All Notes
All Notes
Contents
1 Real numbers 7
1.2 Intervals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2 Sequences 17
3 Functions 31
4 Functions 39
4.4 Limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4.5 Continuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
5 Differentiation 73
6 Integration 91
6.2 Anti-derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
The number systems that we use in calculus are the natural numbers, the integers, the rational numbers,
and the real numbers. Let us describe each of these :
If a number n is an integer, then we use the short-hand n ∈ Z. If n is a positive integer, then we use
n ∈ Z+ . Therefore N = Z+ . Numerous other short-hand notions are used for example, non-negative
integers: Z+ ∪ {0} or N0 .
A decimal number of the form x = 3.16792 is actually a rational number, for it represents
316792
x = 3.16792 = .
100000
8 calculus
A decimal number of the form
m = 4.27519191919 . . . ,
with a group of digits that repeats itself indefinitely, is also a rational number. To see this, notice that
100 · m = 427.519191919 . . .
100m = 427.519191919 . . .
m = 4.27519191919 . . .
99m = 423.244
or
423244
m=
99000
So, as we asserted, m is a rational number or quotient of integers. To indicate recurring decimals we
sometimes place dots over the repeating cycle of digits, e.g., m = 4.2751̇9̇,
19
6 = 3.16̇.
Another kind of decimal number is one which has a non-terminating decimal expansion that does not
keep repeating. An example is π = 3.14159265 . . . . Such a number is irrational, that is, it cannot be
expressed as the quotient of two integers. There are two different short-hand for a set of all irrational
numbers namely I or R \ Q. Clearly, we have
N⊂Z⊂Q⊂R
There is a subclass of real numbers called algebraic numbers denoted by A. These are numbers that
√
can be found from solving an algebraic equation with integer coefficients i.e x2 − 2 = 0 therefore 2 is
an algebraic number. Clealry all rational numbers and some irrational numbers are algebraic numbers.
Is there an example of a number that is not algebraic? Yes and these numbers are called transcendental
numbers. The two most important transcendental numbers are π and e.
(A2) a + b = b + a ∀ a, b ∈ R. (Commutativity)
(A3) a + (b + c) = ( a + b) + c ∀ a, b, c ∈ R. (Associativity)
(A4) R contains an element called identity denoted by 0 such that a + 0 = a for all R (Identity)
(A5) For every a ∈ R, there exists an element of R called the inverse and denoted by − a such that a + (− a) =
(− a) + a = 0. (Inverse)
Axioms of multiplication
(M2) a · b = b · a. (Commutativity)
(M3) a · (b · c) = ( a · b) · c. (Associativity)
(M5) For every a ∈ R, a ̸= 0 there exists an element of R called the inverse and denoted by 1
a such that
1
a· = 1. (Inverse)
a
Distributive law
(D1) a · (b + c) = a · b + a · c. distributivity
Oder axioms
These axioms (A1)–(A5), (M1)–(M5) define an algebraic structure called a field. If we add (O1)–(O4)
we get an ordered field. From these rules we can build more a more theory of numbers and they are
precisely the rules we use solve algebraic equations and inequalities. Finally we add the completeness
axiom. In informal terms the completeness axiom says that real numbers form a continuum with no
gaps between any two real numbers.A more detailed exposition on the subject of real numbers and
application of the above axioms will be discussed in a later course in Real Analysis]
Definition 1.6.
A subset of the real line is called an interval if it contains at least two numbers and all the real numbers between
any of its elements.
Examples :
1. Open interval ( a, b) includes all points between a and b, but not a and b.
3. Half-open interval we mean an open interval ( a, b) together with one of its endpoints. There are two such
intervals are [ a, b) and ( a, b].
. Infinite Intervals. Let a be any number. The set of all points x such that a < x is denoted by ( a, ∞),
the set of all points x such that a ≤ x is denoted by [ a, ∞). Similarly, (−∞, b) denotes the set of all
points x such that x < b and (−∞, b] denotes the set of all x such that x ≤ b.
It is a quantity that gives the magnitude or size of a real number. The absolute value or modulus of a
real number x, denoted by | x |, is given by
(
x, if x ≥ 0
|x| =
− x, if x < 0.
1. | x | ≥ 0
3. −| x | ≤ x ≤ | x |,
4. | − x | = | x |,
5. | x − y| = |y − x |,
7. | xy| = | x | · |y|,
x |x|
8. y = |y| if y ̸= 0.
12. x2 = | x |2
The principle of mathematical induction is used to prove some mathematical statements that are for-
mulated in terms of positive integers. We begin with function called propositional functions. This
function, denoted by P(n) represents some statement about positive integers n. The statement can
either true or false. Our goal in induction is to show that P(n) is true for all values of n > a where a is
fixed number. The steps that we will follow in all induction proofs is given below.
1. Verify the P(a) is true i.e. the statement is true for n = a. This is called the base step.
2. Assume the statement is true for n = k i.e. assume P(k ) is true. This is called the inductive
assumption or inductive hypothesis.
4. Hence conclude by the Principle f Mathematical Induction that P(n) is true for all n ≥ a.
Base Case:
1(1 + 1) 2
LHS = =1 and RHS = = 1,
2 2
Hence the statement is true for n = 1 i.e. P(1) is true.
Inductive step:
Next we assume that the statement holds for n = k, that is P(k) is true,
k ( k + 1)
1+2+···+k = .
2
We use this assumption to prove the statement holds for n = k + 1. Now,
1 + 2 + · · · + k + ( k + 1) = (1 + 2 + · · · + k ) + ( k + 1)
k ( k + 1)
= + (k + 1) (by inductive hypothesis)
2
k ( k + 1) + 2( k + 1)
=
2
k2 + 3k + 2
=
2
(k + 1)(k + 2)
=
2
(k + 1) [(k + 1) + 1]
=
2
Therefore P(k + 1) is true. Hence by the principle of mathematical induction P(n) is true for all n.
Example 1.2.
Prove that for any natural number
1 + 3 + 5 + · · · + 2n − 1 = n2 .
Solution.
Let P(n) be the statement that “1 + 3 + 5 + · · · + 2n − 1 = n2 ”
Base step:
Inductive step:
real numbers 13
1 + 3 + 5 + · · · + 2k − 1 = k2 .
1 + 3 + 5 + · · · + 2n − 1 = n2
Example 1.3.
Prove that 3n > 2n for all natural numbers n.
Solution.
Let P(n) be the statement “3n > 2n for all natural numbers n.”
Base step:
Inductive step:
3k +1 = 3k · 3
> 2k · 3 by inductive hypothesis
> 2k · 2 since 3 > 2
> 2k +1 ,
which is true. Hence, by thr principle of mathematical induction P(n) is true for all n.
Many induction problems that you will come across in mathematics are disguised, that is the problem
may not precisely ask you to use induction to solve the problem. Here is a good example.
14 calculus
Example 1.4.
Which the two expressions n! and 2n is bigger?
Solution.
We can do some trial and error to see what’s going on
n n! 2n n! > 2n ?
1 1 2 No
2 2 4 No
3 6 8 No
4 24 16 Yes
5 120 32 Yes
6 720 64 Yes
It looks like, initially n! < 2n but n! seems to be growing more rapidly. From this, we claim that n! > 2n for al
n ≥ 4. We can then proceed to prove this by induction.
Let P(n) be the statement “ n! > 2n for all n ≥ 4. The base case is when n = 4.
Base step:
From the above table we can see that 4! > 24 . Therefore P(4) is true.
Inductive step:
(k + 1)! = (k + 1)k!
> ( k + 1 ) 2k By the inductive hypothesis
k
> 2·2 Since k + 1 > 4 + 1 > 2
k +1
=2
Therefore P(k + 1) is true. Hence by the principle f mathematical induction we have P(n) is true for all n ≥ 4.
Thus two answer the initial question n! < 2n for n = 1, 2, 3 but n! > 2n for all n ≥ 4.
Example 1.5.
Find all positive integers n such that 22n − 1 is prime number.
real numbers 15
Solution.
Again we can start with some trial and error
It seems that all the values we are getting are not prime. In fact, on closer inspection we can see that all these
cases so far we have a multiple of 3. Can we conclude that 22n − 1 is a multiple 3 for all n. We can use induction
to prove this assertion. Let P(n) be the statement that “22n − 1 is divisible by 3” for n ≥ 1.
Base step:
Inductive step:
Next we assume P(k) is true, that is, for k ≥ 1, 22k − 1 is divisible by 3, i.e., 22k − 1 = 3m, for some m ∈ Z.
22(k+1) − 1 = 4 · 22k − 1
= (1 + 3) · 22k − 1
= ·22k − 1 + 3 · 22k by the inductive hypothesis
2k
= 3m + 3 · 2
Therefore 22(k+1) − 1 is divisible by 3. Thus the statement holds true for n = k + 1. Hence, by induction 22n − 1
is divisible by 3 for all n ≥ 1.
To go back to the original question since 22n − 1 is a multiple of 3 for all n we can conclude n natural number n
exist such that 22n − 1 is prime.
2 Sequences
Example 2.1.
a) 1, 2, 3, 4, 5, · · · , ( Natural numbers)
b) 1, 4, 9, 16, 25, · · · , ( Per f ect squares)
c) 2, 3, 5, 7, 11 · · · , ( Prime numbers)
d) 3, 3, 3, 3, 3, · · · , (Constant sequence)
1 1 1
e) 1, , , , · · · , ( Geometric progression)
2 4 8
1 1 1 1
f) 1, − , , − , , · · · ,
3 5 7 9
g) − 3, 9, −15, 33, −63, · · · ,
There are many ways for representing a general sequence (un ), {un }∞
n=1 , { u n } n∈N e.t.c. Some sequences
have general formula for the n-th term, (un ). The general formulae for the sequences in Example 2.1
are given below.
Example 2.2.
18 calculus
a) un = n
b) u n = n2
c) None
d) un = 3
1
e) u n = 2− n or un =
2n
(−1)n+1
f) un =
2n + 1
g) 1 + (−1)n · 2n+1
Some sequences can also be described using a recursive formula that is the next term is expressed in
terms of previous terms.
Example 2.3.
