Typed Notes - Real Analysis
Typed Notes - Real Analysis
What is this course about? This is “calculus for thinking practitioners". We will lay the founda-
tion for “rules" that you are already very good at. So, we may work very hard to prove something
that you already know, but it is the process which is the focus of the course. You will enjoy it if you
are curious about “why" certain rules work, or “when do they fail", which is the kind of curiosity
that makes a mathematician. For others, think of this as a training in rigor that will prepare you
for the rest of this program.
1.1. The ZFC & Peano axioms. To do analysis on real and complex numbers, we must talk about
what those really are. And rather than take them completely for granted, we must understand
where they come from. For that, we must understand rational numbers, which come from in-
tegers, which themselves come from natural numbers. Is it possible to synthesize a basic set of
rules that eventually yield all the properties of natural numbers that we have been happily using?
Peano (1858-1932) tried to do this:
Definition 1.1 (Peano’s axioms). A set A, is called a Peano set if it satisfies the following axioms or
“rules".
(P1) A contains a distinguished element, which is called 0. In particular A is non-empty.
(P2) There exists a function S from A into itself.
(P3) S(a) ̸= 0 for any a ∈ A.
(P4) S is injective, i.e., if S(a) = S(b) for any a, b ∈ A, then a = b.
(P5) (Principle of Mathematical Induction) if B is a subset of A such that 0 ∈ B , and S(a) ∈ B
whenever a ∈ B , then B = A.
The function S is called a successor function on A.
Class discussion. Explain why the following choices of A and S do not make A a Peano set. You
must explicitly mention which of the above axioms are failing.
a. A = {0, 1}, S(0) = 1 and S(1) = 1
b. A = {0, 0.5, 1, 1.5, 2, 2.5, ...} and S(x) = x + 1.
1
2
It is “clear" to us that the “usual" natural numbers {0, 1, 2, ...} endowed with S(n) = n +1 is a Peano
set. One could take this for granted and proceed ahead (as we have done for many years in
school). If you take only the existence of a Peano set for granted, you may argue as follows: why
don’t we define
1 := S(0);
2 := S(S(0))
and so on. Then using the Principle of Mathematical Induction, we show that the original Peano
set contains nothing extra, and we are almost there (barring the issue of defining addition). How-
ever, this approach forces use to understand what “so on" (definition by recursion) means, and
why the above procedure yields a subset and so on. In fact, we have to ascribe meaning to all
of the italicized words above. In short, we must describe the fundamental rules governing the
formation of sets and functions!
Definition 1.2. A set is a well-defined collection of (mathematical) objects that are called ele-
ments of the set. If A is a set and x is an object, then we use the notation
x∈A
x ∉ A.
Example 1. Going back to our known understanding of sets, let A = {1, 2, {3}}. Then, 2 ∈ A, {2} ⊆ A,
but {2} ∉ A. On the other hand, {3} ∈ A but 3 ∉ A, and therefore, {3} ̸⊆ A.
END OF LECTURE 1
The ZFC Axioms - “Naive" approach. Note that not every collection of mathematical objects is
considered a set (read about Russell’s Paradox)! The following rules determine what is and isn’t
allowed to be called a set. (There are some redundancies in this list of axioms.)
A. Basic. Every mathematical object is a set. (This means that if you know that something is
an element of a set, then that element itself is considered to be a set.)
B. Extension. Two sets, A and B , are said to be equal, i.e., A = B , if they have the same
elements. Practically, to show A = B , one often shows that A ⊆ B and B ⊆ A.
3
A = {0, 1},
B = the set of possible remainders obtained when
dividing a perfect square by 4.
Let a ∈ A, we will show that there is a natural number c such that c 2 ≡ a (mod 4). For
a = 0, take c = 2. For a = 1, take c = 1. Thus, A ⊆ B . On the other hand, let b ∈ B . Then,
b = c 2 − 4q for some natural numbers c and q, and 0 ≤ b ≤ 3. If c is even, then 4 divides
b, which can only happen if b = 0. If c is odd, i.e., c = 2k + 1, then b = (2k + 1)2 − 4q =
4(k 2 +k − q)+1. The only number in {0, 1, 2, 3} of the form 4· +1 is 1. Thus, we have shown
that B ⊆ A.
C. Existence. There exists a set with no elements, called the empty set, denoted by ;.
D. Specification. Let A be a set and P (a) be a “property" that applies to every a ∈ A, i.e., P (a)
is either true or false, then
B = {b ∈ A : P (b) is true}
E. Pairing. Given two sets A, B , there exists a set containing exactly A and B as its elements,
which is denoted by {A, B }. Note that this also gives the existence of {A}.
F. Union. For a set F of sets, there exists a set, called the union of the sets in F , whose
elements are precisely the elements of the elements of F . I.e.,
Consequence. 1. For a non-empty set F of sets, there exists a set, called the intersection
of the sets in F such that
A ⇐⇒ x is an element of everyA ∈ F .
\
x∈
A∈F
G. Power Set. Given a set A, there exists a set P (A) whose elements are precisely the subsets
of A. This set is called the power set of A.
Remark. The axioms so far give the existence of unions, intersections (why?), set differ-
ences, and (finite) Cartesian products (this is more complicated) as sets. In particular,
given two sets A, B , the set A ×B is the set of all ordered pairs (a, b), where a ∈ A and b ∈ B .
4
END OF LECTURE 2
Definition 1.3. A relation from A to B is a subset R of A × B . We say that aRb if and only
if (a, b) ∈ R. The domain of R is
The range of R is
H. Replacement. Skip!
I. Regularity. Skip!
J. Choice. Skip!
Definition 1.4. Given a set A, the successor of A is the set A + = A ∪ {A}. A set which
contains the empty set as an element and the successors of each of its elements is called
an inductive set.
Proof. Homework! □
Theorem 1.6. There exists a unique minimal inductive set. That is, there is an inductive set ω such
that ω ⊆ S for every inductive set S, and if ω′ is an inductive set with the same property, then ω = ω′ .
Proof. Let A be an inductive set, whose existence is guaranteed by the Axiom of Infinity. By the
Axiom of Power Set, P (A) is a set. Then, by the Axiom of Specification,
ω=
\
W.
W ∈F
Step 1. ω is inductive. This is a direct consequence of Lemma 1.5.
5
Step 2. ω is minimal. Let S be an inductive set. Then, by Lemma 1.5, S ∩ A is an inductive set. By
the definition of F , S ∩ A ∈ F . By the definition of intersections, ω ⊆ S ∩ A ⊆ S. Thus, ω ⊆ S (do you
know what this property of ⊆ is called?).
Step 3. ω is unique. By minimality, ω ⊂ ω′ and ω′ ⊂ ω. By the Axiom of Extension, ω = ω′ . □
Theorem 1.7. Let ω be the minimal inductive set. Set 0 := ; and S(A) = A + . Then, ω is a Peano
set.
Theorem 1.8 (The recursion principle). Let X be a non-empty set, f : X → X be a function, and
a ∈ X . Then, there exists a unqiue function F : ω → X such that
(a) F (;) = a
(b) F (A + ) = f (F (A)) for each A ∈ ω.
Remember that the above theorem is claiming the existence of a particular subset of ω× X . When
we define sequences recursively, we are essentially applying the above theorem, as we will see
shortly.
1.2. The Natural numbers. Within the ZFC regime, ω is defined as the set of natural numbers.
But, how do we recover the familiar set of natural numbers from the above discussion? We iden-
tify N and ω as follows
0 := ;
1 := 0+ = {;} = {0}
2 := 1+ = {;, {;}} = {0, 1}
3 := 2+ = {;, {;}, {;, {;}}} = {0, 1, 2}
and so on.
END OF LECTURE 3
Definition 1.9 (Peano Addition). Given m ∈ N, the recursion principle gives the existence of a
unique function summ : N → N such that
(1) summ (0) = m;
(2) summ (n + ) = (summ (n))+ .
Define, for m, n ∈ N,
m + n := summ (n).
Proposition 1.10. 2 + 3 = 5.
Trying to prove 2 + 8 = 10 would be quite tedious, but not so much if we could say that 2 + 8 =
8 + 2. Let us recover the basic properties of addition that we have taken for granted.
Remark. Note that m + = (summ (0))+ = summ (0+ ) = m + 1 by this definition. So will now always
write m + 1 instead of m + .
