
1 ≤ i ≤ n, set
Since α ≥ 0, we have
Ki = αMi
by Homework 1.2. Hence,
n
X n
X
U(αf, P) = Ki ∆i = α Mi ∆i = αU(f, P).
i=1 i=1
Hence,
Similarly,
L(αf ) = αL(f ).
Since f ∈ I[a, b], we have
73
(ii) To prove it for α < 0, it suffices to prove it for α = −1. In that case, consider a
partition P as before. Then, for each 1 ≤ i ≤ n,
sup{αf (x) : xi−1 ≤ x ≤ xi } = − inf{f (x) : xi−1 ≤ x ≤ xi }
Hence,
U(αf, P) = −L(f, P)
Taking an infimum, we have
U(αf ) = inf{U(f, P) : P is a partition of [a, b]}
= − sup{L(f, P) : P is a partition of [a, b]}
= −L(f ).
Similarly,
L(αf ) = −U(f ).
Therefore, U(αf ) = L(αf ), so αf ∈ I[a, b] and
Z b Z b Z b
(αf ) = U(αf ) = −L(f ) = − f =α f.
a a a
Theorem 3.7 (Linearity - II). Let f, g ∈ I[a, b], then (f + g) ∈ I[a, b] and
Z b Z b Z b
(f + g) = f+ g
a a a
Proof.
(i) Let P = {x0 , x1 , . . . , xn } be a partition of [a, b] and for each 1 ≤ i ≤ n, set
Mi = sup(f [xi−1 , xi ])
Ki = sup(g[xi−1 , xi ]), and
Li = sup((f + g)[xi−1 , xi ]).
Then, for any x ∈ [xi−1 , xi ], we have
(f + g)(x) = f (x) + g(x) ≤ Mi + Ki
⇒ Li ≤ Mi + Ki
Xn
⇒ U(f + g, P) = Li ∆i
i=1
Xn n
X
≤ Mi ∆i + Ki ∆i
i=1 i=1
= U(f, P) + U(g, P)
74
(ii) Now let P and P
e be two partitions of [a, b], set S := P ∪ P.
e Then, by part (i) and
Lemma 1.6, we have
Now,
mi := inf(f [xi−1 , xi ]) ≤ f (ci ) ≤ sup(f [xi−1 , xi ]) =: Mi .
75
Therefore,
n
X
L(f, P) = mi ∆i
i=1
Xn
≤ f (ci )∆i
i=1
n
X
≤ Mi ∆i
i=1
= U(f, P).
Xn
⇒ L(f, P) ≤ (F (xi ) − F (xi−1 )) ≤ U(f, P)
i=1
⇒ L(f, P) ≤ F (b) − F (a) ≤ U(f, P).
Taking a supremum on the LHS and infimum on the RHS, we conclude that
Example 4.2. We now give a short proof of Example 2.10. For m ∈ N, consider
f : [a, b] → R be given by f (x) = xm . Then, let F : [a, b] → R be given by
xm+1
F (x) =
m+1
Then, F 0 (t) = f (t) for all t ∈ [a, b], so by Theorem 4.1,
b
bm+1 − am+1
Z
xm dx = F (b) − F (a) = .
a m+1
(ii) If f is continuous at a point c ∈ [a, b], then F is differentiable at c and F 0 (c) = f (c).
Proof.
76
(i) Since f is bounded, there is M > 0 such that |f (x)| ≤ M for all x ∈ [a, b]. So for
any a ≤ x ≤ y ≤ b it follows from Corollary 3.2 that
Z y
f (t)dt ≤ M |y − x|
x
Hence, Z y
f = F (y) − F (x).
x
(ii) Suppose f is continuous at c, and fix > 0. We wish to prove that there is a δ > 0
such that whenever |x − c| < δ,
F (x) − F (c)
− f (c) < .
x−c
77
Therefore, in either case, we see that when |x − c| < δ,
F (x) − F (c)
− f (c) ≤
x−c
Theorem 4.4 (Mean Value Theorem for Integrals). Suppose f : [a, b] → R is continuous
on [a, b]. Then, there exists c ∈ [a, b] such that
Z b
1
f (c) = f (t)dt.
(b − a) a
Proof. Homework.
(i) If P and Q are two primitives of [a, b], then there is a constant C ∈ R such that
P (x) − Q(x) = C
Proof.
(P − Q)0 (x) = 0
for all x ∈ (a, b). So the result follows from the Mean Value Theorem (specifically,
Theorem V.3.1).
78
(ii) Let F : [a, b] → R be given by
Z x
F (x) = f (t)dt.
a
Remark 4.7.