A sequence is given by the following formula.
un+1 = 2un + 1, u1 = 1
u1 = 1
u2 = 2u1 + 1 = 2(1) + 1 = 3
u3 = 2u2 + 1 = 2(3) + 1 = 7
u4 = 2u3 + 1 = 2(7) + 1 = 15
Another famous sequence that can be described using a recursive formula is the Fibonacci sequence.
u n +2 = u n +1 + u n , u1 = 1, u2 = 1
u1 = 1
u2 = 1
u3 = u2 + u1 = 1 + 1 = 2
u4 = u3 + u2 = 2 + 1 = 3
u5 = u4 + u3 = 3 + 2 = 5
u6 = u5 + u4 = 5 + 3 = 8
u7 = u6 + u5 = 8 + 5 = 13
sequences 19
Sequence can be finite or infinite. Finite sequences have a last term but infinite sequences do not.
Finite sequence are not that interesting so we will mainly concentrate on infinite sequences.
For infinite sequences we are interested in knowing the limit of a sequence, i.e. the value that a
sequence is converging to. For example, the sequence
3 4 5 6 7
2, , , , , ,···
2 3 4 5 6
n +1
is approaching zero. The n-th term for this sequence is given by un = n . We say limit of {un } as n
approaches infinity is 1. In shorthand notion, we say
n+1
lim un = 1 or lim =1
n→∞ n→∞ n
If { an } is a convergent sequence, it means that the terms an can be made arbitrarily close to l for n
sufficiently large. To get a feel of what this definition is about, lets look at the limit of un = n1 as n
approaches in infinity using the following steps.
1
1. Choose ε = 10
1 1
|u11 − l | = −0 < = ε if we choose n = 11
11 10
1 1
|u20 − l | = −0 < = ε if we choose n = 20
20 10
1 1
|u100 − l | = −0 < = ε if we choose n = 100
100 10
1 1
|un − l | = −0 <
n 100
1 1
|u120 − l | = −0 < =ε if we choose n = 120
120 10
1 1
|u200 − l | = −0 < =ε if we choose n = 200
200 10
1 1
|u600 − l | = −0 < =ε if we choose n = 600
600 10
5. If we choose ε to be as small as we like can we always find N such that |un − l | < ε. In this case it
clear to see that if we choose N = 1ε , then for all n > N we have
1
|un − l | = − 0 < ε.
n
Example 2.5.
1
Prove that lim = 0.
n→∞ n
Solution.
Let ε > 0, we need to find N (ε) such that
1
|un − l | = − 0 < ε.
n
Now
1 1
−0 =
n n
1
=
n
<ε
Thus if we take N = 1ε . Then for all n > N we have |un − l | < ε i.e. lim 1
= 0.
n→∞ n
Example 2.6.
n+1
Prove that lim = 1.
n→∞ n
sequences 21
Solution.
Let ε > 0, we need to find N (ε) such that
n+1
|un − l | = − 1 < ε.
n
Now
n+1 n+1−n
−1 =
n n
1
=
n
<ε
n +1
if we take N = 1ε . Thus for all n > N we have |un − l | < ε i.e. lim n = 1.
n→∞
Example 2.7.
Find the limit of the following sequence
7 10
4, , , . . .
2 3
Solution.
This sequence can be written as follows
We can definitively prove that lim un = 3 using the formal definition of a limit. That we want to show that for
n→∞
every ε, there exists a natural number N such that for all natural numbers greater than N we have |un − 3|. To
begin, choose ε > 0. Then, we have
1 + 3n
| u n − 3| = −3
n
1
=
n
<ε
if we take N = 1ε . Thus for all n > N we have |un − 3| < ε i.e. lim un = 3.
n→∞
If the limit of a sequence exists, the sequence is called convergent, otherwise, it is called divergent.
Example 2.8.
Use the definition of a limit to prove that
2n − 1 2
lim =
n→∞ 3n + 2 3
22 calculus
Solution.
Let ε > 0, we can find N (ε) such that
2n − 1 2 3(2n − 1) − 2(3n + 2)
− =
3n + 2 3 3(3n + 2)
6n − 3 − 6n − 4
=
3(3n + 2)
−7
=
3(3n + 2)
7
=
3(3n + 2)
7
<
9n
<ε
7
if we take N = 9ε then for all n > N we have
2n − 1 2
− <ε
3n + 2 3
Hence,
2n − 1 2
lim =
n→∞ 3n + 2 3
Example 2.9.
Prove that
2n2
lim =0
n→∞ n3 + 2
Solution.
Let ε > 0, Then we have
2n2
|un − l | = −0
n3 + 2
2n2
=
n3 + 2
2n2
<
n3
2
=
n
<ε
2n2
if we take N = 2ε . Thus for n > N, we have |un − l | < ε i.e. lim n3 +2
=0
n→∞
Example 2.10.
Prove that
sin(n2 )
lim √ =
n→∞ 3
n
Solution.
Let ε > 0, Then we have
sin(n2 )
|un − l | = √
3
−0
n
sin(n2 )
= √3
n
1
< √3
since | sin(θ )| < 1 for all θ
n
<ε
sin(n2 )
1
if we take N = ε3
. Thus for n > N, we have |un − l | < ε i.e. lim √3 n =0
n→∞
1, −1, 1, −1, 1, −1 · · ·
1
Suppose the limit of this sequence is l. Also suppose xn − 1, then xn+1 = −1. Take ε = 2 and since l is
the limit of this sequence, then there exits N such that for all n > N we have
1
| Xn − l | <
2
and
1
| X n +1 − l | <
2
Therefore we have
1
|1 − l | <
2
and
1
| − 1 − l| <
2
Notice that we can write 2 as (1 − l ) − (−1 − l ). Therefore
Theorem 2.3.
Let { an }, {bn } and {cn } be sequence with the following limits
lim an = a
n→∞
lim bn = b
n→∞
lim cn = c, c ̸= 0
n→∞
an lim an a
5. lim = n→∞ =
n→∞ cn lim cn c
n→∞
Proof:
1. We want show that for all ε > 0, there exists N > 0 such that for all n > N we have
|( an + bn ) − ( a + b)| < ε
sequences 25
Since lim an = a it means for ε > 0 there exists a N1 > 0 such that for all n > N1 we have
n→∞
ε
| an − a| <
2
And lim bn = b means for ε > 0 there exists a N2 > 0 such that for all n > N2 we have
n→∞
ε
| bn − b | <
2
Thus if we take N = max{ N1 , N2 } we have for all that ε > 0
3. We want show that for all ε > 0, there exists N > 0 such that for all n > N we have
|k · an − k · a| < ε
Since lim an = a, it means for ε > 0 there exists a N1 > 0 such that for all n > N1 we have
n→∞
ε
| an − a| <
|k| + 1
|k · an − k · a| = |k( an − a)|
= |k|| an − a|
ε
< |k| ·
|k| + 1
< ε for all n > N
4. We want show that for all ε > 0, there exists N > 0 such that for all n > N we have
| a n · bn − a · b | < ε
Since lim bn = b we know by Theorem 2.2 that {bn } is bounded i.e there exists M > 0 such that
n→∞
|bn | < M.
Also an → a and therefore for ε > 0 there exists a N1 > 0 such that for all n > N1 we have
ε
| an − a| <
2M
And bn → b means for ε there exists a N2 > 0 such that for all n > N2 we have
ε
| bn − b | <
2(| a| + 1)
26 calculus
Thus if we take N = max{ N1 , N2 } we have for all that ε > 0
| a n · bn − a · b | = | a n · bn − a · bn + a · bn − a · b |
= |bn ( an − a) + a(bn − b)|
< |bn ( an − a)| + | a(bn − b)|
= |bn || an − a| + | a||bn − b|
ε ε
< M· + | a| ·
2M 2(| a| + 1)
ε ε
< + < ε for all n > N
2 2
|c|
|cn − c| <
2
We also have the following
|c| = |c − cn + cn |
< |c − cn | + |cn |
< |cn − c| + |cn |
|c|
< + |cn |
2
Rearranging we get
|c| |c|
|cn | > |c| − =
2 2
Thus we have
1 2
<
|cn | |c|
1 1 c − cn
− =
cn c c · cn
cn − c
=
c · cn
|cn − c|
=
|c| · |cn |
2
< 2 · |cn − c|
|c|
sequences 27
Going back the fact that lim cn = c again we know that this means there exist N2 > N1 such that for
n→∞
all n > N2 we have
| c |2
|cn − c| < ·ε
2
Thus we now have
1 1 2
− < 2 · |cn − c|
cn c |c|
2 | c |2
< · ·ε
| c |2 2
<ε
Therefore
1 1
lim =
n→∞ cn c
We finally prove that
an a
lim =
n→∞ cn c
Since
1 1
lim an = a and lim =
n→∞ n→∞ cn c
Then by part 4) we have
an 1 1 1 a
lim = lim an · = lim an · lim = a· =
n → ∞ bn n→∞ cn n → ∞ n → ∞ cn c c
6. We want show that for all ε > 0, there exists N > 0 such that for all n > N we have
|k − k| < ε
We proceed as follows: If we take to be any positive number, then for all n > N we have
|k − k| = 0
< ε for any n > N
These properties are useful at breaking down complicated limits into simpler limits.
5 − 2n2
1 3
We want to be able to evaluate limits, for example, of the form lim 2− + 2 or lim .
n→∞ n n n→∞ 4 + 3n + 2n2
Example 2.11.
Evaluate the following limits
1 3
1. lim 2− + 2 .
n→∞ n n
28 calculus
3n2 − 5n
2. lim .
n→∞ 5n2 + 2n − 6
√ √
3. lim ( n + 1 − n).
n→∞
Solution.
1.
1 3 1 1
lim 2− + 2 = lim 2 − lim + 3 lim 2
n→∞ n n n→∞ n→∞ n n → ∞ n
= 2−0+0
=2
2.
lim 3 − n5
3n2 − 5n n→∞
lim =
n→∞ 5n2 + 2n − 6
lim 5 + n2 − 6
n2
n→∞
We know that
5 5
lim 3− = lim 3 − lim = 3−0 = 3
n→∞ n n→∞ n→∞ n
and
2 6 2 6
lim 5+ − 2 = lim 5 + lim − lim 2 = 5 + 0 + 0
n→∞ n n n→∞ n→∞ n n → ∞ n
lim 3 − n5
3n2 − 5n n→∞ 3
lim = =
n→∞ 5n2 + 2n − 6 5
lim 5 + n2 − 6
n2
n→∞
This quantity is undefined, so this approach will not help us solve this problem. This we need to try a new
approach
Whenever we are confronted with a expression of the form
p p
some expression − another expression,
we multiply by term in a way similar to multiplication by complex conjugates (remember high school maths!)