Definition 1.11 (Peano Multiplication.). Given m ∈ N, the recursion principle gives the existence
of a unique function prodm : N → N such that
(1) prodm (0) = 0;
(2) prodm (n + ) = prodm (n) + m.
Define, for m, n ∈ N,
m · n := prodm (n).
Proof. We will only prove the additive part of (6). The rest are for self-study. Note that subtraction
is not available to us! We will fix an arbitrary m, n ∈ N and prove the statement by induction on k.
Consider the case k = 0. We are given that summ (0) = sumn (0), which by definition gives that
m = n. Next, assume the statement holds for k. We must prove it for k + 1. For this, suppose
m + (k + 1) = n + (k + 1). I.e., summ (k + 1) = sumn (k + 1). By definition, this gives that
But the successor map is injective. So, summ (k) = sumn (k). By the induction hypothesis, m = n.
By the principle of mathematical induction, the cancellation law holds for all k ∈ N. Since m, n ∈
N were arbitrary, the claim holds. □
1.3. Field axioms and ordered fields. The set of natural numbers are missing many nice opera-
tions such as subtraction and division, for instance, we cannot solve
x + 3 = 2,
3x = 5
within the set of natural numbers. In fact, we would like to work with number systems where we
can even solve x 2 + 2 = 0 and x 2 − 2 = 0! First, we introduce some abstract concepts that capture
some of these ideas.
x +0 = x
x · 1 = x.
8
x + y = 0.
x y = 1.
Remark. You should be tempted to denote the y in (F5) by −x and the y in (F6) by 1/x. How-
ever, this notation would be meaningless unless we established the uniqueness of y in both the
axioms. These are Theorems 1.1 and 1.7 in the book!
Remark. Define −x as the unique y given by Theorem 1.1. It is called the additive inverse of x.
We write a + (−b) as a − b for convenience.
Similarly, define 1/x as the unique multiplicative inverse of a nonzero x. We write a · (1/b) as
a/b for convenience.
Proof. By commutativity of · (F1), we already have the first equality. Now, by (F4), 1 + 0 = 1 and
x · 1 = x. Thus, by (F3)
x = x · 1 = x · (1 + 0) = (x · 1) + (x · 0) = x + (x · 0).
Definition 1.15. A set A with a relation < is called an ordered set if the following holds:
(O1) For every x, y ∈ A, exactly one of the following three holds: x = y, x < y or y < x.
(O2) Order is transitive. Given x, y, z ∈ F , if x < y and y < z, then x < z.
If x < y, we say that x is strictly less than y. The notation x ≤ y denotes the statement x < y or
x = y, and we say that x is less than or equal to y. Similarly, x > y and x ≥ y are interpreted.
Example 4. N is an ordered set if we define < as follows: m < n if there is a k ∈ N \ {0} such that
m + k = n.
Definition 1.16. An ordered field is a set F that admits two operations +, · and a relation < such
that (F, +, ·) is a field, (F, <) is an ordered set and the following conditions hold.
9
EXERCISE: Read Theorems 1.20-1.25 from Apostol. Our order axioms are DIFFERENT from Apos-
tol’s. His theorems 1.16-1.19 are our order axioms!!
Theorem 1.17 (Apostol, Theorem 1.21). In an ordered field, 0 < 1. We may use Theorem 1.12 from
Apostol: (−a) · b = −(a · b) and (−a) · (−b) = ab.
Proof. From the field axioms, we know that 0 ̸= 1. Thus, by Axiom (O1), either 0 < 1 or 1 < 0. If the
former holds, we are done.
If the latter holds, by (O3), 1 + (−1) < 0 + (−1). Thus, 0 < −1. By (O4), 0 < (−1) · (−1). Thus,
0 < 1, which is a contradiction.
Thus, 0 < 1 hold. □
END OF LECTURE 5
1.4. Upper bound and least upper bound. Let (F, +, ·, <) be an ordered field (you can think of the
usual real numbers, with the usual +, · and < as the key example).
Example 5. Let S = {x ∈ F : 0 ≤ x ≤ 1} and T = {x ∈ F : 0 ≤ x < 1}. Both S and T are bounded above.
In both cases, 1 is an upper bound. In the cases, of S, 1 is the maximum of S.
Remark. Upper bounds clearly need not be unique. However, if a bounded set has a maximum,
then the maximum must be unique (why?).
Definition 1.19. Let S ⊆ F be a bounded above set. An element b ∈ F is said to be a least upper
bound of S or supremum of S if
(i ) b is an upper bound of S,
(i i ) for any a ∈ F such that a < b, a is not an upper bound of S. In other words, given a < b,
there exists an s ∈ S such that a < s.
Theorem 1.20. Let S ⊂ F be a bounded set that admits a least upper bound. Suppose b 1 and b 2 are
least upper bounds of S. Then, b 1 = b 2
10
Proof. Note that there are three possiblities: either b 1 = b 2 , b 1 < b 2 or b 2 < b 1 . There is nothing to
prove in the first case.
Suppose b 2 < b 1 . Then, by item (i i ) of Definition 1.19 applied to b 1 , b 2 is not an upper bound
of S. However, by item (i ) of Definition 1.19 applied to b 2 , b 2 is an upper bound of S. This is a
contradiction.
Suppose b 1 < b 2 . The same argument holds with the roles of b 1 and b 2 exchanged. □
1.5. The set of real numbers. We assume the existence of a set R which admits operations +, ·
and a relation < so that (F1)-(F6) and (O1)-(O4) are satisfied and the set admits the least upper
bound property, i.e.,
(LUB) every bounded set S ⊆ R admits a least upper bound.
Some special subsets of R are:
(1) any x > 0 is called a positive number, and any x < 0 is called a negative number,
(2) N = {0, 1, 2, ...}, the set of natural numbers,
(3) P={1,2,...}, the set of positive natural numbers,
(4) Z = N
n ∪ {−n : n ∈ P},othe set of integers,
p
(5) Q = q
: q ∈ P, p ∈ Z , the set of rational numbers,
(6) Q = R \ Q, the set of irrational numbers.
′
Theorem 1.21 (Archimedean propery). Let x, y ∈ R such that x > 0. There exists a positive integer
n such that nx > y.
Clearly, S is non-empty. Suppose the claim was not true. Then nx ≤ y for all n ∈ P and y us ab
upper bound of S. Thus, by the l.u.b. property of R. b = sup S exists. Now, since b − x < b, by
item (i i ) of Definition 1.19, there exists an n ∈ P such that b − x < nx. Thus, b < (n + 1)x ∈ S. This
contradicts the fact that b is an upper bound of S. □
END OF LECTURE 6
12
Going forward, we will allow all the properties of real numbers that we have used so far, in-
cluding exponentiation, the absolute value function, etc.
2.1. Sequences.
Given a sequence, we would like to rigorously capture the idea that the terms of the sequence
are approaching a single real number.
Definition 2.2. A sequence {a n }n∈N in R is said to be convergent if there exists an L ∈ R such that
for every ε > 0. there exists an Nε,L ∈ N such that
a n → L as n → ∞.
Theorem 2.3. Let {a n }n∈N be a convergent sequence in R. Suppose ℓ1 and ℓ2 are limits of {a n }n∈N .
Then, ℓ1 = ℓ2 .
Proof. Let ε > 0. By definition, there exists an N1 ∈ N such that |a n − ℓ1 | < ε/2 for all n ≥ N1 .
Similarly, there exists an N2 ∈ N such that |a n − ℓ2 | < ε/2 for all n ≥ N2 . Let N = max N1 , N2 . Then,
But, ε > 0 was arbitrary. HW. If 0 ≤ x < ε for all ε > 0, then x = 0.
Note. This proof can also be done via contradiction. □
Example 7. (a) Let p > 0 be fixed. Let a n = 1/n p , n ∈ P. Then, {a n }n∈P is convergent and 0 is a
limit.
13
Proof. Let ε > 0. By the Archimedean property of R applied to x = ε1/p and y = 1, there is an
N = Nε ∈ P such that
N ε1/p > 1.
Thus, for n ≥ N , ¯ ¯
¯ 1 1 1
¯ n p − 0¯ = n p ≤ N p < ε.
¯
¯ ¯
END OF LECTURE 7
Example 8. (b) a n = (−1)n , n ∈ N. Then, {a n }n∈N is divergent.
Proof. Suppose {a n }n∈N is convergent and admits a limit L ∈ R. Let ε = 1. By definition, there
exists an N ∈ N such that
|a n − L| < 1 ∀n ≥ N .