(i) The Fundamental Theorem of Calculus (specifically Corollary 4.6) is so useful, it
is usually taught in school as the definition of the integral. i.e. The integral of a
function is identified with its anti-derivative and Equation VI.1 is used to compute
integrals.
(ii) Leibniz used the symbol Z
f (t)dt = P (x) + C (VI.2)
79
Then, by the Fundamental Theorem of Calculus, Q is differentiable on (a, b) and for any
x ∈ (a, b)
Q0 (x) = f (g(x))g 0 (x)
Moreover, h := P ◦ g is also differentiable and for any x ∈ (a, b)
Q(x) − h(x) = C
C = Q(a) − h(a) = 0 − 0.
(iii) [Functional Equation] ln(ab) = ln(a) + ln(b) for all a, b ∈ (0, ∞).
Proof.
(i) Obvious.
(ii) This follows by the Fundamental Theorem of Calculus (Theorem 4.1 and Theo-
rem 4.3) since ln(1) = 0.
80
(iii) Note that
Z ab Z a Z ab
1 1 1
ln(ab) = dt = dt + dt
1 t 1 t a t
by the Additivity of the integral (Theorem 3.3). Consider g : [1, b] → [a, ab] given
by g(s) := as. Then, by the Substitution Theorem (Theorem 4.8), we see that
Z ab Z g(b) Z b
1 1 1
dt = dt = · adt = ln(b)
a t g(1) t 1 at
(ii) Uniqueness: Note that ln0 (x) = 1/x for all x ∈ R, so ln is strictly increasing on R+
(by Theorem V.3.1). So if a1 < a2 , then ln(a1 ) < ln(a2 ). This implies uniqueness.
(ii) exp(0) = 1.
81
(iii) [Functional Equation] exp(a + b) = exp(a) exp(b) for all a, b ∈ R.
Proof.
(iv) Suppose a, b ∈ R are such that a < b. Choose c, d ∈ R+ so that ln(c) = a and
ln(d) = b. If c ≥ d, then since ln is strictly incerasing, it would follow that
a = ln(c) ≥ ln(d) = b.
This is not the case, so it must happen that c < d, so exp(a) < exp(b).
(v) Fix c ∈ R+ and > 0. We may assume that is small enough so that (c−, c+) ⊂
R+ . We wish to find a δ > 0 so that
Note that δ > 0 because ln is strictly increasing. Now suppose x ∈ R is such that
|x − c| < δ. Then,
Similarly, exp(x) > exp(c) − as well. Hence, | exp(x) − exp(c)| < as desired.
82
(vi) Fix c ∈ R and consider the difference quotient
exp(xn ) − 1
lim = 1.
n→∞ xn
Choose yn ∈ R+ such that ln(yn ) = xn , so that exp(xn ) = yn . Since exp is
continuous by part (v), we know that
lim yn = exp(0) = 1.
n→∞
exp(xn ) − 1 yn − 1
lim = lim
n→∞ xn n→∞ ln(yn )
1
= lim
n→∞ ln(yn )−ln(1)
yn −1
1
= lim 0 = 1.
n→∞ ln (1)
Remark 5.6.
(i) By Proposition 5.3, there is a unique positive real number e ∈ R such that ln(e) =
1, so that exp(1) = e.
(ii) We now consider the function equation exp(a + b) = exp(a) exp(b). Since exp(1) =
e, it follows that for any n ∈ N,
83
(iv) Therefore, we may combine parts (i) and (ii) and see that
exp(n) = en
Now exp(1/n) ∈ R+ , so by the uniqueness of the nth root (see Proposition I.4.6),
it follows that
exp(1/n) = e1/n .
(i) ex := exp(x).
Remark 5.8.
(i) Note that this definition agrees with the conclusion of Remark 5.6 for rational x,
but is simply the definition for irrational x.
(ii) For any a > 0, we have some basic properties which may be proved using Propo-
sition 5.5: For any a, b > 0 and x, y ∈ R,
(a) ln(ax ) = x ln(a).
(b) (ab)x = ax bx .
(c) ax ay = ax+y .
(d) (ax )y = axy = (ay )x .
ln(y)
(e) If a 6= 1, then y = ax if and only if x = ln(a)
.
84
VII. Sequences of Functions
1. Pointwise and Uniform Convergence
Definition 1.1. Let A ⊂ R be set and let (fn ) be a sequence of functions on A (i.e.
For each n ∈ N, fn : A → R is a function). We say that (fn ) converges pointwise to a
function f : A → R if, for each x ∈ A,
p
If this happens, we write fn →
− f.