Thus we have
√ √
√ √ √ √ n+1+ n
n+1− n = n+1− n· √ √
n+1+ n
sequences 29
Proof: Since lim an = l it means that for all ε > 0 there exists N1 > 0 such that
n→∞
Definition 3.1 (b) A function f is strictly increasing on an interval I if for all points x and y in I with
Definition 3.1 (c) A function f is monotone decreasing on an interval I if for all points x and y in I with
Definition 3.1 (d) A function f is strictly decreasing on an interval I if for all points x and y in I with
Definition 3.1 (e) A function f is monotone on interval I if f is either monotone increasing or monotone
decreasing.
Example 3.1.
Consider the function
f ( x ) = (2x − 1)( x + 5)
We observe that f is increasing on the interval (−9/4, ∞) and is decreasing on the interval (−∞, −9/4).
32 calculus
Definition 3.2 (Bounded Functions).
A function f is
1. bounded above if there is a real number M such that f ( x ) ≤ M for all points x in its domain. The number
M is then called an upper bound of f .
2. bounded below if there is a real number m such that f ( x ) ≥ m for all points x in its domain. The number
m is then called a lower bound of f .
3. bounded if f is bounded above and below, that is, there exist real numbers M and m such that m ≤ f ( x ) ≤ M
for all points x in its domain.
Example 3.2.
The function f ( x ) = x + 3 is bounded in −1 ≤ x ≤ 1. An upper bound is 4 (or any number greater than 4). A
lower bound is 2 (or any number less than 2).
Observe that boundedness depends on the domain. For example, the function h( x ) = tan x is bounded
on the interval 0, π4 but not bounded on the interval 0, π2
Elementary Functions
f ( x ) = a 0 x n + a 1 x n −1 + · · · + a n −1 x + a n (3.1)
where a0 , a1 , . . . , an are constants and n is a positive integer called the degree of the polynomial provided a0 ̸= 0.
Example.
x5 + 10x3 − 2x + 1 is a polynomial of degree 5.
Example.
An example of a rational function is
x3 + x + 5
f (x) =
x2 − 3x − 4
is a rational function. Since ( x + 1)( x − 4) = 0 for x = −1 and x = 4, the domain of f is the set of all real
numbers except −1 and 4.
functions 33
Example.
Examples of power functions are
1
y=
x
1
y = x2
2
y = x3
Definition 3.6.
A piecewise defined function is a function described by using different formula on different parts of its domain.
Example.
1.
−1,
x<0
f (x) = 0, x=0
x + 2, x>0
x
34 calculus
2.
− x + 2, x<0
g( x ) = 2
x , 0≤x≤1
4, x > 1.
Transcendental Functions
1. Exponential function, f ( x ) = a x , a ̸= 0, 1.
10 y
x y = 2x
−3 −2 −1 1 2 3
4
y = log2 ( x )
2
1 2 3 4 5
−2
−4
−6
3. Trigonometric functions
sin x
sin x, cos x, tan x = , csc x, cot x, sec x
cos x
. The graph of y = sin x is given below
y
y = sin x
1
x
100 200 300
−1
π/2 y = arcsin x
x
−1 1
−π/2
36 calculus
5. Hyperbolic Functions,
y = tanh x
x
−3 −2 −1 1 2 3
−1
Definition 3.7 (Even and Odd Functions).1. Let f ( x ) be a real-valued function of a real variable. Then f is
even if
f ( x ) = f (− x )
y
y = cos x
1
x
−300 −200 −100 100 200 300
−1
f ( x ) = f (− x )
functions 37
y
y = sin x
1
x
−300 −200 −100 100 200 300
−1
| x |, x2 , x4 , cos x, cosh x
Definition 3.8.
Determine whether the following function is odd or even
3x
f (x) =
x2 + 1
Solution.
Observe that
3(− x ) 3x
f (− x ) = 2
=− 2 = − f ( x ).
(− x ) + 1 x +1
The function is odd.
A function f can be combined with another function g by means of arithmetic operations to form other
f
functions, the sum f + g, difference f − g, product f g and quotient g are defined as :
1. ( f + g)( x ) = f ( x ) + g( x ).
2. ( f − g)( x ) = f ( x ) − g( x ).
3. ( f g)( x ) = f ( x ) g( x ).
f f (x)
4. g ( x ) = g(x) , g( x ) ̸= 0.
38 calculus
f
Example: If f ( x ) = 2x2 − 5 and g( x ) = 3x + 4. Find f + g, f − g, f g, g .
Solution:
Definition 4.1 (b) A function f is strictly increasing on an interval [ a, b] if for all points x and y in [ a, b] with
Definition 4.1 (c) A function f is monotone decreasing on an interval [ a, b] if for all points x and y in [ a, b]
with
x < y then f ( x ) ≥ f (y)
Definition 4.1 (d) A function f is strictly decreasing on an interval [ a, b] if for all points x and y in [ a, b]
with
x < y then f ( x ) > f (y)
Definition 4.1 (e) A function f is monotone on interval [ a, b] if f is either monotone increasing or monotone
decreasing.
Example 4.1.
Consider the function
f ( x ) = (2x − 1)( x + 5)
We observe that f is increasing on the interval (−9/4, ∞) and is decreasing on the interval (−∞, −9/4).
40 calculus
Definition 4.2 (Bounded Functions).
A function f is
1. bounded above if there is a real number M such that f ( x ) ≤ M for all points x in its domain. The number
M is then called an upper bound of f .
2. bounded below if there is a real number m such that f ( x ) ≥ m for all points x in its domain. The number
m is then called a lower bound of f .
3. bounded if f is bounded above and below, that is, there exist real numbers M and m such that m ≤ f ( x ) ≤ M
for all points x in its domain.
Example 4.2.
The function f ( x ) = x + 3 is bounded in −1 ≤ x ≤ 1. An upper bound is 4 (or any number greater than 4). A
lower bound is 2 (or any number less than 2).
Observe that boundedness depends on the domain. For example, the function h( x ) = tan x is bounded
on the interval 0, π4 but not bounded on the interval 0, π2
Elementary Functions
f ( x ) = a 0 x n + a 1 x n −1 + · · · + a n −1 x + a n (4.1)
where a0 , a1 , . . . , an are constants and n is a positive integer called the degree of the polynomial provided a0 ̸= 0.
Example.
x5 + 10x3 − 2x + 1 is a polynomial of degree 5.
Example.
An example of a rational function is
x3 + x + 5
f (x) =
x2 − 3x − 4
is a rational function. Since ( x + 1)( x − 4) = 0 for x = −1 and x = 4, the domain of f is the set of all real
numbers except −1 and 4.
functions 41
Example.
Examples of power functions are
1
y=
x
1
y = x2
2
y = x3
Definition 4.6.
A piecewise defined function is a function described by using different formula on different parts of its domain.
Example.
1.
−1,
x<0
f (x) = 0, x=0
x + 2, x>0
x
42 calculus
2.
− x + 2, x<0
g( x ) = 2
x , 0≤x≤1
4, x > 1.
Transcendental Functions
1. Exponential function, f ( x ) = a x , a ̸= 0, 1.
10 y
x y = 2x
−3 −2 −1 1 2 3
4
y = log2 ( x )
2
1 2 3 4 5
−2
−4
−6
3. Trigonometric functions
sin x
sin x, cos x, tan x = , csc x, cot x, sec x
cos x
. The graph of y = sin x is given below
y
y = sin x
1
x
100 200 300
−1
π/2 y = arcsin x
x
−1 1
−π/2
44 calculus
5. Hyperbolic Functions,
y = tanh x
x
−3 −2 −1 1 2 3
−1
Definition 4.7 (Even and Odd Functions).1. Let f ( x ) be a real-valued function of a real variable. Then f is
even if
f ( x ) = f (− x )
y
y = cos x
1
x
−300 −200 −100 100 200 300
−1
f ( x ) = f (− x )
functions 45
y
y = sin x
1
x
−300 −200 −100 100 200 300
−1
Example.
Definition 4.8.
Determine whether the following function is odd or even
3x
f (x) =
x2 + 1
Solution.
Observe that
3(− x ) 3x
f (− x ) = =− 2 = − f ( x ).
(− x )2 + 1 x +1
The function is odd.
A function f can be combined with another function g by means of arithmetic operations to form other
f
functions, the sum f + g, difference f − g, product f g and quotient g are defined as :
1. ( f + g)( x ) = f ( x ) + g( x ).
2. ( f − g)( x ) = f ( x ) − g( x ).
3. ( f g)( x ) = f ( x ) g( x ).
46 calculus
f f (x)
4. g (x) = g( x )
, g( x ) ̸= 0.
5. ( f ◦ g)( x ) = f g( x )
f
Example: If f ( x ) = 2x2 − 5 and g( x ) = 3x + 4. Find f + g, f − g, f g, g .
Solution:
4.4 Limits
if the value of f ( x ) gets arbitrarily closer to L as x gets closer and closer to a. For example,
lim x2 = 9
x →3
since x2 gets arbitrarily close to 9 as x approaches as close as one wishes to 3. That is if we can choose
any ε > 0 such that | f ( x ) − L| < ε we can always find another positive real number, called δ such
| x − a| < δ. This leads us to a formal definition of limit of a function.
In other words, if f ( x ) can get as close to L “as we like”, then we can always demonstrate that x gets
close to a
functions 47
Example 4.3.
Use the ε − δ definition to show that
lim( x + 1) = 2
x →1
Solution.
Looking at figure below we see that as x → 1 we have f ( x ) → 2.
x y = x+1
−3 −2 −1 1 2 3
−2
−4
However on a precise limit definition is able to definitively prove that such a limit exist.
To be certain that lim f ( x ) = 2, we need to show that ε > 0, there exists a δ > 0 such that
x →1
We proceed as follows;
| f ( x ) − L| = |( x + 1) − 2|
= | x + 1 − 2|
= | x − 1|
| f ( x ) − L | = | x − 1| < δ = ε
Example 4.4.
Use the ε − δ definition to show that
lim(4x − 3) = 5
x →2
Solution.
Looking at figure below we see that as x → 2 we have f ( x ) → 5.
48 calculus
10 y
x y = (4x − 3)
−3 −2 −1 1 2 3
−5
To show that lim f ( x ) = 5, we need to show that for a given ε > 0, we can find δ > 0 such that,
x →2
To proceed we have
| f ( x ) − L| = |(4x − 3) − 5|
= |4x − 8|
= 4| x − 2|
< 4δ
ϵ
= ε if we take δ = .
4
Thus if we take ϵ
4 then whenever | x − 2| < δ, we have | f ( x ) − 5| < ε.
Example 4.5.