Thus, by the triangle inequality,
|a 2N − a 2N +1 | ≤ |a 2N − L| + |a 2N +1 − L| < 1 + 1 = 2.
Definition 2.4. A sequence {a n }n∈N is said to be bounded if there exists an M > 0 such that
|a n | < M
for all n ∈ N.
Proof. (=>) Case 1. Let {a n } be an increasing and bounded sequence. Then, there exists an M > 0
such that
|a n | < M ∀n.
In other words −M < a n < M for all n. Let S = {a n : n ∈ N}. S is nonempty and bounded above. By
LUB, b = sup S exists in R.
Let ε > 0. By the definition of the supremum and monotonicity, ∃N ∈ N such that
b − ε < a N ≤ an ∀n ≥ N .
14
|a n − b| < ε.
Example 9. We show that limn→∞ n = +∞. Let R ∈ R. If R ≤ 0, then n ≥ R for all n ≥ 1. So we may
choose NR = 1. If R > 0, then by the Archimedean property, there is a positive natural number
N such that N > R. Choose NR = N in this case, we get that n > R for all n ≥ NR . Thus, we have
shown that for any R ∈ R, there is an NR ∈ N such that |a n | > R for all n ≥ NR .
Theorem 2.8 (Limit Laws). Let {a n } and {b n } be convergent sequences in R with limits a and b,
respectively.
(1) For any c ∈ R, {a n + c} converges to a + c and {c a n } converges to c a.
(2) The sequence {a n + b n } converges to a + b.
(3) The sequence {a n b n } converges to ab.
(4) Suppose b ̸= 0 and ∃M ∈ N such that b n ̸= 0 ∀n ≥ M then {1/b n }n≥M converges to 1/b.
(5) Suppose b ̸= 0 and ∃M ∈ N such that b n ̸= 0 ∀n ≥ M then {a n /b n }n≥M converges to a/b.
|b n − b| < ε1 = |b|/2.
15
Thus,
Remark. The converse statements generally do not hold in the above theorem, mostly because
the convergence of sequences such as {a n +b n }, {a n b n } and {a n /b n } may not in general imply the
convergence of the individual sequences {a n } and {b n }.
END OF LECTURE 8
Definition 2.9. A infinite series of real numbers is a formal expression of the form
∞
X
a0 + a1 + a2 + · · · or an .
n=0
P∞
Given a series n=0 a n , its sequence of partial sums (sops) is the sequence {s n }n∈N given by
sn = a0 + · · · + an , n ∈ N.
s1 = 1
1
s2 = 1 +
2
1 1 1 1 1 1
s4 = 1+ + + ≥ 1+ + +
2 3 4 2 4 4
1 1 1 1 1 1 1 1
s8 = 1+ + + + +···+ ≥ 1+ + +
2 3 4 5 8 2 2 2
..
.
1 1 1 1 1
s 2k ≥ 1+ + 2 + 4 + · · · + 2k−1 k = 1 + k .
2 2 8 2 2
For any m ∈ R, there is some k ∈ N such that s 2k > m. Thus, {s n } is divergent. □
P∞ 1
(b) Claim. n=1 n 2 is convergent.
Remark. We have seen an example of a telescoping series above. These are series of the form
P∞
c n − c n+1 for some sequence {c n }. Note that, in this case, s n = c 1 − c n+1 , so the convergence of
n=1
the series entirely depends on the convergence of {c n }.
P∞ n 1
(c) (Geometric) Claim. Let −1 < x < 1. Then, n=0 x converges and its sum is 1−x
. For |x| ≥ 1,
P∞ n 1
n=1 x = 1−x
diverges.
In the case, when |x| > 1, (1 + |x| − 1)n > n(|x| − 1). Thus, for any R ∈ R, by the Archimedean
principle, there exists an N ∈ N such that |x|N > R. Thus, {x n } is unbounded, and therefore,
divergent.
P∞ n
Returning to the series n=0 x . When |x| < 1, we use the limit laws of convergent sequences
to say that
1
. lim s n =
1−x n→∞
Proof. Let ℓ = a n . Let ε > 0. Then, there exists N ∈ N such that |s n − ℓ| < ε/2 for all n ≥ N .
P
END OF LECTURE 9
We will work with series with non-negative terms. Note that the sops for such series are
monotonically increasing, so the task of showing convergence reduces to showing the (upper)
boundedness of the sops.
Theorem 2.11 (Comparison Test). Suppose there exist constants M ∈ N and C > 0 such that
0 ≤ an ≤ C bn ∀n ≥ M .
P P P P
Then, a n if b n is convergent. In other words, b n diverges if a n diverges.
P P
Proof. Let {s n } and {t n } denote the sequence of partial sums of a n and b n respectively. Since
P
b n converges, there exists an N > M and an L > 0 such that t n ≤ L for all n ≥ N . Now,
sn ≤ C tn ∀n ≥ N .
s1 = 1
1 1 1
s3 = 1 + + ≤ 1 + 2
2p 3p 2p
1 1 1 1 1 1
s7 = 1+ p + p + p +···+ p ≤ 1+2 p +4 p
2 3 4 7 2 4
..
.
1 1 1
s 2k −1 ≤ 1 + 2 p
+ · · · + 2k−1 p(k−1) < .
2 2 1 − 2p−1
Exercise: complete the proof. □
Many other tests are derived from the comparison test. We will note two here, but prove only
one.
P
Theorem 2.12 (Ratio Test). Let a n be a series of non-negative terms. Suppose
a n+1
lim = L.
n→∞ a n
END OF LECTURE 10
Proof. Case 1. L < 1. Choose an r such that L < r < 1. Choosing ε = r − L >0, we obtain an N ∈ N
a n+1
such that < L + ε = r for all n ≥ N . Thus,
an
a N +1 < a N r
a N +2 < a N +1 r < a N r 2
a N +3 < a N +2 r < a N r 3
..
.
a N +k < aN r k .
In other words, for n ≥ N , a n ≤ ar NN r n = cr n . By the Comparison Test, and the convergence of the
geometric series r n , r < 1, we have the convergence of a n .
P P
Case 2. L > 1. Choose R such that 1 < R < L. Then, choosing ε = L − R, we have that for some
a n+1
N ∈ N, > L − ε = R > 1 for all n ≥ N . Thus, a n+1 > a n for all n ≥ N . The sequence {a n } cannot
an P
converge to 0. Thus, a n diverges.
19
P 1 n +1 P1
Case 3. L = 1. converges while limn→∞ = 1. On the other hand, diverges while
n2 n n
n2 + 1
limn→∞ = 1. □
n2
P
Theorem 2.13 (Root Test). Let a n be a series of non-negative terms. Suppose
p
n
lim a n = R.
n→∞
Proof. Exercise!
1/n 2 . How does one compute limn→ n 1/n without using
P P
For Case (3), we can take 1/n and
Fms? □
P P
Theorem 2.14 (Limit Laws for Series). Suppose a n and b n converge with sums a and b respec-
tively. Then, for constants ℓ and m, ℓa n + mb n converges to l a + mb. Suppose |a n | and |b n |
P P P
Proof. Exercies! □
b n = a n + |a n |.
P
Observe that 0 ≤ b n ≤ 2|a n |. Thus, by the comparison test, b n converges. Now, by the limit laws
P P
for convergent series, a n = (b n − |a n |) converges. □
P (−1)n
Example 12. Claim. n is convergent.
20
s 1 = −1
1 1
s 3 = −1 + − > s1
2 3
1 1 1 1
s5 = −1 + − + − > s 3
2 3 4 5
..
.
µ ¶ µ ¶ µ ¶
1 1 1 1
s 2k+1 = −1 + + − + + · · · − < 0.
2 3 4 2k + 1
Thus, {s 2k+1 } being a bounded increasing sequence, converges to some limit, say ℓ.
1
s2 = −
2
1 1 1
s 4 = −1 + − + < s 2
2 3 4
..
.
µ ¶ µ ¶ µ ¶
1 1 1 1 1
s 2k = −1 + − + − + · · · + ≥ −1.
2 3 4 5 2k
Thus, {s 2k } being a bounded decreasing sequence, converges to some limit, say m. Moreover,
1
s 2k+1 = s 2k + 2k+1 . So, by limit laws for sequences, ℓ = m. Exercise: why does this suffice to claim
that {s n } converges? □
Remark. The estimate in AST allows us to estimate sums of alternating series within any pre-
(−1)n
scribed error. For instance, to know ∞
P
n=1 n up to an error of 0.01. I need to find n so that
1
|S − s n | < .