Remark 1.2.
p
(i) fn →
− f if and only if, for each x ∈ A and each > 0, there exists N ∈ N such that
for all n ≥ N .
fn (1) = 1
lim fn (x) = 0
n→∞
(iii) Consider the same functions as in part (ii) and consider two points x = 3/4 and
y = 1/4 and let = 1/4. Then,
(a)
1
f2 (y) = <
16
so that for all n ≥ 2, |fn (y) − f (y)| < .
85
(b) However,
f4 (x) ≈ 0.31, and f5 (x) ≈ 0.23
so that for all n ≥ 5, |fn (x) − f (x)| < .
Therefore, the value of N chosen (as in part (i)) depends on the choice of x and .
Definition 1.3. Let A ⊂ R be a set and (fn ) be a sequence of functions on A. We say
that (fn ) converges uniformly to a function f : A → R if for each > 0, there exists
N ∈ N such that
|fn (x) − f (x)| <
u
for all x ∈ A. If this happens, we write fn →
− f.
Remark 1.4.
(i) Note that in the definition above, given > 0 the value of N chosen does not depend
on the point x ∈ A. Therefore, Definition 1.3 is stronger than Definition 1.1.
u p
(ii) In particular, if fn →
− f , then fn →
− f must hold. However, the converse does not
hold.
u
Lemma 1.5. Suppose each fn is a continuous on A and fn →
− f , then f is continuous
on A.
Proof. Fix c ∈ A and > 0. We wish to find a δ > 0 so that
|x − c| < δ ⇒ |f (x) − f (c)| < .
To do this, choose N ∈ N so that
|fn (x) − f (x)| < /3
for all n ≥ N and all x ∈ A. Now, fN is continuous, so there is a δ > 0 such that
|x − c| < δ ⇒ |fN (x) − fN (c)| < /3.
Then,
|f (x) − f (c)| ≤ |f (x) − fN (x)| + |fN (x) − fN (c)| + |fN (c) − f (c)|
< + +
3 3 3
= .
So f is continuous at c. This is true for any c ∈ A, so f is continuous on A.
Example 1.6. Consider fn : [0, 1] → R given by fn (x) := xn . Then, we saw in
Remark 1.2 that (fn ) converges pointwise to
(
0 : if 0 ≤ x < 1
f (x) :=
1 : if x = 1
86
2. Distances between Functions
Remark 2.1.
d(x, y) := |x − y|.
(ii) Given a set A ⊂ R and two functions f, g : A → R, we may measure the distance
between them as
d∞ (f, g) := sup |f (x) − g(x)|.
x∈A
(iii) Alternatively, if A = [0, 1] is an interval and f and g are integrable, we may define
it as the ‘area’ between the two graphs, by
Z 1
d1 (f, g) := |f (t) − g(t)|dt.
0
(iv) There are many different definitions like these, but we will focus on one that
depends on one crucial fact (see Theorem IV.3.3) that every continuous function
on a closed interval is bounded.
(i) We define
C[a, b] := {f : [a, b] → R : f is continuous}.
In other words, each element of C[a, b] is a function.
Note that the is a well-defined real number because f and g are both bounded
functions.
u
Proposition 2.3. For a sequence (fn ) ⊂ C[a, b], we have fn →
− f if and only if, for
each > 0, there exists N ∈ N such that
d(fn , f ) <
for all n ≥ N .
In other words, uniform convergence of a sequence is captured by this distance d(·, ·).
Proof. This follows from Definition 1.3 and Definition 2.2.
87
(End of Day 25)
Proposition 2.4. For any three functions f, g, h ∈ C[a, b], we have
(i) d(f, g) ≥ 0.
(ii) d(f, g) = 0 holds if and only if f (x) = g(x) for all x ∈ [a, b].
(iii) d(f, g) = d(g, f ).
(iv) [Triangle Inequality] d(f, g) ≤ d(f, h) + d(g, h).
A function that satisfies all these conditions is called a metric and (C[a, b], d) is called a
metric space.
Proof.
(i) Obvious because d(f, g) is the supremum of a set of non-negative real numbers.
(ii) Obvious because d(f, g) = 0 holds if and only if |f (x) − g(x)| = 0 for all x ∈ [a, b].
(iii) Obvious.