Show that lim( x2 + 1) = 2.
x →1
Solution.
To show that lim f ( x ) = 2, we need to show that for a given ε > 0, we can find δ > 0 such that,
x →1
To proceed
| f ( x ) − 2| = | x 2 + 1 − 2|
= | x 2 − 1|
= |( x + 1)( x − 1)|
= | x + 1|| x − 1|
< | x + 1| δ
functions 49
In order to determine how small the δ we choose can be, we need to put a bound on the term x + 1. To do this,
we take an initial value δ = 1 (you could choose any number for initial δ but 1 is convenient!).
| x − 1| < 1
−1 < x − 1 < 1
0<x<2
1 < x+1 < 3
| f ( x ) − 2| = | x 2 + 1 − 2|
= | x + 1|| x − 1|
< | x − 1| δ
< 3δ since 1 < x + 1 < 3
<ε
Example 4.6.
Show that lim( x2 + 3x ) = 10.
x →2
Solution.
Let ε > 0, our goal is to find δ > 0 such that
To proceed
| f ( x ) − 10| = | x2 + 3x − 10|
= |( x + 5)( x − 2)|
= | x + 5|| x − 2|
We need to put a bound on the term ( x + 5). To do this, we take an initial value δ = 1. Then
| x − 2| < 1
−1 < x − 2 < 1
1 < x < 3.
6 < x+5 < 8
. Therefore,
| f ( x ) − 10| = | x2 + 3x − 10|
= | x + 5|| x − 2|
< 8δ
If we take δ = min 1, 8ε , then |( x2 + 3x ) − 10| < ε whenever | x − 2| < δ. Hence the limit is proved.
50 calculus
Example 4.7.
Prove that
2 2
lim =
x →5 x 5
Solution.
Let ε > 0, our goal is to find δ > 0 such that
2
| f ( x ) − | < ε whenever | x − 5| < δ.
5
To proceed
2 2 2
f (x) − = =
5 x 5
|2(5 − x )|
=
5| x |
|2( x − 5)|
=
5| x |
2| x − 5|
=
5| x |
1 2δ
<
|x| 5
1
We need to put a bound on the term |x| . Taking an initial value δ = 1. Then
| x − 5| < 1
−1 < x − 5 < 1
4<x<6
1 1 1
< <
6 x 4
Therefore
2 |2( x − 5)|
f (x) − =
5 5| x |
1 2δ
<
|x| 5
1 2δ
< ·
4 5
2δ
<
20
δ
=
10
2 2
If we take δ = min {1, 10ϵ}, then x = 5 < ε whenever | x − 5| < δ. Hence the limit is proved.
Example 4.8.
Prove that
x+4 1
lim =−
x →−6 2−x 4
functions 51
Solution.
Let ε > 0, our goal is to find δ > 0 such that
1
f (x) − − <ε whenever | x − (−6)| < δ.
4
To proceed
x+4 1
| f ( x ) − L| = +
2−x 4
4( x + 4) − (2 − x )
=
4(2 − x )
3 x+6
=
4 x−2
3 | x + 6|
=
4 | x − 2|
3δ 1
<
4 | x − 2|
1
Next we need to find a bound for | x −2| . Take an initial value of δ = 1. Then
| x − (−6)| = | x + 6| < 1
−1 < x + 6 < 1
−7 < x < −5
−9 < x − 2 < −7
7 < | x − 2| < 9
1 1 1
< <
9 | x − 2| 7
Therefore
x+4 1
| f ( x ) − L| = −
2−x 4
3 x+6
=
4 x−2
3 δ
< ·
4 7
3
= δ
28
x +4
Take δ = min 1, 28 1
3 ε , then 2− x − 4 < ε whenever | x − (−6)| < δ. Hence the limit is proved.
Example 4.9.
Use ϵ − δ definition to prove that
√
lim ( 19 − x ) = 3
n→10
Solution.
Let ε > 0, our goal is to find δ > 0 such that
√
| f ( x ) − L| = | 19 − x − 3|
√ √ √ √
√ √ ( x + a − b) · ( x + a + b)
x+a− b= √ √
x+a+ b
x+a−b
= √ √
x+a+ b
√
This for 19 − x − 3 this becomes
√ √
19 − x − 9
√ √
√ √ 19 − x + 9
= ( 19 − x − 9) · √ √
19 − x + 9
√ √ √ √
( 19 − x − 9) · ( 19 − x + 9)
= √ √
19 − x + 9
19 − x − 9
= √ √
19 − x + 9
10 − x
= √ √
19 − x + 9
Therefore
√
| f ( x ) − L| = | 19 − x − 3|
10 − x
= √ √
19 − x + 9
| x − 10|
= √ p
| 19 − x + 9|
1 √
As before, we have to look for a bound for √
19− x + 9
. We begin by taking a preliminary δ = 1. This implies
Therefore, we have
√
| f ( x ) − L| = | 19 − x − 3|
| x − 10|
= √ p
| 19 − x + 9|
| x − 10|
< √
8+3
1
< √ δ
8+3
n √ o
If we take δ = min 1, ( 8 + 3)ε , we have | f ( x ) − L| < ε whenever | x − 10| < δ. Hence the limit is proved.
Example 4.10.
Use ϵ − δ definition to prove that
3
lim =3
n →1 x
Solution.
Let ε > 0, our goal is to find δ > 0 such that
To proceed
3
| f ( x ) − L| = −3
x
3 − 3x
=
x
|3 − 3x |
=
|x|
| x − 1|
=3
|x|
1
< 3δ
|x|
1
As before, we have to look for a bound for |x| . We begin by taking a preliminary δ = 1. This implies
−1 < | x − 1| < 1
0<x<2
1
Unfortunately we can’t find a bound for |x| using δ = 1 as before. So we try δ = 12 . Then we have
1 1
− < | x − 1| <
2 2
1 3
<x<
2 2
2 1
< <2
3 x
54 calculus
Therefore, we have
3
| f ( x ) − L| = −3
x
1
< 3δ
|x|
< 3δ · 2
= 6δ
3
If we take δ = min 1, 6ε , we have x − 3 < ε whenever | x − 1| < δ. Hence the limit is proved.
Theorem 4.1.
Let lim f ( x ) exists then it unique.
x→a
Theorem 4.2.
If the limit of function f at point a exists then f is bounded i.e. there exists a δ > 0 and real number M > 0
such that
| f ( x )| < M for all | x − a| < δ
Theorem 4.3.
If lim f ( x ) = A, lim g( x ) = B, and h( x ) = k. Then
x→a x→a
(a) lim h( x ) = k.
x→a
1 1
(d) lim = , provided B ̸= 0.
x→a g( x ) B
f (x) lim f ( x ) A
(e) lim = x→a = , provided B ̸= 0.
x→a g( x ) lim g( x ) B
x→a
Proof:
(a) Trivial
functions 55
(b) We need to show that for all ε > 0 there exists δ > 0 such that
f ( x ) + g( x ) − ( A + B) < ε whenever | x − a| < δ
Since lim f ( x ) = A then for every ε > 0, there exists a δ1 > 0 such that
x→a
ε
f (x) − A < whenever | x − x0 | < δ1
2
and lim g( x ) = B means that for every ε > 0, there exists a δ2 > 0 such that
x→a
ε
g( x ) − B < whenever | x − a| < δ2
2
To proceed, we have
f ( x ) + g( x ) − ( A + B) = f ( x ) − A + g( x ) − B
= f ( x ) − A + g( x ) − B
To show that lim [ f ( x ) + g( x )] = ( A + B), we need to show that for each ε > 0 there exists δ > 0
x→a
such that
[ f ( x ) + g( x )] − ( A + B) < ε
We know that lim f ( x ) = A means that for each ε > 0 there exists δ1 > 0 such that
x→a
ε
g( x ) − B <
2
whenever | x − a| < δ1 . Also lim g( x ) = B means that for each ε > 0 there exists δ2 > 0 such that
x→a
ε
g( x ) − B <
2
whenever | x − a| < δ2 . Now, taking δ = min{δ1 , δ2 } then
[ f ( x ) + g( x )] − ( A + B) = [ f ( x ) − A] + [ f ( x ) − B]
< [ f ( x ) − A] + [ f ( x ) − B]
ε ε
= +
2 2
=ε
(c) We need to show that for all ε > 0 there exists δ > 0 such that
ε
f (x) − A < whenever | x − a| < δ1
2(| B| + 1)
Also since f has a limit at a, then f is bounded [see Theorem 4.2 ] i.e. there exists M such that
| f ( x )| < M in the vicinity of a.
Next we have lim g( x ) = B. That means that for every ε > 0, there exists a δ2 > 0 such that
x→a
ε
g( x ) − B < whenever | x − a| < δ2
2M
To proceed, we have
f ( x ) · g( x ) − A · B = f ( x ) · g( x ) − f ( x ) · B + f ( x ) · B − A · B
= f ( x ) g( x ) − B + B f ( x ) − A
< f ( x ) g( x ) − B + B f ( x ) − A
= | f ( x )|| g( x ) − B| + | B|| f ( x ) − A|
< M| g( x ) − B| + | B|| f ( x ) − A|
ε ε
< M· + | B| ·
2M 2(| B| + 1)
ε ε
< +
2 2
=ε
if we take δ = min{δ1 , δ2 }.
(d) Since the limit of g exists then for every ε > 0 there exists δ1 > 0 such that whenever | x − a| < δ1 we
have
1
| g( x ) − B| < | B|
2
We can now proceed as follows
| B| = | g( x ) − g( x ) + B|
< | g( x ) − B| + | g( x )|
1
< | B| + | g( x )|
2
Thus rearranging we have
1
| g( x )| > | B|
2
Again since the limit of g exists then for every ε > 0 there exists δ2 > 0 such that whenever
| x − a| < δ2 we have
1
| g ( x ) − B | < | B |2 ε
2
functions 57
1 1 B − g( x )
− =
g( x ) B g( x ) · B
| B − g( x )|
=
| g( x ) · B|
| g( x ) − B|
=
| g( x ) · B|
| g( x ) − B|
=
| g( x )| · | B|
1 1
lim =
x→a g( x ) B
We also know from (b) that the limit of a product of two functions is product the two respective
limits, therefore we have
f (x) 1 1 A
lim = lim f ( x ) · = A· =
x→a g( x ) x→a g( x ) B B
(f) Let h( x ) = c. Then lim h( x ) = c by part (a) and by part (c) we have
x→a
lim c f ( x ) = lim h( x ) · f ( x ) = c · A
x→a x→a
Example 4.11.