100
Take n = 99, or the sum of the first 99 terms.
END OF LECTURE 11
21
Definition 3.1. Given a real number p and an ε > 0, the ε-neighborhood of p is the open interval
Definition 3.2. (a) Given a function f that is defined on some I = (a, p)∪(p, b), a < b, we say that
f has a limit L as x approaches p if:
for every ε > 0, there is a δ > 0 such that whenever 0 < |x − p| < δ, we have that | f (x) − L| < ε. OR
for every ε > 0, there is a δ > 0 such that whenever x ∈ Nδ (p) \ {p}, f (x) ∈ Nε (L).
This is denoted by
lim f (x) = L.
x→p
——————–Skipped one-sided limits!!——————
Given a function f : (a, p) → R we say that f has a left-hand limit L as x approaches p if:
for every ε > 0, there is a δ > 0 such that whenever p − δ < x < p, we have that | f (x) − L| < ε.
This is denoted by
lim f (x) = L.
x→p −
Given a function f : (p, b) → R we say that f has a right-hand limit L as x approaches p if:
for every ε > 0, there is a δ > 0 such that whenever p < x < p + δ, we have that | f (x) − L| < ε.
This is denoted by
lim f (x) = L.
x→p +
——————————————————————
Example 13. (1) (Constant functions) f (x) = c, for some fixed c ∈ R. Let p ∈ R. Let ε > 0. Let δ = 1
(or any ohter positive number!). Then, whenever 0 < |x − p| < 1, we have that | f (x) − c| = 0 < ε.
Thus, for every p ∈ R,
lim f (x) = c.
x→p
(2) (Identity function) f (x) = x. Fix p ∈ R. Let ε > 0. Choose δ = ε. Then, whenever |x − p| < δ,
we have that | f (x) − f (p)| = |x = p| < δ = ε. Since, ε > 0 and p ∈ R were arbitrary, we have that
limx→p x = p for all p ∈ R.
22
p p
(3) f (x) = x, x > 0. We will show that limx→p f (x) = p, p > 0. Let ε > 0.
p p
Scrapwork. We want to produce a δ > 0 such that if 0 < |x − p| < δ, then | x − p| < ε. Note
that
p p |x − p| 1
| x − p| = p p ≤ p |x − p|.
| x + p| p
END OF LECTURE 12
p
Example 14. Let δ = pε > 0. Then, for 0 < |x − p| < δ,
p p |x − p| 1
(3.1) | x − p| = p p ≤ p |x − p| < ε.
x+ p p
Since ε > 0 was arbitrary, we are done.
(4) f (x) = x1 , x ̸= 0. Then, limx→0 f (x) does not exist. Suppose it did, and was L. Then, for any
ε > 0, there exists a δ > 0 such that whenever 0 < |x| < δ, |1/x − L| < ε. In particular, for any
nonzero x, y ∈ (−δ, δ), we must have that
Let ε = 1/2 and δ correspond to this ε. Then, by the Archimedean property of R, there is an
1
N ∈ N such that N δ > 1, or N
< δ. Choose, x = N and y = N + 1. Then. 0 < |x|, |y| < δ, but
|1/x − 1/y| = |N + 1 − N | = 1 > 1/2.
Theorem 3.4 (Limit laws for functions). Let f and g be functions such that limx→p f (x) = A and
limx→p g (x) = B . Then,
(1) limx→p ( f (x) ± g (x)) = A ± B .
(2) limx→p ( f (x)g (x)) = AB .
(3) limx→p ( f (x)/g (x)) = A/B , if B ̸= 0.
| f (x)g (x) − AB | = | f (x)g (x) − f (x)B + f (x)B − AB | ≤ | f (x)||g (x) − B | + |B || f (x) − A|.
ε
Let ε1 = . Let δ1 > 0 be such that whenever 0 < |x − p| < δ1 , then | f (x) − A| < ε1 . In
(2|B | + 1)
particular,
| f (x)| < |A| + ε1 = M .
ε
Let ε2 = . Let δ2 > 0 be such that whenever 0 < |x − p| < δ2 , then |g (x) − A| < ε2 .
2M
23
END OF LECTURE 13
3.2. Continuity.
Definition 3.5. Let S ⊂ R, $ f : S → R and p ∈ S. We say that f is continuous at p if: for every ε > 0,
there is a δ > 0 such that whenever |x − p| < δ and x ∈ S, we have that | f (x) − f (p)| < ε, in other
words
lim f (x) = f (p).
x→p
Theorem 3.6 (Algebraic combinations of continuous functions). Let f , g be functions that are
continuous at p. Then, f ± g , f g and f /g (when g (p) ̸= 0) are continuous at p.
Example 15. (a) Based on our previous computations, constant functions and the identity func-
tion ( f (x) = x) are continuous on all of R. Using the above theorem, every polynomial, i.e., func-
tion of the form
p(x) = a n x n + · · · a 0 ,
where a 0 , ..., a n are constants, is continuous. Every rational function, i.e., function of the form
p(x)
r (x) = ,
q(x)
where p and q are polynomials, is continuous on the set {x ∈ R : p(x) ̸= 0 and q(x) ̸= 0}. Note.
r (x) may be extended to a continuous function on certain points x, where p(x) = q(x) = 0, but a
priori, the function r (x) itself is considered undefined on such points! E.g., the function
x2 − 1
r (x) =
x −1
is only defined and continuous on R \ {1}. The function
x 2 −1 , x ̸= 1,
R(x) = x−1
2, x = 1.
is defined on all of R, and is in fact also continuous everywhere (but this requires proof at x = 1).
24
(b) f (x) = x r for any r ≥ 0 is continuous on (0, ∞). In HW 5, you will show that limx→p x n = p n for
every n ∈ N and p ∈ R. Now, let q > 0. There exists a natural number n such that q ≤ n. Thus,
0 ≤ y q − 1 ≤ y n − 1, y >1
0 ≤ 1 − yq ≤ 1 − yn, y ≤ 1.
By the squeezing principle (HW05), limx→1 x q = 1. Now, for any a > 0, note that
|x q − a q | = a q ¯(x/a)q − 1¯ .
¯ ¯
(c) Trigonometric functions will have the meaning (in terms of triangles) that they did in school
(angles are measured in radians!). You may assume all the trigonometric formulas from school.
You will establish their continuity in HW5.
END OF LECTURE 14
Theorem 3.7 (Compositions). Given f : A → R and g : B → R, assume that ran( f ) ⊂ B . Then,
g ◦ f : A → R given by g ◦ f (x) = g ( f (x)) is well-defined. If f is continuous at p ∈ A, and g is
continuous at q = f (p), then g ◦ f is continuous at p.
Proof. Let ε > 0. We must produce a δ > 0 so that whenever |x − p| < δ and x ∈ A, we have that
By the continuity of g at q = f (p), there is a τ such that whenever |y − q| < τ and y ∈ B , then
Now, set ε2 = τ. By the continuity of f at p, there is a δ > 0 such that whenever |x − p| < δ and
x ∈ A, we have that
| f (x) − f (p)| < τ.
Thus, whenever |x − p| < δ and x ∈ A, we have that | f (x) − q| < τ and f (x) ∈ B . Thus, by (3.2),
Before we proceed to the next theorem about continuous functions, let us prove a result about
sequences.
Lemma 3.8 (Comparison for sequences). Suppose {a n }, {b n } ⊂ R are convergent sequences such
that a n ≤ b n , then L 1 = limn→∞ a n ≤ limn→∞ b n = L 2 .
25
Proof. Suppose not. Say that L 1 > L 2 . Let ε = (L 1 − L 2 )/2. Then, there exists an N1 ∈ N and N2 ∈ N
such that L 1 − ε < a n and b n < L 2 + ε for all n ≥ max{N1 , N2 }. I.e., b n < (L 1 + L 2 )/2 < a n for all
n ≥ max{N1 , N2 }. This is a contradiction. □
A common test used to detect zeros of a polynomial are to look for places where the polyno-
mial “changes sign". This comes from an important property of continuous functions!
Theorem 3.9 (The Intermediate Value Theorem). Let f : [a, b] → R be a continuous function.
Suppose y is a real number lying between f (a) and f (b). Then, there is a c ∈ [a, b] such that f (c) =
y.