(iv) Note that for any x ∈ [a, b], we have
|f (x) − g(x)| ≤ |f (x) − h(x)| + |h(x) − g(x)| ≤ d(f, h) + d(g, h)
So the RHS is an upper bound for the set {|f (x) − g(x)| : x ∈ [a, b]}, so it follows
that
d(f, g) = sup |f (x) − g(x)| ≤ d(f, h) + d(g, h).
x∈[a,b]
Note that this is a well-defined real number and it satisfies all the properties of
Proposition 2.4.
Remark 2.6.
(i) Note that C[a, b] ⊂ B[a, b].
(ii) If (fn ) in C[a, b] converges uniformly to a function f ∈ B[a, b], then f ∈ C[a, b] by
Lemma 1.5.
(iii) You will study metric spaces next semester. In that context, the distance function
d is a metric on C[a, b] and on B[a, b]. Moreover, C[a, b] is a subspace of B[a, b]
and it is closed (it contains all its limit points).
88
3. Weierstrass’ Approximation Theorem
Remark 3.1. We introduce a little probability theory. If you have not seen any prob-
ability theory before, you may skip this remark and take Equation VII.3 (see below) as
a fact.
(i) Fix a real number p ∈ [0, 1], and consider an experiment with two possible out-
comes: A success occurs with probability p, and a failure occurs with probability
(1 − p). Our experiment is modelled by a sample space (the set of all possible
outcomes),
Ω := {0, 1}
where 0 represents a failure and 1 represents a success, together with a probability
function P : P(Ω) → R given by
P ({0}) = (1 − p)
P ({1}) = p
P (∅) = 0
P (Ω) = 1
(ii) Fix n ∈ N, and suppose this experiment is repeated n times. Then, the sample
space becomes
Ω := {(a1 , a2 , . . . , an ) : ai ∈ {0, 1}}
and the probability function P : P(Ω) → R is given by
P (∅) = 0, and
P ({(a1 , a2 , . . . , an )}) = pk (1 − p)n−k
where k is the number of 1’s occurring in (a1 , a2 , . . . , an ), with the understanding
that for any A ⊂ Ω, X
P (A) = P ({ω}).
ω∈A
89
(iv) A random variable is any function Z : Ω → R. In our case, we will be interested
in two random variables:
(a) X = Xn,p defined by
n
X
X(a1 , a2 , . . . , an ) = the number of successes in (a1 , a2 , . . . , an ) = ai .
i=1
(vi) Given a random variable Z, there are two useful real numbers associated to it, the
mean µZ and the variance σZ2 . We will not define the variance here, but the mean
is given by X
µZ = E(Z) = zP (Z = z) (VII.2)
z∈R
(vii) In our situation, we will need two facts concerning the random variable Y = Yn,p :
p(1−p)
(a) µY = p and σY2 = n
.
(b) We also need Chebyshev’s inequality: For δ > 0, define
D := {ω ∈ Ω : |Y (ω) − p| > δ}
90
Then,
σY2 p(1 − p)
P (D) = P (|Y − µY | > δ) < 2
=
δ nδ 2
Moreover, the function p 7→ p(1 − p) has its maximum value at p = 1/2 (by
Calculus), so
1
P (|Yn,p − p| > δ) < (VII.3)
4nδ 2
Lemma 3.2. Fix δ > 0, p ∈ [0, 1], and n ∈ N, let
k
S := k ∈ {0, 1, . . . , n} : −p >δ
n
Then,
X n 1
pk (1 − p)n−k <
k∈S
k 4nδ 2
Proof. Let Y = Yn,p be the random variable defined in Remark 3.1, and define the set
D := {ω ∈ Ω : |Y (ω) − p| > δ}
so that
1
P (D) <
4nδ 2
by Chebyshev’s inequality. However, ω = (a1 , a2 , . . . , an ) ∈ D if and only if |Y (ω) − p| >
δ. For each 0 ≤ k ≤ n, let
Moreover,
D = (D ∩ C0 ) t (D ∩ C2 ) t . . . t (D ∩ Cn ).
Hence,
n
X
P (D) = P (D ∩ Bk )
k=0
X n
= pk (1 − p)n−k
k∈S
k
91
Remark 3.3. Suppose f : [0, 1] → R is a function. For n ∈ N and p ∈ [0, 1], let Y = Yn,p
be the random variable defined above.
(i) Since Y (ω) ∈ [0, 1] for all ω ∈ Ω, we get a new random variable f (Y ) : Ω → R
given by f ◦ Y . i.e.
f (Y )(ω) = f (Y (ω))
for all ω ∈ Ω.