Evaluate the following limits
(a) lim( x + 3)
x →1
1
(b) lim ,
x →1 x+2
(c) lim( x2 − 7x + 5).
x →8
x2 − 4
(d) lim
x →2 x−2
58 calculus
Solution.
1 lim 1 1 1 1
x →1
2. lim = = = = ,
x →1 x+2 limx→1 ( x + 2) lim x + lim 2 1+2 3
x →1 x →1
2 2
3. lim( x − 7x + 5) = 8 − 7(8) + 5 = 13
x →8
x2 − 4 ( x + 2)( x − 2)
4. lim = lim = lim( x + 2) = 4.
x →2 x−2 x →2 x−2 x →2
Exercise 4.1.
Simplify the following limits
1. lim( x4 + 3x − 5)
x →2
x4 + x2 − 1
2. lim
x →1 x2 + 5
√
3. lim 2x2 − 1
x →5
2 +3x −2
5. lim ex
x →1
Not all limits can be found using the above properties. Some examples are given below
Example 4.12.
x−2
(a) lim
x →2 x2 − 4
1
sin θ − 2
(b) limπ
θ→ 6 θ − π6
x − 8x2
(c) lim
x →∞ 12x 2 + 5x
1
(d) lim (ln x ) x
x →∞
1
(e) lim(ex + x ) x
x →0
(f) lim x x
x →0+
√ √
5+h− 5
(g) lim
h →0 h
functions 59
In the table below, the second column shows the “limits” if we naively apply the properties above. The
third column shows the correct limit.
(a)
x−2 x−2
lim 2
= lim
x →2 x −4 x → 2 ( x − 2)( x + 2)
1
= lim
x →2 x + 2
1
=
4
(h)
√ √ √ √ √ √
5+h− 5 5+h− 5 5+h+ 5
lim = lim ·√ √
h →0 h h →0 h 5+h+ 5
√ √ √ √
( 5 + h − 5)( 5 + h + 5)
= lim √ √
h →0 h ( 5 + h + 5)
5+h−5
= lim √ √
h →0 h ( 5 + h + 5)
h
= lim √ √
h →0 h ( 5 + h + 5)
1
= lim √ √
h →0 5+h+ 5
1
= √ √
( 5 + 5)
1
= √
(2 5
60 calculus
We will revisit the evaluation of the rest of the limits in Section 5.3 on the L’Hôpitals Rule.
So far we have considered the limit of a function f as x → a from either side of a. We can also compute
limits as x → a from from only one side. If the approach is from the right, the limit is a right-hand
limit or limit from the right and is denoted by lim f ( x ). Similarly, a left-hand limit is also called a
x → a+
limit from the left and is denoted by lim f ( x ). The definitions of such limits are
x → a−
Example 4.13.
Consider the following function.
− x2 + 1, x<0
f (x) =
x + 2, x>0
x
functions 61
lim f ( x ) = lim (− x2 + 1) = 0 + 1 = 1
x →0− x →0−
lim f ( x ) = lim ( x + 2) = 0 + 2 = 2
x →0+ x →0+
Example 4.14.
Consider the following function.
x2 − 1, x<0
g( x ) =
x − 1, x>0
lim g( x ) = lim ( x2 − 1) = 0 − 1 = −1
x →0− x →0−
lim g( x ) = lim ( x − 1) = 0 − 1 = −1
x →0+ x →0+
In Example 4.13 the left and right limits are unequal and in Example 4.14 the limits are equal. Clearly,
when the left-hand limit and right-hand limits are equal, the limit exist and where the left-hand and
right-hand limits are unequal the limit does not exist. Also notice that the limit at x = 0 is −1, even
though the function g has a different value at x = 0.
Theorem 4.4.
Let f be a function. Then we have
lim f ( x ) = L
x→a
62 calculus
if, and only if, both one-sided limits exist and be equal to L, that is,
lim f ( x ) = L
x → a−
lim f ( x ) = L
x → a+
Example 4.15.
|x|
Show that lim does not exist.
x →0 x
Solution.
Notice that
| x | xx = 1, if x≥0
=
x −x = −1, if x < 0.
x
Thus
|x|
lim = lim (−1) = −1.
x →0− x x →0−
|x|
lim = lim (1) = 1
x →0+ x x →0+
Since
|x| |x|
lim ̸= lim
x →0+ x x →0− x
|x|
it means that lim does not exist.
x →0 x
Then,
lim g( x ) = L
x→a
This theorem is sometimes called the Sandwich Theorem. It states that if a a function h is sandwiched
between two functions which are approaching L then h must also approach L.
Example 4.16.
2x ≤ g( x ) ≤ x4 − x2 + 2
Evaluate lim g( x )
x →1
functions 63
lim 2x = 2(1) = 2
x →1
2. Prove that
θ cos θ ≤ sin θ ≤ θ
θ cos θ sinθ θ
≤ ≤
θ θ θ
sinθ
cos θ ≤ ≤1
θ
sin θ
Let f (θ ) = cos θ, g(θ ) = θ and h(θ ) = 1. We know that f (θ ) ≤ g(θ ) ≤ h(θ ) and
sin θ
lim =1
θ →0 θ
64 calculus
√ √
xesin( x ) ≥ 0. On the other hand we have sin
π
(b) Observe that x ≥ 0 for all x ≥ 0. Therefore, π
x ≤ 1,
thus
esin( x ) ≤ e1
π
Therefore
√ √
xesin( x ) ≤ e x
π
lim f ( x ) = lim 0 = 0
x →0+ x →0+
and
√
lim h( x ) = lim e x = 0
x →0+ x →0+
=0
functions 65
We now go on to look at at one of the most important concept regarding real numbers: That is the
subject of continuity.
4.5 Continuity
Notice in the previous chapter that the definition of limit does to require the function to be defined at
the point of the limit for example if
sin x
f (x) =
x
lim f ( x ) = lim
x →0 x →0
Therefore
sin x
=1
x
but f is undefined at x = 0. In some cases the function is defined at point but lim f ( x ) ̸= f (c) An
x →c
example of this given below
Example 4.18.
Consider the following function.
2
x ,
x<1
h( x ) = 2, x=1
x, x>1
x
66 calculus
Notice that
lim h( x ) = lim ( x2 ) = 12 = 1
x →2− x →1−
lim h( x ) = lim ( x ) = 1
x →2+ x →1+
Since left-hand and right-hand limits exist and are both equal to 1. It means the
lim h( x ) = 1
x →1
The function in 4.18 passes at Condition [(a)] since h(1) is defined. It also passes Condition [(b)] since
lim h( x ) exists, but fails at Condition [(c)] since lim h( x ) = 1 ̸= h(1) since h(1) = 2
x →1 x →1
Example 4.19.
Determine whether the following functions are continuos at the point x = −1.
1. f ( x ) = x
2. f ( x ) = | x + 1|
3.
1 x < −1
f (x) =
−1 x ≥ −1
4.
−| x | x ̸ = −1
f (x) =
1 x = −1
1
5. f ( x ) = x +1
Solution.1. f ( x ) = x
functions 67
2. f ( x ) = | x + 1|
x
68 calculus
Condition [(a)] passes but conditions [(b)] and [(c)] fail. Therefore f is not continuous.
−| x | x ̸ = −1
4. f ( x ) =
1 x = −1
Condition [(a)] and [(b)] pass but conditions [(c)] fails. Therefore f is not continuous.
1
5. f ( x ) = x +1 All three conditions fail. Therefore f is not continuous.
The conditions in Definition 4.12 allow us to use ε − δ notions rephrase the definition of continuity.
Example 4.20.
Use the ε − δ definition to show that the following functions are continuous at the given points c.
(a) f ( x ) = 3x + 2, c=2
(b) f ( x ) = x2 , c=1
(c) f ( x ) = x2 + 2x + 1, c=2
√
(d) f ( x ) = x, c = 4
Solution.
functions 69
To proceed, we have
| f ( x ) − f (2)| = |(3x + 2) − 8|
= |3x − 6|
= 3| x − 2|
< 3δ
ε
< ε if we take δ =
3
Thus f ( x ) = 3x + 2 is continuous at x = 2.
2. We want to show that for each ε > 0 there exist δ > 0 such that if
To begin, we have
| f ( x ) − f (1)| = | x2 − (1)2 |
= | x 2 − 1|
= |( x − 1)( x + 1)|
| x − 1| < 1
−1 < x − 1 < 1
0<x<2
1 < x+1 < 3
1 < | x + 1| < 3
| f ( x ) − f (1)| = | x2 − (1)2 |
= |( x − 1)( x + 1)|
< 3| x − 1|
< 3δ
nε o
< ε if we take δ = min ,1
3
Therefore f ( x ) = x2 is continuous at x = 1.
3. We want to show that for each ε > 0 there exist δ > 0 such that if
= | x2 + 2x + 1 − 9|
= | x2 + 2x − 8|
= |( x + 4)( x − 2)|
To find a bind for | x + 4| we proceed as follows; we first take δ = 1 i.e. | x − 2| < 1 then
−1 < x − 2 < 1
1<x<3
5 < x+4 < 7
5 < | x + 1| < 7
4. We want to show that for each ε > 0 there exist δ > 0 such that if
To begin, we have
√ √
| f ( x ) − f (4)| = | x − 4|
√ √ √ √
( x − 4) · ( x + 4)
= √ √
x+ 4
x−4
= √ √
x+ 4
| x − 4|
= √ √
x+ 4
To find a bind for √ 1√ we proceed as follows; we first take δ = 1 i.e. | x − 4| < 1 then
x+ 4
−1 < x − 4 < 1
3<x<5
√
√ √
3< x< 5
√ √ √ √ √ √
3+ 4 < x+ 4 < 5+ 4
1 1 1
√ √ < √ √ < √ √
5+ 4 x+ 4 3+ 4
functions 71
| x − 4|
| f ( x ) − f (4)| = √ √
x+ 4
| x − 4|
< √ √
3+ 4
1
< √ √ δ
3+ 4
√
< ε if we take δ = ( 3 + 2)ε
n √ o
Thus i we take δ = min 1, ( 3 + 2)ε) we have | f ( x ) − f (4)| < ε we have | x − 4| < δ. Therefore
√
f ( x ) = x is continuous at x = 4.
1. k · f ,
2. f ± g,
3. f · g,
f
4. g provided g( x ) ̸= 0 ∀ x.
Proof: The same as proofs for limits of functions in Theorem 4.3, except that in the following
lim f ( x ) = A, lim g( x ) = B
x→a x→a
we have
A = f ( a ), B = g( a)
■
Example.