We will assume that f (a) ≤ y ≤ f (b). The case f (b) ≤ y ≤ f (a) can be handled similarly.
Proof. f y = f (a) or y = f (b), then c = a or c = b, respectively, works. So, we may assume that
f (a) < y < f (b). Let
S = {x ∈ [a, b] : f (x) < y} ⊂ [a, b].
c := sup S exists.
We first show that f (c) ≤ y. Let n ∈ P. Since c − 1/n is not an upper bound of S, there is an x n ∈ S
such that
1
c−
< x n ≤ c.
n
Thus, by the above lemma, limn→∞ x n = c. By the above lemma and sequential characterization
of continuity, y ≥ limn→∞ f (x n ) = f (c).
Next, we show that y ≤ f (c). Since b is an upper bound of S, c ≤ b. Moreover, f (c) ≤ y < f (b),
so c ̸= b. Let N ∈ N such that c + 1/n < b for all n ≥ N . Then, f (c + 1/n) ≥ y. By the sequential
char. of cont. and the above lemma, f (c) = limn→ f (c + 1/n) ≥ y. □
Corollary 3.10 (Bolzano’s theorem). Let f : [a, b] → R be a continuous function such that f (a)
and f (b) assume opposite signs (so, are non-zero). Then, there is at least one c ∈ (a, b) such that
f (c) = 0.
Then, g (0) = f (0) − f (π) = −( f (π) − f (2π)) = −g (π). Either g (0) = 0, or g takes opposite signs at
0 and π. In either case: f takes the same values on a pair of anti-podal points. This is the 1-
dimensional case of the Borsuk–Ulam theorem. This is the basis of statements such as, "At any
given time, there are two polar opposite points on the Earth’s equator where the temperature is
exactly the same!"
END OF LECTURE 15
————Skip!————–
Theorem 3.11 (Existence of nth roots). Given a positive integer n and a positive number a, there
is exactly one positive nth root of a.
Proof. Let c > 1 such that a ∈ [0, c]. Let f (x) = x n . Then, f (0) = 0 and f (c) = c n > c > a > 0. By IVT,
there is some b ∈ (0, c) such that b n = a. Now, since 0 < x < y ⇒ x n < y n , we are done. □
————Skip!————–
f (x) ≥ L ∀x ∈ S.
| f (x)| ≤ M ∀x ∈ S.
Theorem 3.13. A function f : S → R is bounded if and only if it is bounded above and bounded
below on S.
Example 16. The function 1/x, x ̸= 0, is neither bounded above nor bounded below on its do-
main. However, it is bounded below on (0, 1).
Theorem 3.14. Let f : [a, b] → R be a continuous function. Then f is bounded on [a, b].
Proof. Given any closed interval I = [c, d ] with midpoint e, if f is bounded on I − = [c, e] and I + =
[e, d ], then f is bounded on [c, d ]. Suppose f is not bounded on [a, b]. Let I 0 = [a, b] and a 0 = a.
Then, by the contra-positive of the above statement, f is either not bounded on I 0− = [a, (a +b)/2]
or on I 0+ = [(a + b)/2, b]. Pick the one on which it is not bounded (pick left to break a tie), and call
it I 1 . Call its left endpoint a 1 . Continuing this way, inductively, we obtain a sequence of intervals
27
+ −
I 0 = [a 0 , b 0 ],...,I k = [a k , b k ],..., so that f is unbounded on I k and I k is either I k−1 or I k−1 . Since we
are halving the intervals at each stage, |b j − a j | < (b − a)/2 j .
Let s = sup{a j : j ∈ N}. Justify why s exists, and show that s ∈ [a, b].
By continuity, there is a δ > 0 such that if x ∈ Nδ (s) ∩ [a, b], then | f (x) − f (s)| < 1. I.e., | f (x)| <
1 + | f (s)| on x ∈ Nδ (s) ∩ [a, b]. Let N ∈ N such that for all n ≥ N , a n ∈ Nδ/2 (s) ∩ [a, b]. This works
because of the monotonicity of {a n }. Choose N ′ ≥ N such that for all n ≥ N ′ , (b − a)/2n < δ/2.
Thus,
|b n − s| < |b n − a n | + |a n − s| < δ ∀n ≥ N ′ .
This contradicts the fact that f is unbounded on these intervals! □
END OF LECTURE 16
Definition 3.15. A function f : S → R is said to have a global maximum (minimum) at p ∈ S if
f (p) ≥ (≤) f (x) for all x ∈ S.
Theorem 3.16 (Extreme value theorem). Let f : [a, b] → R be a continuous function. Then, f
attains a global maximum and a global minimum in [a, b].
Proof. Let s = sup{ f (x) : x ∈ [a, b]}. We show that there is some c ∈ [a, b] such that f (c) = s.
Suppose not. Let g (x) = s − f (x). Then, g is continuous and g (x) > 0 on [a, b]. Thus, 1/g (x) is
bounded on [a, b]. Thus, by the above theorem, there is an M > 0 such that 1/g (x) ≤ M for all
x ∈ [a, b]. Thus, s − f (x) ≥ 1/M for all x ∈ [a, b]. Thus, s − (1/M ) ≥ f (x) on [a, b]. This contradicts
the fact that s is the least upper bound.
The first part of the argument applied to − f on [a, b] shows that there exists a d ∈ [a, b] such
that − f (d ) ≥ − f (x) for all x ∈ [a, b]. Thus, f (x) ≤ f (d ) on [a, b]. □
Corollary 3.17. Let f : [a, b] → R be a continuous function. Then, f ([a, b]) = [min[a,b] f (x), max[a,b] f (x)].
Theorem 3.18 (Composition for limits, based on HW05). Let f and g be functions such that
Moreover, suppose that for some δ > 0, if 0 < |x − p| < δ, then | f (x) − L| > 0. Then,
4. D IFFERENTIATION
Fun fact: continuous functions can be very weird. They can have “corners" or “be jagged" at
every point!
Definition 4.1. Given a function f : (a, b) → R and p ∈ (a, b) we say that f is differentiable at p if
the limit
f (p + h) − f (p) f (x) − f (p)
lim = lim
h→0 h x→p x −p
′ ′
exists, in which case, we call the above limit f (p). f (p) is also sometimes known as the rate of
change of f at p. If f is differentiable at each p ∈ (a, b), it is said to be differentiable on (a, b), and
f ′ : (a, b) → R is the derivative function of f .
Example 17. (a) (Constant functions) f (x) = c, c ∈ R, is differentiable on R and f ′ (x) ≡ 0. For any
x ∈ R and h ̸= 0,
f (x + h) − f (x)
= 0.
h
(b) (Linear functions) f (x) = ax +b, x, a, b ∈ R, is differentiable at each x ∈ R, and f ′ (x) = a. Proof:
homework!
(c) (Natural powers) f (x) = x n , n ∈ P. We will use the formula (telescoping sum)
n−1
a n − b n = (a − b) a k b n−1−k .
X
k=0
Note that, for any x ∈ R and n ∈ P,
(x + h)n − x n n−1
(x + h)k x n−1−k .
X
=
h k=0
Theorem 4.2. Let f : (a, b) → R be a function that is differentiable at p ∈ (a, b). Then, f is contin-
uous at p.
lim
x→ f (x)= f (p).
□
29
Remark. The converse need not be true. Consider f (x) = |x|. The function is continuous (why?),
but observe that
f (h) − f (0) 1, h > 0,
=
h −1, h < 0.
Thus, taking {(−1)n n −1 }, we see that the derivative at 0 does not exist.
Theorem 4.3 (Algebra of derivatives). Let f and g be defined on (a, b) and differentiable at p ∈
(a, b). Then,
(1) so is f + g , and ( f + g )′ (p) = f ′ (p) + g ′ (p),
(2) (product rule) so is f g , and ( f g )′ (p) = f (p)g ′ (p) + f ′ (p)g (p),
g (p) f ′ (p)− f (p)g ′ (p)
(3) (quotient rule) if g (p) ̸= 0, so is f /g , and ( f /g )′ (p) = g 2 (p)
.
Proof. We only prove the quotient rule in the special case of f ≡ 1. Note that
1 1
g (p+h)
− g (p) g (p + h) − g (p) 1 1
=− .
h h g (p) g (p + h)
By the continuity of g at p and the algebra of limits of functions,
1 1
g (p+h)
− g (p) g (p + h) − g (p) 1 1 g ′ (p)
lim = − lim =− .
h→0 h h→0 h g (p) g (p + h) g (p)2
□
Example 18. We now have the differentiability of polynomials and rational functions in their
domains of definition.