(ii) We need one important fact about the mean of f (Y ). It is given by the expression
X
E(f (Y )) = f (y)P (Y = y)
y∈R
Definition 3.4. Let f : [0, 1] → R be a function, and let n ∈ N. The nth Bernstein Polynomial
associated to f is defined as
n
X k n k
Bn (x) := f x (1 − x)n−k
k=0
n k
92
then E ≤ E1 + E2 , where
X
k n k
E1 := f (x) − f x (1 − x)n−k , and
n k
k∈S
/
X
k n k
E2 := f (x) − f x (1 − x)n−k
k∈S
n k
by Equation VII.1.
(ii) We now need to estimate E2 . Note that there exists M > 0 such that
|f (x)| ≤ M
Choose N ∈ N so that
M
< .
N δ2
Then, by Lemma 3.2, for all n ≥ N , we would have
E2 < /2.
Observe that the choice of N only depends on and δ. Since δ only depends on
(and not on x, because of uniform continuity), it follows that
n
X k n k
f (x) − f x (1 − x)n−k <
k=0
n k
93
Corollary 3.6 (Weierstrass’ Approximation Theorem). For any continuous function
u
f : [a, b] → R, there is a sequence (pn ) of polynomials such that pn →
− f.
ϕ(t) := a(1 − t) + tb
94
VIII. Series of Functions
1. Series of Real Numbers
Definition 1.1. Let (an ) be a sequence of real numbers.
k
X
sk := ai .
i=1
(ii) The sequence (sk ) is called the (infinite) series associated to (an ). The series is
often denoted by
∞
X
“ ai ”
i=1
Note that this is only a symbol, and really represents the limit
“ lim sk ”
k→∞
(iii) If the sequence (sk ) converges to s ∈ R, then we say that the series is convergent,
and we write ∞
X
s := ai .
i=1
If the sequence (sk ) does not converge, we say that the series is divergent.
Proposition
P∞ 1.3 (Geometric Series). If −1 < r < 1 and an := rn , then the series
i=1 ai converges.
95
Proof. Define the partial sums by
k
1 − rk+1
X r
sk := ai = r = (1 − rk+1 ).
i=1
1−r 1−r
P∞ P∞
Proposition 1.4. If i=1 ai = a and i=1 bi = b are two convergent series, then for
any α ∈ R,
∞
X
(αai + bi ) = αa + b
i=1
(ii) limn→∞ Tn = 0.
The sequence (Tn ) defined above is called the tail of the series.
Proof. Consider the parial sums
k
X
sk := ai
i=1
(i) Fix n ∈ N, and consider the partial sums of the series defining Tn :
k
X
tk := an+i
i=1
96
We WTS that (tk ) is convergent, so observe that
k
X
tk = an+i = sn+k − sn−1
i=1
since (sn+k )∞
k=1 is a subsequence of (sk ). In other words,
Tn = a − sn−1 .
2. Tests of Convergence
P∞
Theorem 2.1 (Test of Divergence). If a series i=1 ai is convergent, then limn→∞ an =
0.
Proof. Write
k
X
sk := ai
i=1
an = sn − sn−1
Example 2.2.
P∞
(i) If an = 1 for all n ∈ N, then i=1 ai diverges.
97
Theorem 2.3 (Integral Test). Let f : [1, ∞) → R be a non-negative, decreasing func-
tion. For each k ∈ N, define
k
X
sk := f (i)
i=1
Z k
tk := f (x)dx.
1
Then, the sequence (sk ) is convergent if and only if the sequence (tk ) is convergent.
so both sequences are increasing. By the Monotone Sequence Theorem (Theorem II.2.5),
each sequence is convergent if and only if it is bounded above.
On the interval [1, k], f is Riemann Integrable since it is monotone (by Proposition VI.2.7).
Consider the partition P = {1, 2, . . . , k}. Since f is decreasing,
k−1
X
U(f, P) = f (i) = sk−1
i=1
Xk
L(f, P) = f (i) = sk − f (1)
i=2
Now,
(i) If (sk ) is convergent, it is bounded above. In that case, (tk ) is also bounded above.
Since (tk ) is monotone increasing, it must be convergent by the Monotone Sequence
Theorem (Theorem II.2.5).
(ii) Similarly, if (tk ) is convergent, it is bounded above. The same argument as above
shows that (sk ) must also be convergent.
Example 2.4.
98
Observe that f is non-negative and decreasing. Moreover,
Z k
tk = f (x)dx = ln(k)
1
P∞
so (tk ) is unbounded. By the Integral Test, i=1 ai diverges.
(ii) If an = 1/n2 , then consider f : [1, ∞) → R given by
1
f (x) = .
x2
Again f is non-negative and decreasing. However, by FTOC,
Z k
−1 k 1
tk = f (x)dx = |1 = 1 − .