Let f ( x ) = 3x + 2 and g( x ) = x2 + 2x + 1. Both functions are continuous at x = 2 [see Examples (4.20) and
(4.20)]. Then the following are also continuous at x = 2
1. (3x + 2) − ( x2 + 2x + 1) = − x2 + x + 1
2. (3x + 2)( x2 + 2x + 1)
3x +2
3. x2 +2x +1
Example.
1. The function f ( x ) = x2 is continuous on the interval [−3, 5] since it is continuous at every point on the
interval [−3, 5]. In fact f is continuous on R.
2. The function g( x ) = x12 is not continuous on the interval [−3, 5] since it is not continuous at x = 0. However,
g is continuous on R \ {0}.
In simple terms the intermediate value theorem states that a continuous function takes on every inter-
mediate value between the function values f ( a) and f (b).
Example 4.21.
Show that
√
ln x = x − x
Solution.
√
Let f ( x ) = ln x − x + x. The function f is continuous on the interval [2, 3]. Further,
√
f (2) = ln 2 − 2 + 2 = 0.10736 . . . > 0
√
f (3) = ln 3 − 3 + 3 = −0.16933 . . . < 0
and also f (2) ̸= f (3). Set M = 0, therefore by Intermediate value theorem there exist c ∈ (2, 3) such that
√
f (c) = 0 i.e. ln x = x − x has a root in (2, 3)
5 Differentiation
The gradient of f at the point x is equal to the gradient of the red line. We can approximate this
gradient using the secant line [the grey line]. That is
f ( x + h) − f ( x ) f ( x + h) − f ( x )
gradient of secant line = =
( x + h) − h h
The gradient of secant line becomes a better approximation of the gradient at point x if we choose a
smaller h. Taking the limit as h approach 0 we get obtain the exact value of the gradient of of f at point
x. This gradient is called the derivative.
Solution.
Using definition (5.1) we have
f ( x0 + h ) − f ( x0 )
f ′ ( x0 ) = lim
h →0 h
[( x0 + h)3 − ( x0 + h)] − ( x03 − x0 )
= lim
h →0 h
[(1 + h)3 − (1 + h)] − (13 − 1)
= lim
h →0 h
(1)3 + 3(1)2 h + 3(1) h2 + h3 − 1 − h
= lim
h →0 h
2 3
2h + 3h + h
= lim
h →0 h
= lim(2 + 3h + h2 )
h →0
= 2.
Example 5.2.
Find the derivative (if it exists) of f ( x ) = | x | at x = 0.
Solution.
|0 + h | − |0|
f ′ (0) = lim
h →0 h
|h|
= lim
h →0 h
We are familiar with this function from Example 4.15. We concluded then that this limit does not exist. Therefore,
the function f is not differentiable at x = 0.
Notice in Example 5.2, that the derivative does not exist at a point where there is a “sharp” corner.
Thus, if a function has a derivative on an interval it means the function does not have “kinks” any-
where on the interval i.e. its smooth. Therefore the terms “smooth” and “differentiable” are used
interchangeably.
differentiation 75
Example 5.3.
Find the derivative of f ( x ) = 2x2 + 3x − 2
Solution.
Let x ∈ R
f ( x + h) − f ( x )
f ′ ( x ) = lim
h →0 h
2( x + h)2 + 3( x + h) − 2 − 2x2 + 3x − 2
= lim
h →0 h
2
2x + 4xh + 2h2 + 3x + 3h − 2 − 2x2 + 3x − 2
= lim
h →0 h
4xh + 2h2 + 3h
= lim
h →0 h
= lim(4x + 2h + 3)
h →0
= 4x + 3
Example 5.4.
√
Let f ( x ) = x, find the derivative of f on [0, ∞).
Solution.
Let x > 0.
f ( x + h) − f ( x )
f ′ ( x ) = lim
h →0 h
√ √
x+h− x
= lim
h →0 h
√ √ √ √
x+h− x x+h+ x
= lim ·√ √
h →0 h x+h+ x
(h
= lim √ √
h →0 h ( x + h + x)
( x + h) − x
= lim √ √
h →0 h ( x + h + x)
1
= lim √ √
h →0 x+h+ x
1 1
= √ √ = √ .
x+ x 2 x
Example 5.5.
Let f ( x ) = x n where n is any real number. Prove that
f ′ ( x ) = nx n−1
76 calculus
Proof:
f ( x + h) − f ( x )
f ′ ( x ) = lim
h →0 h
( x + h)n − x n
= lim
h →0 h
(n0 ) x n + (n1 ) x n−1 h + (n2 ) x n−2 h2 + · · · + (nn)hn − x n
= lim
h →0 h
x n + nx n−1 h + (n2 ) x n−2 h2 + · · · + (nn)hn − x n
= lim
h →0 h
nx n−1 h + (n2 ) x n−2 h2 + · · · + (nn)hn
= lim
h →0 h
n n −2 n n −1
= lim nxn−1 + x h+···+ h
h →0 2 n
= nx n−1 + 0 + · · · + 0
= nxn−1
Example 5.6.
Find the derivative of g( x ) = sin x on R
Solution.
g( x + h) − g( x )
g′ ( x ) = lim
h →0 h
sin( x + h) − sin x
= lim
h →0 h
sin x cos h + cos x sin h − sin x
= lim
h →0 h
sin x (cos h − 1) + sin h cos x
= lim
h →0 h
sin x (cos h − 1) sin h cos x
= lim +
h →0 h h
sin x (cos h − 1) sin h cos x
= lim + lim
h →0 h h →0 h
cos h − 1 sin h
= sin x · lim + cos x · lim
h →0 h h →0 h
cos h − 1 sin h
= sin x · lim + cos x · lim
h →0 h h →0 h
sin h cos h−1
Recall from Examples 4.16 and 4.17 that lim h = 1 and lim h = 0. Therefore,
h →0 h →0
g′ ( x ) = sin x · 0 + cos x · 1
= cos x
differentiation 77
f ′ ( x ),
y′ ,
dy
,
dx
d
( f ( x ))
dx
d
The symbol dx is called differentiation operator. The process of finding derivatives of functions is
called differentiation.
′
1. k =0
′
= f ( x )′ ± g( x )′
2. f ( x ) ± g( x )
′
3. [ f ( x ) · g( x )] = f ( x ) · g′ ( x ) + g( x ) · f ′ ( x ), Product rule
i′
g( x )· f ′ ( x )− f ( x )· g′ ( x )
h
f (x)
4. g( x )
= ( g( x ))2
. Quotient rule
5. [ f ◦ g]′ ( x ) = f ′ g( x ) · g′ ( x ).
Proof:
1. Let g( x ) = k. Then
g( x + h) − g( x )
g′ ( x ) = lim
h →0 h
k−k
= lim
h →0 h
0
= lim
h →0 h
=0
2. To prove that
′
= f ( x )′ ± g( x )′
f ( x ) ± g( x )
:
78 calculus
′ f ( x + h) + g( x + h) − f ( x ) + g( x )
f ( x ) + g( x ) = lim
h →0 h
f ( x + h) − f ( x ) + g( x + h) − g( x )
= lim
h →0 h
f ( x + h) − f ( x ) g( x + h) − g( x )
= lim +
h →0 h h
f ( x + h) − f ( x ) g( x + h) − g( x )
= lim + lim
h →0 h h →0 h
= f ′ ( x ) + g′ ( x )
3. To prove that
′
[ f ( x ) · g( x )] = f ( x ) · g′ ( x ) + g( x ) · f ′ ( x ) :
f ( x + h) · g( x + h) − f ( x ) · g( x )
( f · g)′ ( x ) = lim
h →0 h
f ( x + h) · g( x + h) − f ( x + h) · g( x ) + f ( x + h) · g( x ) − f ( x ) · g( x )
= lim
h →0 h
f ( x + h) · g( x + h) − f ( x + h) · g( x ) + f ( x + h) · g( x ) − f ( x ) · g( x )
= lim
h →0 h
f ( x + h) · g( x + h) − g( x ) + g( x ) · f ( x + h) − f ( x )
= lim
h →0 h
!
g( x + h) − g( x ) f ( x + h) − f ( x )
= lim f ( x + h) · + g( x ) ·
h →0 h h
! !
g( x + h) − g( x ) f ( x + h) − f ( x )
= lim f ( x + h) · + lim g( x ) ·
h →0 h h →0 h
! !
g( x + h) − g( x ) f ( x + h) − f ( x )
= lim [ f ( x + h)] · lim + lim [ g( x )] · lim
h →0 h →0 h h →0 h →0 h
= f ( x ) · g′ ( x ) + g( x ) · f ′ ( x )
4. To prove that
′
g( x ) · f ′ ( x ) − f ( x ) · g′ ( x )
f (x)
= .
g( x ) ( g( x ))2
differentiation 79
′ f f
f g ( x + h) − g ( x )
( x ) = lim
g h →0 h
f ( x +h)
g( x +h)
− gf ((xx))
= lim
h →0 h
f ( x +h)· g( x )− f ( x )· g( x +h)
g( x +h)· g( x )
= lim
h →0 h
f ( x + h) · g( x ) − f ( x ) · g( x + h)
= lim
h →0 g( x + h) · g( x ) · h
f ( x + h) · g( x ) − f ( x ) · g( x + h)
1
= lim · lim
h →0 g( x + h) · g( x ) h →0 h
f ( x + h) · g( x ) − f ( x ) · g( x ) + f ( x ) · g( x ) − f ( x ) · g( x + h)
1
= lim · lim
h →0 g( x + h) · g( x ) h →0 h
[ f ( x + h) x − f ( x )] · g( x ) − f ( x ) · [ g( x + h) − g( x )]
1
= lim · lim
h →0 g( x + h) · g( x ) h →0 h
[ f ( x + h) x − f ( x )] · g( x ) f ( x ) · [ g( x + h) − g( x )]
1
= 2
· lim − lim
[ g( x )] h→0 h h →0 h
[ f ( x + h) x − f ( x )] g( x + h) − g( x )
1
= 2
· g ( x ) · lim − f ( x ) · lim
[ g( x )] h →0 h h →0 h
g( x ) · f ′ ( x ) − f ( x ) · g′ ( x )
= 2
[ g( x )]
5. To prove that
[ f ◦ g]′ ( x ) = f ′ g( x ) · g′ ( x )
We begin with
[ f ◦ g]( x + h) − [ f ◦ g]( x )
[ f ◦ g]′ ( x ) = lim
h →0 h
f [ g( x + h)] − f [ g( x )]
= lim
h →0 h
f [ g( x + h)] − f [ g( x )] g( x + h) − g( x )
= lim ·
h →0 g( x + h) − g( x ) h
f [ g( x + h)] − f [ g( x )] g( x + h) − g( x )
= lim · lim
h →0 g( x + h) − g( x ) h →0 h
′
′
= f g( x ) · g ( x )
■
80 calculus
5.1 Differentiation formulae and techniques
It’s all high school stuff. We won’t do it here but you are expected know to all the formulae such as
′
= n · x n −1
xn
1.