END OF LECTURE 18
Definition 4.4. Let f : A → B be a one-to-one and onto function. f is said to be invertible on A,
and given, any y ∈ B , there is a unique x y ∈ A such that f (x y ) = y. Define the inverse of f function
f −1 : B → A as
f −1 (y) = x y .
Then, f −1 ( f (x)) = x for all x ∈ A, and f −1 ( f (y)) = y for all y ∈ B .
Now, let x, y ∈ [a, b] such that x < y. If x = a, we already have that f (x) < f (y) by (4.1). Now,
suppose a < x < y. Then, applying the above argument to f restricted to the domain [a, y], we
have that f (a) < f (x) < f (y). Thus, f is strictly increasing on [a, b]
Case 2. f (a) > f (b). Apply Case 1 to − f to obtain that f is strictly decreasing. □
END OF LECTURE 19
Proof. (i i ) We have already shown that f is strictly monotone. Assume, WLOG (why?), that f is
strictly increasing. Then, by IVT, we have that J = [ f (a), f (b)].
Case 1. p ∈ (a, b). Let q = f (p). Let d = 21 min{p − a, b − p} > 0. Let ε > 0. Suppose ε ≤ d . Then
Thus,
p − ε < f −1 (y) < p + ε.
If ε > d , choose δε as δd . Then, whenever |y − q| < δd , we have that | f −1 (y) − f −1 (q)| < d < ε.
Case 1. p = a or p = b. Homework!
31
f −1 (q + k) − f −1 (q) h(k) 1
= = f (p+h(k))− f (p)
.
k f (p + h(k)) − f (p)
h(k)
By continuity of f −1 , limk→0 h(k) = 0. Since h(k) ̸= 0, whenever k ̸= 0, we have by the composition
rule and the limit laws that
f −1 (y + k) − f −1 (y) 1 1 1
lim = lim f (x+h(k))− f (x)
= f (x+h)− f (x)
= ′ (x)
.
k→0 k k→0 limh→0 f
h(k) h
□
Example 19. The above theorem immediately gives the continuity and differentiability of f (x) =
1
x 1/n for x > 0 and n ∈ P. Moreover, f ′ (x) = nx (n−1)/n .
By the algebra of derivatives and induction (or the chain rule — which we haven’t done yet!),
¢p
one obtains the continuity and differentiability of f (x) = x 1/q , x > 0, for each p ∈ Z and q ∈ P.
¡
p
Moreover, f ′ (x) = q x (p−q)/q .
Example 20. We may now define inverse trigonometric functions. As convention, we invert
the sine function on [−π/2, π/2], the cosine function over [0, π] and the tangent function over
(−π/2, π/2).
END OF LECTURE 20
Theorem 4.6. Let f : (a, b) → R and g : (c, d ) → R such that f ((a, b)) ⊂ (c, d ). Say f is differentiable
at p and g is differentiable at q = f (p). Then, g ◦ f is differentiable at p, and
Define
g (q+ℓ)−g (q) , ℓ ̸= 0, q + ℓ ∈ (c, d )
ℓ
G(ℓ) =
g ′ (q), ℓ = 0.
Since limℓ→0 G(ℓ) = G(0), G is continuous at ℓ = 0. Note that
g (q + ℓ) − g (q) = ℓG(ℓ).
Thus,
h(p + k) − h(p) g (q + ℓ) − g (q) ℓ(k)G(ℓ(k))
= = .
k k k
Now, limk→0 ℓ(k)
k
= f ′ (p), by assumption, and limk→0 G(ℓ(k)) = G(0) = g ′ (q) by the continuity of ℓ
and G at 0. Thus, we are done. □
Definition 4.7. Let f : A → R. We say that f attains a local maximum (minimum) at a ∈ A if there
exists a δa > 0 such that f (a) ≥ f (x) ( f (a) ≤ f (x)) for all x ∈ A ∩ Nδa (a).
Theorem 4.8 (Extreme=>Critical). Let f : (a, b) → R. Let c ∈ (a, b) such that f is differentiable at c
and f attains a local extremum at c. Then, f ′ (c) = 0.
END OF LECTURE 21
Proof. We assume (WLOG) that f attains a local maximum at c. Then, there is a δ > 0 such that
f (c) ≥ f (x) for all x ∈ Nδ (c) ∩ (a, b). We may further assume that Nδ (c) ⊂ (a, b).
By the sequential characterization of limits, we have that for any sequence {c n } ⊂ (a, b) such
that limn→∞ c n = c, we have that
f (c n ) − f (c)
lim = f ′ (c).
n→∞ cn − c
33
The theorem is most helpful in the following way: to identify potential points of local extrema
within open intervals where f is differentiable. Thereafter, one needs to do some local analysis
at the potential points. For the local analysis, one inspects the sign of the derivative ‘just a bit
before’ and ‘just a bit after’ the point.
Theorem 4.9 (Mean Value Theorem). Let f : [a, b] → R be continuous on [a, b] and differentiable
f (b)− f (a)
on (a, b). Then, there is some c ∈ (a, b) such that f ′ (c) = b−a
.
Proof. Special case. f (a) = f (b). This is referred to as Rolle’s theorem. By the extreme value
theorem, there exist c 1 , c 2 ∈ [a, b] where f attains its global extrema on [a, b]. If at least one of c 1
or c 2 are in (a, b), we are done. If f (a) = f (b) = c 1 = c 2 . Then, f is a constant function, in which
case, f ′ ≡ 0 = on [a, b].
END OF LECTURE 22
f (b)− f (a)
General case. We want to apply Rolle’s to a function G such that G ′ (c) = 0 ⇐⇒ f ′ (c)− b−a
= 0.
f (b)− f (a)
Let g (x) = f (x) − x b−a . Then, g is cont. on [a, b] and differentiable on (a, b). Moreover,
g (a) = b f (a) − a f (b) = g (b). Thus, by Rolle’s theorem, there is some c ∈ (a, b) such that g ′ (c) = 0,
i.e.
f (b) − f (a)
f ′ (c) = .
b−a
34
□
Applications of MVT
Theorem 4.10. Let f : [a, b] → R be continuous on [a, b] and differentiable on (a, b). Then
(1) if f ′ (x) > 0 for all x ∈ (a, b), f is strictly increasing on [a, b],
(2) if f ′ (x) < 0 for all x ∈ (a, b), f is strictly decreasing on [a, b],
(3) if f ′ (x) = 0 for all x ∈ (a, b), f is a constant function on [a, b].
Proof. We prove only (1) and leave the rest as an exercise. Let x, y ∈ [a, b] such that x < y. Apply
the MVT on f restricted to [x, y], we have the existence of some z ∈ (x, y) such that
f (y) − f (x)
0 < f ′ (z) = .
y −x
Thus, f (y) > f (x). □
Remarks. (1) What if f ′ (c) > 0 or f ′ (c) < 0 for some c ∈ (a, b)?
(2) What about the converse of each part above? Remember that x 3 is a strictly increasing func-
tion.
(3) You are allowed to use the correct variations of the above result.
Example 22. The above theorem gives the first derivative test whose formal statement is given
as Theorem 4.8 in Apostol. Let us return to f (x) = (x − 1)2 + |x| + x 3 /3. Note that
x 2 + 2x − 3 = (x − 1)(x + 3), x < 0,
′
f (x) = ³ p ´³ p ´
x − ( 2 − 1) x + 2 + 1) , x > 0.
Test at x = −3: f ′ (x) > 0 on (−4, −3) and f ′ (x) < 0 on (−3, 0), so f (x) ≤ f (−3) for all x(−4, 0). So,
−3 is a point of local max.
p
Test at x = 0: f ′ (x) < 0 on (−1, 0), and f ′ (x) < 0 on (0, 2 − 1). So, f (x) > f (0) > f (y) for every
p
x ∈ (−1, 0) and y ∈ (0, 2 − 1). Thus, x = 0 is not a point of local max/min.
p
Test at x = 2 − 1 Homework!