1 x k
Since (tk ) converges, it follows from the Integral Test that ∞
P
i=1 ai converges as
well.
Definition 2.5. A series ∞
P P∞
i=1 ai is said to be absolutely convergent if the series i=1 |ai |
is convergent.
Theorem 2.6. If ∞
P
i=1 ai is absolutely convergent, then it is convergent.
99
Theorem 2.7 (Comparison Test). Suppose (an ) and (bn ) are two sequences with non-
negative entries (i.e. an ≥ 0 and bn ≥ 0 for all n ∈ N). Suppose that
Then, ∞
P
i=1 ai converges as well.
Proof. We may assume WLOG that an ≤ bn for all n ∈ N (see Remark 1.2). Consider
the partial sums
k
X
sk := ai
i=1
Xk
tk := bi
i=1
sk ≤ tk
for all k ∈ N, and limn→∞ tk =: t exists. By the Monotone Sequence Theorem (The-
orem II.2.5), we know that (tk ) is bounded above (indeed, t = sup{tk : k ∈ N}).
Therefore,
sk ≤ t
for all k ∈ N, so (sk ) is bounded above. By the Monotone Sequence theorem, (sk ) is
convergent, which proves the result.
Theorem 2.8 (Ratio Test). Suppose (an ) is a non-negative sequence of real numbers
such that
an+1
lim =: L
n→∞ an
exists.
Proof.
100
(i) Suppose L < 1, fix a number r ∈ (L, 1). Then, there exists N ∈ N such that, for
all n ≥ N , we have
an+1
<r
an
an+1 an
⇒ n+1 < n
r r
So the sequence
aN aN +1
, ,...
rN rN +1
is a decreasing sequence, so that
an aN
≤
rn rN
aN
for all n ≥ N . Hence, if C := rN
, then
an ≤ Crn
P∞
Since
P∞ i=1 Cri converges, it follows from the Comparison Test (Theorem 2.7) that
i=1 ai also converges.
101
3. Power Series
Definition 3.1. Let (fn ) be a sequence of functions on a set A ⊂ R.
(i) For each k ∈ N, define the k th partial sum as gk : A → R given by
k
X
gk (x) = fi (x).
i=1
(ii) The sequence (gk ) is called the series associated to (fn ). The series is often denoted
by
X∞
“ fi ”
i=1
Note that, by Remark VII.1.4, if the series converges uniformly, then it converges
pointwise.
Theorem 3.2 (Weierstrass’ M-Test). Suppose (fn ) is a sequence of functions on a set
A ⊂ R. Suppose that there exist real numbers Mn ≥ 0 such that
(i) 0 ≤ |fn (x)| ≤ Mn for all x ∈ A and all n ∈ N.
P∞
(ii) i=1 Mi converges in R.
Proof.
(i) By condition (ii) and the Comparison Test (Theorem 2.7),
∞
X
|fi (x)|
i=1
102
(ii) Now consider
n
X ∞
X
f (x) − fi (x) = fi (x)
i=1 i=n+1
X∞
≤ |fi (x)|
i=n+1
X∞
≤ Mi
i=n+1
The RHS is now the tail of convergent series (see Proposition 1.5), so for a given
> 0, there exists N ∈ N such that
∞
X
Mi <
i=n+1
holds for all n ≥ N and for all x ∈ A. Therefore, the series converges uniformly
to f .
Definition 3.3. Fix an interval I := (−R, R) ⊂ R and a sequence (cn ) of real numbers.
For each n ∈ N, define fn : I → R by
fn (x) := cn xn .
If the series ∞
X
ci x i
i=1
103
For any fixed x ∈ R, we apply the Ratio Test.
|an+1 | |x|
lim = lim =0
n→∞ |an | n→∞ n + 1
(ii) Similarly,
∞
X (−1)i x2i+1
s(x) := , and
i=0
(2i + 1)!
∞
X (−1)i x2i
c(x) :=
i=0
(2i)!
converge pointwise for each x ∈ R.
(End of Day 30)
Theorem 3.5. Suppose a power series
∞
X
ci x i
i=1
104
(ii) Moreover, if 0 < S < R, then let
PA := (−S, S). Then, if tP
:= S/R, then the above
argument shows that the series i=1 ci x is dominated by ∞
∞ i i
i=1 t . The Weierstrass
M-Test (Theorem 3.2) now says that the series converges uniformly on A.
Note that if x = R or x = −R, we cannot say anything conclusive about the convergence
of the series.