′
2. cos x = − sin x
′ f ′ (x)
3. ln f ( x ) = f (x)
4. Implicit differentiation
8. etc. etc.
So you should dust up your high school notes on differentiation (and integration!).
Proof: Recall that a function is differentiable a point x0 if the following limit exist
f ( x ) − f ( x0 )
f ′ ( x0 ) = lim .
x → x0 x − x0
Since this limit exist it means f (c) is defined. Next we we show that lim f ( x ) = f ( x0 ). Cosinder all x
x → x0
with x ̸= c. Then we have
f ( x ) − f ( x0 )
f ( x ) − f ( x0 ) = ( x − x0 ) ·
x − x0
Taking limits of both sides we have
f ( x ) − f ( x0 )
lim ( f ( x ) − f ( x0 )) = lim ( x − x0 ) ·
x → x0 x → x0 x − x0
f ( x ) − f ( x0 )
= lim ( x − x0 ) · lim
x → x0 x → x0 x − x0
= 0 · f ′ ( x0 )
=0
Therefore we have
lim ( f ( x ) − f ( x0 )) = 0
x → x0
differentiation 81
or
lim f ( x ) − lim f ( x0 ) = 0
x → x0 x → x0
lim f ( x ) − f ( x0 ) = 0
x → x0
Finally, we have
lim f ( x ) = f ( x0 )
x → x0
Thus, the three condition of Definition 4.12 are satisfied i.e f is continuous. ■
f ( a) = f (b)
f (b) − f ( a)
f ′ (c) = ,
b−a
Geometrically, this means that there exists a point c in ( a, b) such that the tangent line to f at c is
parallel to the secant line joining the points ( a, f ( a)) and (b, f (b)).
82 calculus
Observe that Rolle’s theorem is a special case of the Mean value theorem. Another name for the Mean
value theorem is Lagrange’s Theorem.
The following theorem shows some of the immediate consequences of the Mean value theorem.
Theorem 5.5.
Let f be continuous n [ a, b] and differentiable on ( a, b).
Proof:
(a) Recall from Definition 4.1 (a) that a function is monotone increasing on [ a, b] if for all points x and y
in [ a, b] with
x < y then f ( x ) ≤ f (y)
Let x1 , x2 ∈ ( a, b) with x1 < x2 . Then f is continuous on [ x1 , x2 ] and differentiable on ( x1 , x2 ). Thus
by mean value theorem there exists c ∈ ( x1 , x2 ) such that
f ( x2 ) − f ( x1 )
f ′ (c) = ,
x2 − x1
Since f ′ ( x ) ≥ 0 for all x ∈ ( a, b). We have
f ( x2 ) − f ( x1 )
≥0
x2 − x1
=⇒ f ( x2 ) − f ( x1 ) ≥ 0
=⇒ f ( x2 ) ≥ f ( x1 )
differentiation 83
(c) Recall from Definition 4.1 (c) that a function is monotone decreasing om [ a, b] if for all points x and
y in [ a, b] with
x<y then f ( x ) ≥ f (y)
f ( x2 ) − f ( x1 )
f ′ (c) = ,
x2 − x1
f ( x2 ) − f ( x1 )
≤0
x2 − x1
=⇒ f ( x2 ) − f ( x1 ) ≤ 0
=⇒ f ( x2 ) ≤ f ( x1 )
Some applications of the mean value theorem are shown in the following examples.
Example 5.7.
Use The Mean Value Theorem to show that
Solution.
1. Consider the function f ( x ) = cos x is continuous and differentiable for all x on some interval [ a, b]. Thus by
the Mean Value Theorem there exists a
cos a − cos b
f ′ (c) = (cos c)′ = (− sin c) =
a−b
84 calculus
We know that | sin c| ≤ 1. Therefore, taking modulus of both sides, we get
cos a − cos b
= | sin c| ≤ 1,
a−b
Hence
| cos a − cos b| ≤ | a − b|
2. Let f ( x ) = tan−1 x. Then we have f ′ ( x ) = 1+1x2 . The function f is also continuous and differentiable for x
in some interval ( a, b). By the Mean Value Theorem there exist c such that a < c < b, and
f (b) − f ( a)
f ′ (c) =
b−a
Since f ′ (c) = 1
1+ c2
we have
a<c<b
a < c2 < b2
2
1 + a2 < 1 + c2 < 1 + b2
1 1 1
2
< 2
<
1+b 1+c 1 + a2
1 − 1
tan b − tan a− 1
1
< <
1 + b2 b−a 1 + a2
b−a b−a
2
< tan−1 b − tan−1 a < .
1+b 1 + a2
1. Let f ( x ) = ln x. Then f ′ ( x ) = 1x . The function f is continuous and differentiable for all x > 0. Take
0 < a < b then by the Mean Value Theorem, there exists c ∈ ( a, b) such that
f (b) − f ( a) ln b − ln a
f ′ (c) = =
b−a b−a
1 ln b − ln a
=
c b−a
differentiation 85
a<c<b
1 1 1
< <
b c a
1 ln b − ln a 1
< <
b b−a a
b−a b−a
< ln b − ln a <
b a
a b b
1 − < ln < − 1.
b a a
a b b
2. Setting a = 5 and b = 6 into 1 − b < ln a − 1, we have
< a
5 6 6
1 − < ln < −1
6 5 5
1 1
∴ < ln 1.2 < .
6 5
We conclude the chapter of differentiation by revisiting the problems in Example 4.12. Recall that we
called the limits of the form 00 , ∞ 0
∞ , 0 , the indeterminate form. The following theorem called L’Hôpitals
1
Rule is used to evaluate these limits. 1
Although this
rule is named
Theorem 5.6 (L’Hôpital’s rule). after the French
mathematician
Suppose f and g are differentiable and g′ ( x ) ̸= 0 on an open interval I that contains a (except possibly at a). Guillaume de
Suppose that L’H0̂pital, it
lim f ( x ) = 0 and lim g( x ) = 0 was developed
x→a x→a by the Swiss
or mathemati-
cian Johann
lim f ( x ) = ±∞ and lim g( x ) = ±∞ Bernoulli.
x→a x→a
Then
f (x) f ′ (x)
lim = lim ′
x→a g( x ) x→a g ( x )
Example 5.8.
Evaluate the following limits
sin 2x
(a) lim .
x →0 sin 5x
e3x
(b) lim .
x →∞ x
86 calculus
1
sin θ − 2
(c) limπ
θ→ 6 θ − π6
x−2
(d) lim
x →2 x2 − 4
x − 8x2
(e) lim
x →∞ 12x 2 + 5x
Solution.
(a) Let f ( x ) = sin 2x and g( x ) = sin 5x. Then the functions f and g are differentiable with f ′ ( x ) = 2 cos 2x
and g′ ( x ) = 5 cos 5x. We also have limx→0 sin 2x = 0 and limx→0 sin 5x = 0. Therefore by L’Hôpital’s rule
rule we have
sin 2x f (x) f ′ (x) 2 cos 2x 2·1 2
lim = lim = lim ′ = lim = =
x →0 sin 5x x →0 g ( x ) x →0 g ( x ) x →0 5 cos 5x 5·1 5
(b) Let f ( x ) = e3x and g( x ) = x. Both f and g are differentiable with f ′ ( x ) = 3e3x and g′ ( x ) = 1. We also
have limx→∞ f ( x ) = ∞ and limx→∞ g( x ) = ∞. Therefore by L’Hôpital’s rule we have
e3x 3e3x
lim = lim = lim 3e3x = ∞
x →∞ x x →∞ 1 x →∞
(c) Again this limit is of the form 00 . Therefore, L’Hôpital’s rule we have
1
sin θ − 2 cos θ
limπ = limπ
θ→ 6 θ − π6 θ→ 6 1
= limπ cos θ
θ→ 6
π
= cos
√ 6
3
=
2
∞
(d) This limit is of the form ∞. Therefore, L’Hôpital’s rule we have
x−2 ( x − 2) ′
lim = lim 2
x →2 2
x −4 x →2 ( x − 4 ) ′
1
= lim
x →2 2x
1
=
4
−∞
(e) This limit is of the form ∞ . Therefore, L’Hôpital’s rule we have
x − 8x2 ( x − 8x2 )′
lim = lim
2
x →∞ 12x + 5x x →∞ (12x 2 + 5x )′
1 − 16x
= lim
x →∞ 24x + 5
differentiation 87
−∞
This limit is still of the form ∞ . Therefore we apply L’Hôpital’s rule again.
(1 − 16x )′
= lim
x →∞ (24x + 5)′
−16
= lim
x →∞ 24
2
=−
3
∞
If the indeterminate form is of the form 00 , 0∞ , ∞0 and 1∞ , we first transform it into the form 0
0 or ∞
before applying L’Hôpital’s rule.
Example 5.9.
1
(a) lim (ln x ) x
x →∞
1
(b) lim(ex + x ) x
x →0
(c) lim x x
x →0+
Solution.
1
1. This limit is of the form ∞0 . So first we let f ( x ) = (ln x ) x . Next we let g( x ) = ln f ( x ). Therefore,
g( x ) = ln f ( x )
1
= ln(ln x ) x
1
= ln(ln x )
x
ln(ln x )
=
x
ln(ln x ) ∞
Now lim x is of the form ∞. Therefore
x →∞
ln(ln x )
lim g( x ) = lim
x →∞ x →∞ x
!
′
[ln(ln x )]
= lim Using L’Hôpital’s rule
x →∞ ( x )′
!
1
x
= lim 1
1
x →∞
ln x
!