Definition 4.11. Let f : (a, b) → R be a differentiable function such that f ′ : (a, b) → R is also
differentiable. Then, the ( f ′ )′ is denoted by f ′′ and is called the second derivative of f . One may
defined the nth derivative of f inductively, and denote it by f (n) . If f (n) exists for all n ∈ N, we say
that f is infinitely differentiable on (a, b)
END OF LECTURE 23
Definition 4.12. If a function f has an nth derivative at x 0 , its nth Taylor polynomial at x 0 is the
polynomial
x f (n) (x 0 )
P n0 (x) = P n (x) = f (x 0 ) + f ′ (x 0 )(x − x 0 ) + f ′′ (x 0 )(x − c)2 + · · · + (x − x 0 )n .
n!
P nc can be used to approximate the function f . It would be useful to under stand what the
error term looks like. There are many known forms of the error term. We mention one.
Theorem 4.13 (Taylor’s theorem). Let f : (a, b) → R be (n + 1)-times differentiable on (a, b). Let
x 0 ∈ (a, b). Then, for any x ∈ (a, b), there exists a c(x) between x and x 0 such that
x f (n+1) (c(x))
f (x) = P n0 (x) + (x − x 0 )n+1 .
(n + 1)!
Idea: Want to apply Rolle’s theorem to a function G such that
x
′ (n+1) f (x) − P n0 (x)
G (c) = 0 ⇐⇒ f (c) = (n + 1)! .
(x − x 0 )n+1
Could try
x
(n) f (x) − P n0 (x)
f (t ) − t (n + 1)! .
(x − x 0 )n+1
But G(y) ̸= G(c). We will skip the proof, but work out an example.
Question from class. What if we wish to approximate f by P n+1 instead of P n ?
Answer. The best we can say is that there is a function h = h x0 : (a, b) → R such that lim h(x) = 0
x→x 0
and
x
(4.2) 0
f (x) = P n+1 (x) + h(x)(x − x 0 )n+1
for x ∈ (a, b). Thus, we don’t get as much information on the form of the error term as we do in
Taylor’s theorem. Now, if f was (n + 2)-times differentiable, then by Taylor’s theorem, h in (4.2)
would become
f (n+2) (c(x))
h(x) = (x − x 0 )
(n + 2)!
Example 23. Let f (x) = cos(x). We compute the 3rd Taylor polynomials of f at c = 0. Note that
f (0) = 1, f ′ (0) = 0 and f ′′ (0) = −1. Then,
1 2
P 3 (x) = 1 − x .
2!
36
5. I NTEGRATION
Our goal will be to construct a “reasonable" theory of “area under a graph". There are two
equivalent theories of integration that appear in introductory calculus books — that of Riemann
integration and Darboux integration. We will follow Apostol and discuss Darboux integration.
The word “reasonable" above refers to the fact that we want area, which is a function whose
inputs are sets and outputs are real numbers, to satisfy certain rules (axioms) such as positivity,
finite additivity, monotonicity under contaiments, etc. It is possible that for any such theory,
the areas of certain sets just cannot be measured, i.e., the sets are not ‘measurable’! You will
encoutner all of this more formally in a measure theory course, but we will see some indication
of this in this course as well.
we will perform integration on a closed interval, i.e., a set of the form [a, b], a < b.
P = {x 0 , ..., x n }
such that a = x 0 < x 1 < · · · < x n−1 < x n = b. The partition P determines n subintervals, [x j −1 , x j ],
1 ≤ j ≤ n, of [a, b]. We refer to [x j −1 , x j ] as the j t h interval of P .
(ii) Given a partition P of [a, b], a refinement of P is a partition P ′ such that, as sets, P ⊆ P ′ ,
i.e., every subinterval determined by P ′ is a subset of some subinterval determined by P .
(iii) Given two partitions P = {x 0 < x 1 < · · · < x n } and Q = {y 0 < y 1 < · · · < y m } of [a, b], their
common refinement is the partition
Remark. It is a (tedious to prove) fact that for any ℓ ∈ {1, ..., N } there is a unique j (ℓ) ∈ {1, ..., n}
and unique k(ℓ) ∈ {1, ..., m} such that
Definition 5.2. A function f : [a, b] → R is called a step function if there is a partition P = {x 0 <
· · · < x n } such that s is constant on each open subinterval of P , i.e., for each 1 ≤ k ≤ n, there is a
real number s k such that
s(x) = s k for x ∈ (x k−1 , x k ).
At the end points of these intervals, the function may be defined as anything.
Example 24. The floor and ceiling functions restricted to closed and bounded intervals. Inifnitely
many steps are not allowed in the above definition.
38
Remark. The sum, difference, product and quotient (when defined) of two step functions, with
partitions P and Q is also a step function with partition P ∪ Q. We demonstrate this with a few
pictures, since the proof is long and boring.
Definition 5.3. Let s : [a, b] → R be a step function with partition P = {x 0 < ... < x n }. Its integral is
Rb
denoted by a s(x)d x and is given by
n
X
s k (x k − x k−1 ).
k=1
Ra Rb
We also define b s(x)d x to be − a s(x)d x.
END OF LECTURE 26
Remark. The partition of a step funciton is not uniquely determined. A refinement of P would
also work, but it would leave the integral unaffected.
Theorems 1.2-1.8 state certain key properties of the above definitions such as
Z b Z b Z b
(c 1 s(x) + c 2 t (x))d x = c 1 s(x)d x + c 2 t (x)d x,
a a a
Z b Z b
s≤t ⇒ s(x)d x ≤ t (x)d x,
a a
and
Z kb Z b
s(x/k)d x = k s(x)d x.
ka a
These can formally proved using the hints listed in 1.15. See also Theorem 11.2.16 in Tao.
To integrate a general function, f : [a, b] → R, we want to approximate it with step functions
(from above and below), i.e., we want to consider two classes of functions
Definition 5.4.
For an unbounded function, such as 1/x on [0, 1], the latter set is empty! So, we stick to
bounded functions.
Lemma 5.5. Let f : [a, b] → R be a bounded function, i.e., there is an M > 0 such that
Proof. Let g (x) = −M , x ∈ [a, b], and h(x) = M , x ∈ [a, b]. Then, g ∈ S f and h ∈ T f . Moreover, g ≤ t
for all t ∈ T f and h ≥ s for all s ∈ S f . Thus,
Z b Z b
−M (b − a) = g (x)d x ≤ t (x)d x, ∀t ∈ T f
a a
Z b Z b
M (b − a) = h(x)d x ≥ s(x)d x, ∀s ∈ S f .
a a
□
END OF LECTURE 27
Definition 5.6. Let f : [a, b] → R be a bounded function. The lower integral of f is the quantity
½Z b ¾
I ( f ) = sup s(x)d x : s ∈ S f ,
a
and the upper integral of f is the quantity
½Z b ¾
I ( f ) = inf t (x)d x : t ∈ T f .
a
We say that f is Riemann integrable if I ( f ) = I ( f ), in which case this quantity is called the integral
Rb
of f over [a, b] and denoted by a f (x)d x. We set
Z a Z b
f (x)d x = − f (x)d x,
b a
Z a
f (x)d x = 0.
a
Example 25. Show that every step function is integrable and there is no ambiguity in using the
Rb
symbol a s(x)d x for a step function.
s n (x) = f (x j −1 ), x ∈ [x j −1 , x j ),
t n (x) = f (x j ), x ∈ [x j −1 , x j ).
40
END OF LECTURE 28
We will take the following properties of Riemann integration for granted: linearity with re-
spect to integrand (Thm. 1.16); additivity with respect to interval of integration (Thm 1.17); in-
variance under translation (Thm 1.18); epxansion/contraction of interval of integration (Thm
1.19); comparison (Thm 1.20). Try to prove these for extra practice, or see Section 1.27 for the
proofs.
41
Definition 5.8. A function f : A → R is s.t.b. uniformly continuous on A if, for every ε > 0, there
is a δ > 0 such that whenever |x − y| < δ, x, y ∈ A, we have that | f (x) − f (y)| < ε.
Example 27. (1) f (x) = x is uniformly continuous on R. Note that δ = ε works for each ε > 0.
(2) f (x) = x 2 is not uniformly continuous on R. We need to produce an ε > 0 such that for every
δ, there exist x δ and y δ such that |x δ − y δ | < δ, but | f (x δ ) − f (y δ )| ≥ ε.
Choose ε = 1. Let δ > 0. Let x δ = δ1 and y δ = δ1 + δ2 . Then, |x δ − y δ | < δ, but
| f (x δ ) − f (y δ )| = 1 + δ2 /4 > 1.