Proof. Define
A := {|x| : the series converges at x}
Then, A ⊂ R≥0 and A is non-empty because |y1 | ∈ A.
We know that the series does not converge at y2 . So if x ∈ R such that |x| > |y2 |, then
the series cannot converge at x by Theorem 3.5. Therefore, A ⊂ [0, |y2 |].
R := sup(A)
exists.
(i) If z ∈ R is such that |z| < R, then there exists x ∈ R such that the series converges
at x and |z| < |x|. So by Theorem 3.5, the series must converge absolutely at z.
(ii) If z ∈ R is such that |z| > R, then the series cannot converge at z because otherwise
this would force |z| ∈ A, contradicting the definition of R.
Definition 3.7. This number R described in Corollary 3.6 is called the radius of convergence
of the power series. We adopt the following convention:
105
(End of Day 31)
P∞ i
Corollary 3.8. Consider a power series i=1 ci x as above, and assume that
cn+1
L := lim
n→∞ cn
Proof.
0, then let an := cn xn , then applying the Ratio Test (Theorem 2.8) to the
(i) If L 6=P
series ∞ i=1 ai , we wish to determine when
an+1
1 > T := lim = L|x|.
n→∞ an
Hence, the series converges whenever L|x| < 1, so it converges on the interval
(−1/L, 1/L)
(ii) If L = 0, then the same argument shows that the series converges for each x ∈ R,
so R = ∞.
Example 3.9.
Here, ci = i!1 , so
cn+1
L := lim =0
n→∞ cn
so R = +∞ (i.e. the power series converges at each x ∈ R).
106
Here, ci = 1 for all i ∈ N, so
cn+1
L = lim =1
n→∞ cn
so R = 1 as well. So the power series converges on (−1, 1) but diverges on
(−∞, −1) ∪ (1, ∞).
(a) At x = 1, the series clearly diverges.
(b) At x = −1, the series diverges since the sequence of partial sums does not
converge (Verify this!).
Hence the series converges on (−1, 1).
(iii) Consider the power series
∞
X xi
i=1
i
Here, ci = 1/i for all i ∈ N, so
cn+1
L = lim =1
n→∞ cn
so R = 1 as well. So the power series
(a) converges on (−1, 1),
(b) diverges on (−∞, −1) ∪ (1, ∞),
(c) At x = 1, the series diverges by Example 2.4,
(d) At x = −1, the series converges (we have not proved this, but it is a con-
sequence of the Alternating Series Test - see [Apostol Calculus, Theorem
10.14]).
Hence, the series converges on [−1, 1).
Remark 3.10. All the results for a power series centered at 0 also hold for power series
centered at other points.
(i) Suppose a power series
∞
X
ci (x − a)i
i=1
107
(iii) Every such power series has a radius of convergence R, which defines aan interval
I := (a − R, a + R) where the series converges, and the set
A := (−∞, a − R) ∪ (a + R, ∞)
where the series diverges. We cannot determine how the series behaves at the
points (a − R) or (a + R).
4. Taylor’s Series
Remark 4.1. Let I := (a, b) be a fixed open interval in R. Given a function f : I → R,
when can f be expressed as a power series? To answer this, we recall Taylor’s Theorem
(Theorem V.4.7):
f (n+1) (c)
f (x) = Pn (x) + (x − x0 )n+1 .
(n + 1)!
108
for all n ≥ N and all x ∈ I. To see this, we use Taylor’s Theorem (Theorem V.4.7) to
write
f (n+1) (c)
f (x) − Pn (x) = (x − x0 )n+1
(n + 1)!
for some point c ∈ I. By hypothesis,
An+1
|f (x) − Pn (x)| ≤ (b − a)n+1 (VIII.1)
(n + 1)!
The corresponding series
∞
X Ai (b − a)i
i=1
i!
converges by the ratio test (see Example 3.4), so by the Test of Divergence (Theorem 2.1),
An+1
lim (b − a)n+1 = 0.
n→∞ (n + 1)!
109
(iii) Similarly, if
eix + e−ix
f (x) = cos(x) =
2
Then,
x2 x4 x6
f (x) = 1 − + − + ...
2! 4! 6!
(iv) Note that in each example above, the series converges uniformly by Theorem 3.5.
(End of Day 32)
(v) We now give an example to show how Theorem 4.3 may fail. Define f : (−1, 1) → R
by ( 2
e−1/x : if x 6= 0
f (x) =
0 : if x = 0
Then, f is differentiable and
( 2
2x−3 e−1/x : if x 6= 0
f 0 (x) =
0 : if x = 0.
and this converges pointwise to f (x) on I. Therefore, the Taylor’s series may
converge even if the hypothesis of Theorem 4.3 is not satisfied.