FACT. Every continuous function on a closed and bounded interval is, in fact, uniformly contin-
uous! We will not prove this in class, however, this can be proved using a similar technique as was
used for showing that continuous functions on closed and bounded intervals are bounded.
mj ≤ f ≤ Mj.
s n (x) = m j , x ∈ [x j −1 , x j )
s n (b) = f (b).
and
t n (x) = M j , x ∈ [x j −1 , x j )
t n (b) = f (b).
42
Then,
Z b Z b
s(x)d x ≤ I ( f ) ≤ I ( f ) ≤ t (x)d x.
a a
Thus,
b−a Xn
0 ≤ I(f )− I(f ) ≤ M j − m j < ε(b − a).
n j =1
Since ε > 0 was arbitrary, we are done! □
END OF LECTURE 29
Theorem 5.10 (M integrals). Let f : [a, b] → R be continuous. Then, there is some c ∈ [a, b] such
that Z b
f (c)(b − a) = f (x)d x.
a
Theorem 5.11. Let a < b. Let f : [a, b] → R be Riemann integrable. Then, so is | f |, and
¯Z b ¯ ¯Z a ¯ Z b
¯ ¯ ¯ ¯
¯
¯ f (x)d x ¯¯ = ¯¯ f (x)d x ¯¯ ≤ | f (x)|d x.
a b a
Theorem 5.12 (The first Fundamental Theorem of Calculus). Let f : [a, b] → R be a Riemann
integrable function. Let F : [a, b] → R be the function given by
Z x
F (x) = f (t )d t .
a
Then, F is a continuous function on [a, b]. Moreover, if f is continuous at p ∈ (a, b), then F is
differentiable at p, and F ′ (p) = f (p).
Proof. Since f is bounded, there exists an M > 0 such that −M ≤ f (x) ≤ M for all x ∈ [a, b]. Now,
let x, y ∈ [a, b] such that x < y. Then,
Z y
F (y) − F (x) = f (t )d t ≤ M (y − x).
x
Similarly, −M (y − x) ≤ F (y) − F (x). Thus,
Interchanging the role of x and y, we get the same conclusion if y < x. You will show in the next
assignment that the above condition implies continuity of F .
Now, we assume that f is continuous at some p ∈ (a, b). Let h ∈ R such that p + h ∈ (a, b).
Then,
F (p + h) − F (p) 1
Z p+h
= f (x)d x
h h p
1
Z p+h
= f (p) + f (x) − f (p)d x
h p
1
Z p+h
= f (p) + ( f (x) − f (p))d x
h p
= f (p) +G(h).
We wish to show that limh→0 G(h) = 0, in other words, given ε > 0, there is a δ > 0 such that
whenever 0 < |h| < δ, we have that |G(h)| < ε.
Let ε > 0. By the continuity of f at p, there exists a δ > 0 such that whenever x ∈ (a, b) and
|x − p| < δ, we have that | f (x) − f (p)| < ε. Thus, using the previous theorem
¯R ¯
¯ Z p+h
¯1
¯ 1 ¯¯ p+h f (x) − f (p)d x ¯¯ , h > 0,
¯ h ¯p
¯ ( f (x) − f (p))d x ¯¯ = ¯
¯h
p − 1 ¯¯R p+h f (x) − f (p)d x ¯¯ , h < 0,
h p
R
1 p+h ¯ f (x) − f (p)¯ d x, h > 0,
¯ ¯
h p
≤ < ε.
− 1 R p ¯ f (x) − f (p)¯ d x, h < 0,
¯ ¯
h p+h
□
Remarks. (1) If f is not continuous at p, that does not necessarily imply that F is not differen-
tiable at p. For instance, in the next assignment, you will show that if
0, x ∈ [−1, 0) ∪ (0, 1],
f (x) =
1, x = 0,
then f is integrable,r and F (x) = 0 for all x ∈ [−1, 1]. However, F ′ (0) ̸= f (1). On the other hand,
consider
0, x ∈ [−1, 0),
f (x) =
1, x ∈ [0, 1],
Then,
and 0, x ∈ [−1, 0),
F (x) =
x, x ∈ [0, 1],
44
Remark. If F and G are two primitives of f on (a, b), then (F − G)′ = 0. Thus, F = G + C for some
constant C .
x r +1
Example 28. sin(x) and cos(x) are primitives of cos(x) and − sin(x), respectively. r +1 is a primi-
r
tive of x on (0, ∞) for all r ∈ R \ {1}.
By the First FTOC, we also know that
Z x 1
F (x) = dt, x > 0,
1 t
1
is a primitive of f (x) = x
on (0, ∞).
Theorem 5.14 (The second Fundamental Theorem of Calculus). Let I be an open interval, ¯ and
f : I → R be a function that admits an anti-derivative F on I . Let [a, b] ⊆ I , and suppose f ¯ is
¯
[a,b]
Riemann integrable. Then,
Z b
f (x)d x = F (b) − F (a).
a
END OF LECTURE 31
By taking common refinements, we may assume that s ε and t ε are step functions with respect
to the same partition P = {x 0 < · · · < x n }, i.e., there exist s 1 , ..., s n ∈ R and t 1 , ..., t n ∈ R such that
s ε (x) = s j and t ε (x) = t j for all x ∈ (x j −1 , x j ).
Now, by MVT, there exist c j ∈ (x j −1 , x j ) such that
n n n
F ′ (c j )(x j − x j −1 ) =
X X X
F (b) − F (a) = F (x j ) − F (x j −1 ) = f (c j )(x j − x j −1 ).
j =1 j =1 j =1
merely means that the derivative of sin(x) is cos(x). This symbol (without limits) has nothing to
R
do with integration a priori. You should simply read “ f (x)d x" as "the general primitive of f "
The two fundamental theorems tell us that:
R Rx
1. For a continuous f , f (x)d x = a f (t )d t +C .
Rb R ¯b
2. a f (x)d x = f (x)d x ¯
¯
a
The FTOC’s allow us to convert theorems about differentiabilty to theorems about integrabil-
¤′
ity. For instance, suppose F ′ = f . We know that F ◦ g (x) = F ′ (g (x))g ′ (x) = f (g (x))g ′ (x). Not
£
46
The above integral exist due the continuity of f of J . By the FFTOC, we have that F is differen-
tiable on J , and
F ′ (x) = f (x), x ∈ J.
Since g (I ) ⊂ J , we have that F ◦ g is well-defined and differentiable on I . By the chain rule,
for all x ∈ [a, b]. Thus, F ◦ g is an anti-derivative of ( f ◦ g )g ′ on I . By the SFTOC, applied twice,
Z b Z g (b)
′
f (g (x))g (x)d x = F (g (b)) − F (g (a)) = f (x)d x.
a g (a)
One can similarly obtain the formula for Integration by Parts as a consequence of the Funda-
mental Theorems of Calculus and the Product Rule for derivatives, but we will skip that.
END OF LECTURE 32
5.1. Logarithm and exponentiation. For any b > 0, we wish to define logb (x) as that real number
z such that
(5.1) b z = x.
However, in this course, we are yet to properly define what b z means if z is an irrational number.
Moreover, why should such a z exist? Why should it be unique. Let us extract the key features of
the desired log function from (5.1).
47
Say, z = " logb (x)" and w = " logb (y)", i.e., b z = x and b w = y. Then, if powers are to behave as
expected, b z+w = x y, i.e., logb (x) + logb (y) = z + w = " logb (x y)". Thus, we seek a function ℓ such
that
ℓ(x y) = ℓ(x) + ℓ(y)
for all x, y in its domain. Some observations based on this functional equation:
• If 0 ∈ dom(ℓ), then taking x = y = 0, we have that ℓ(0) = 2ℓ(0) ⇒ ℓ(0) = 0. Taking only
y = 0, we have that ℓ(0) = ℓ(x) for all x ∈ dom(ℓ). Thus, we only get the constant zero
function.
• If 1, −1 ∈ dom(ℓ), then taking x = y = 1 gives that
ℓ(1) = 0,
yℓ′ (y x) = ℓ′ (x).
The choice ℓ (1) gives the trivial solution. For any non-zero choice of ℓ′ (1), ℓ(x)/ℓ′ (1) gives
′
Proof. □
Definition 5.18. Let e denote the unique positive number such that ln(e) = 1. For any x ∈ R, let
exp(x) denote the unique positive number y such that ln(y) = x, i.e., the function exp(x) is the
inverse of ln(x).
The constant e was defined earlier using an infinite series. This turns out to be the same e!
Proof. □
a x = exp(x ln a).
END OF LECTURE 33