110
5. Review
Remark 5.1. Some important results/ideas:
(i) Chapter 1:
(a) There is no x ∈ Q such that x2 = 2 ( Theorem I.1.1.
(b) Least Upper Bound Property (Definition I.3.3).
(c) Archimedean Property of R (Theorem I.4.3).
(d) Density of Q in R (Theorem I.4.5).
(e) Existence and Uniqueness of roots (Proposition I.4.6).
(ii) Chapter 2:
(a) Definition of convergence of a sequence (Definition II.1.3).
(b) Algebra of Limits (Theorem II.1.8).
(c) Limits and order (Proposition II.2.1).
(d) Monotone Sequence Theorem (Theorem II.2.5)
(e) Bolzano-Weierstrass Theorem (Theorem II.3.4)
(f) Cauchy Criterion for convergence of a sequence (Theorem II.4.5).
(iii) Chapter 3:
(a) Definition of Open and Closed sets (Definition III.1.2 and Definition III.2.4).
(b) Closure of a set (Theorem III.3.4).
(c) Definition of Compact set (Definition III.4.1).
(d) Heine-Borel Theorem (Theorem III.4.4).
(iv) Chapter 4:
(a) Epsilon-Delta Definition of Limit (Theorem IV.1.5).
(b) Definition of Continuous Function (Theorem IV.2.2).
(c) Continuous Functions on a Compact set and Extreme Value Theorem (Corol-
lary IV.3.4).
(d) Intermediate Value Theorem (Theorem IV.4.2)
(e) Compactness and Uniform Continuity (Theorem IV.5.7).
(v) Chapter 5:
(a) Rules of Differentiation (Theorem V.1.5 and Theorem V.1.6).
(b) Mean Value Theorem (Theorem V.2.5).
(c) First and Second Derivative Tests for Extrema (Proposition V.3.2 and Propo-
sition V.3.4).
111
(d) Taylor’s Theorem (Theorem V.4.7).
(vi) Chapter 6:
(a) Definition of Riemann Sums (Definition VI.1.1)
(b) Riemann Integrability Criterion (Theorem VI.2.3).
(c) Continuous Functions are Integrable (Theorem VI.2.8).
(d) Function with finitely many discontinuities is Integrable (Theorem VI.3.5)
(e) Fundamental Theorem of Calculus (Theorem VI.4.1 and Theorem VI.4.3).
(f) Properties of Logarithm and Exponential Function (Proposition VI.5.2 and
Proposition VI.5.5)
(vii) Chapter 7:
(a) Uniform limit of continuous functions (Lemma VII.1.5)
(b) Uniform convergence in terms of supremum metric (Proposition VII.2.3).
(c) Weierstrass’ Approximation Theorem (Corollary VII.3.6).
(viii) Chapter 8:
(a) Definition of convergence of series (Definition 1.1).
(b) Tail of convergent series (Proposition 1.5).
(c) Test of Divergence (Theorem 2.1).
(d) Integral Test (Theorem 2.3)
(e) Comparison Test (Theorem 2.7).
(f) Ratio Test (Theorem 2.8).
(g) Definition of Convergence of a series of functions (Definition 3.1)
(h) Weierstrass’ M-Test (Theorem 3.2).
(i) Convergence of Power Series (Theorem 3.5) and Radius of Convergence.
(j) Taylor’s Series examples (Example 4.4).
112
IX. Instructor Notes
(i) Instead of following Rudin, I decided to take a more accessible approach by fol-
lowing [Abbot] and [Apostol Calculus] (the latter for the Fundamental Theorem
of Calculus and Series).
(ii) This meant that I had to go slow (especially at the start, in the first three chapters)
which impacted me at the end as I was unable to do the Arzela-Ascoli theorem,
and was unable to really explore the idea of power series. In the future, it may be
prudent to go a little faster through the early material.
(iii) That said, I think the course was well-received and the grades indicate that most
students found the pace to be alright.
(iv) I also did not discuss the more general Riemann-Stieltjes integral because it did
not seem like the added generality was really necessary (plus, I think the main
ideas are already in the usual Riemann integral).
113
Bibliography
[Abbot] S. Abbott, Understanding Analysis, Springer (2001) (Available here)
[Apostol Calculus] T. Apostol, Calculus Vol 1 (2nd Ed), John Wiley and Sons (1967)
114