1
x ln x
= lim
x →∞ 1
=0
f ( x ) = e g( x )
88 calculus
This implies that
Therefore
1
lim (ln x ) x = lim f ( x ) = lim eg(x) = e0 = 1
x →∞ x →∞ x →∞
1
2. This limit is of the form 1∞ . So first we let f ( x ) = (ex + x ) x . Next we let g( x ) = ln f ( x ). Therefore,
g( x ) = ln f ( x )
1
= (e x + x ) x
1
= ln(ex + x )
x
ln(ex + x )
=
x
ln(ex + x )
′ e x +1
Now lim x is of the form 00 . First we note that [ln(ex + x )] = ex + x . Therefore,
x →0
ln(ex + x )
lim g( x ) = lim
x →0 x →0 x
(e + x ) ′
x
= lim Using L’Hôpital’s rule
x →0 ( x )′
x !
e +1
ex + x
= lim
x →0 1
ex + 1
= lim
x →0 ex + x
=2
Therefore lim g( x ) = 2. We also know that
θ →0
f ( x ) = e g( x )
Therefore
1
lim(ln x ) x = lim f ( x ) = lim eg(x) = e2 = e2
x →0 x →0 x →0
g( x ) = ln f ( x )
= ln x x
= x ln x
ln x
= 1
x
differentiation 89
−∞
Now lim ln1 x is of the form ∞ .
x →0 x
!
ln x
lim( x ) = lim 1
g x →0
x
!
(ln x )′
= lim 1 ′
Using L’Hôpital’s rule
x →0
x
!
1
x
= lim
x →0 − x12
= lim(− x )
x →0
=0
f ( x ) = e g( x )
Therefore
1
lim(ln x ) x = lim f ( x ) = lim eg(x) = e0 = 1
x →0 x →0 x →0
6 Integration
The concept of integration is motivated by a goal to find area under the graph.
a b x
We can approximate this area by rectangles. Suppose we subdivide the interval into 4 interval as shown
below.
In this case we have 4 rectangles. Each rectangle has height given by the left limit and width is given
by diving [ a, b] into 4 parts. Thus each width is
b−a
∆=
4
a = x0 x1 x2 x3 b x
a = x0 x1 ··· x n −1 b x
Notice that as we increase the subdivisions we improve the accuracy of the approximate area. Thus we
exact area is obtained by letting n → ∞. This leads to the definition of a deifinite integral.
integration 93
Definition 6.1.
Let f be a function defined on a closed interval [ a, b] and let xi∗ ∈ [ xi−1 , xi ]. Then the definite integral of f from
Z b
a to b, denoted by f ( x ) dx, is defined by
a
Z b n
a
f ( x ) dx = lim
n→∞
∑ f (xi∗ ) · ∆x.
i =0
The numbers a and b are called the lower and upper limits of the definite integral, respectively. If the
limit exists, the function f is said to be integrable on the interval.
We will derive definite integrals for simple powersof x. To do this we list the following simple series
which can be proved by Principle of mathematical induction.
Exercise 6.1.
Use mathematical induction to prove that
n
n ( n +1)
1. ∑ k = 1 + 2 + · · · + n = 2
k =1
n
n(n+1)(2n+1)
2. ∑ k2 = 12 + 22 + · · · + n2 = 6
k =1
n
n2 ( n +1)2
3. ∑ k3 = 13 + 23 + · · · + n3 = 4
k =1
Example 6.1.
Evaluate the following integrals.
R1
1. 0
x dx
R5
2. 2
x2 dx
R1
3. 0
x3 dx
x0 = 0
1 1
x1 = x0 + =
n n
1 2
x2 = x1 + =
n n
.. ..
.= .
i
xi =
n
94 calculus
Therefore
i
f ( xi ) = xi =
n
1 n
b−a
Z
0
x dx = lim
n→∞
∑ f (xi∗ ) · n
i =1
n
1−0
= lim
n→∞
∑ f (xi∗ ) · n
i =1
n
i 1
= lim
n→∞
∑n·n
i =1
!
1 n
n2 i∑
= lim · i
n→∞
=1
1 n ( n + 1)
= lim
n→∞ n2 2
!
1 n2 (1 + n1 )
= lim
n→∞ n2 2
!
(1 + n1 )
= lim
n→∞ 2
1
=
2
x0 = 2
3 3
x1 = x0 + = 2+
n n
3 6
x2 = x1 + = 2 +
n n
.. ..
.= .
3i
xi = 2 +
n
Therefore
2
12i 9i2
3i
f ( xi ) = xi2 = 2+ = 4+ + 2
n n n
integration 95
2
x dx = lim
n→∞
∑ f (xi∗ ) · n
i =1
n
12i 9i2
3
= lim
n→∞
∑ 4+
n
+ 2 ·
n n
i =1
n n n
12 36i 27i2
= lim
n→∞
∑ n
+ lim ∑ 2 + lim ∑ 3
n→∞ n n→∞ n
i =1 i =1 i =1
n
12 36 n 27 n
= lim
n→∞
∑ n
+ lim 2 · ∑ i + lim 3 · ∑ i2
n→∞ n n→∞ n
i =1 i =1 i =1
12 36 n(n + 1) 27 n(n + 1)(2n + 1)
= lim n · + lim 2 · + lim 3 ·
nn→∞ n→∞ n 2 n→∞ n 6
2 1
36 n (1 + 2 ) 3 1
27 n (1 + n )(2 + n1 )
= 12 + lim 2 · + lim 3 ·
n→∞ n n n→∞ n 6
= 12 + 18 + 9
= 39
3. Exercise
Solution.
96 calculus
1.
Z b n
a
k dx = lim
n→∞
∑ k · ∆x
i =1
n
b−a
= k lim
n→∞
∑ n
i =1
b−a
= k·n·
n
= k(b − a)
2. In the case, ∆ = b− a
n = a− a
n = 0. Therefore,
Z a n n
a
f ( x ) dx = lim
n→∞
∑ f (xi∗ ) · ∆x = nlim
→∞
∑ f (xi∗ ) · 0 = 0.
i =1 i =1
3.
Z b n
a
f ( x ) dx = lim
n→∞
∑ f (xi∗ ) · ∆x
i =1
n
b−a
= lim
n→∞
∑ f (xi∗ ) · n
i =1
n
a−b
= − lim
n→∞
∑ f (xi∗ ) · n
i =1
Z a
=− f ( x ) dx
b
4.
Z b n
a
k f ( x ) dx = lim
n→∞
∑ k f (xi∗ ) · ∆x
i =1
n
b−a
= k lim
n→∞
∑ f (xi∗ ) · n
i =1
Z a
=k f ( x ) dx
b
5. We give proof for f ( x ) + g( x ) . The one for f ( x ) − g( x ) is similar.
Z b n
∑ f ( xi∗ ) + g( xi∗ ) · ∆x
f ( x ) + g( x ) dx = lim
a n→∞
i =1
n h i
= lim
n→∞
∑ f ( xi∗ ) · ∆x + g( xi∗ ) · ∆x
i =1
n n
= lim
n→∞
∑ f (xi∗ ) · ∆x + nlim
→∞
∑ g(xi∗ ) · ∆x
i =1 i =1
Z b Z b
= f ( x ) dx + g( x ) dx
a a
integration 97
6.
Z b n
a
f ( x ) dx = lim
n→∞
∑ f (xi∗ ) · ∆x
i =1
n
b−a
= lim
n→∞
∑ f (xi∗ ) · n
i =1
n
b−c+c−a
= lim
n→∞
∑ f (xi∗ ) · n
i =1
n
c−a b−c
= lim ∑ f ( xi ) ·
∗
+
n→∞ n n
i =1
n
c−a b−c
= lim ∑ f ( xi∗ ) · + f ( xi∗ ) ·
n→∞ n n
i =1
n n
c−a b−c
= lim
n→∞
∑ f (xi∗ ) · n
+ lim ∑ f ( xi∗ ) ·
n → ∞ n
i =1 i =1
Z c Z b
= f ( x ) dx + f ( x ) dx
a c
7.
Z b n
a
f ( x ) dx = lim
n→∞
∑ f (xi∗ ) · ∆x
i =1
f ( xi∗ ) · ∆x ≥ 0
Therefore
n
∑ f (xi∗ ) · ∆x ≥ 0
i =1
Hence
Z b n
a
f ( x ) dx = lim
n→∞
∑ f (xi∗ ) · ∆x
i =1
≥0
The independent variable x in a definite integral is called a dummy variable of integration. The value
of the integral does not depend on the symbol used. In other words,
Z b Z b Z b Z b
f ( x ) dx = f (r ) dr = f (s) ds = f (t) dt
a a a a
and so on.
6.2 Anti-derivatives
That is, for a given function f , we wish to find another function F for which F ′ ( x ) = f ( x ) for all x on
some interval.
Definition 6.2.
A function F is said to be an anti-derivative of a function f if F ′ ( x ) = f ( x ) on some interval.
There is always more than one anti-derivative of a function. For instance, in the foregoing example,
F1 ( x ) = x2 − 1 and F2 ( x ) = x2 + 10 are also anti-derivatives of f ( x ) = 2x since F1′ ( x ) = F2′ ( x ) = f ( x ).
Indeed, if F is an anti-derivative of a function f , then so is G ( x ) = F ( x ) + C, for any constant C. This
is a consequence of the fact that
d
G′ ( x) = ( F ( x ) + C ) = F ′ ( x ) + 0 = F ′ ( x ) = f ( x ).
dx
Thus, F ( x ) + C stands for a set of functions of which each member has a derivative equal to f ( x ).
Theorem 6.2.
If G ′ ( x ) = F ′ ( x ) for all x in some interval [ a, b], then
G(x) = F(x) + C
The anti-derivative of a fucntion f is called the indefinite integral of f with respect to x. The function
f is called the integrand. The process of finding an anti-derivative is called anti-differentiation or
integration. The number C is called a constant of integration.
Z Z
1. c f ( x ) dx = c f ( x ) dx.
Z Z Z
2. [ f ( x ) ± g( x )] = f ( x ) dx ± g( x ) dx.
integration 99
1
Z
3. x n dx = x n+1 + C for n ̸= −1.
n+1
Z
4. a dx = ax + C.
Z
5. cos x dx = sin x + C.
Z
6. sin x dx = − cos x + C.
Z
7. sec2 x dx = tan x + C.
Z
8. e x dx = e x + C.
1
Z
9. dx = ln | x | + C.
x
1
Z
10. √ dx = sin−1 x + C.
1 − x2
1
Z
11. dx = tan−1 x + C.
1 + x2
The fundamental theorem of calculus provides the connection between the differential calculus and
the integral calculus.
Theorem 6.3.
Let f be continuous on [ a, b] and let F be any function for which F ′ ( x ) = f ( x ) (i.e. F is the antiderivative of f ).
Then Z b
f ( x ) dx = F (b) − F ( a). (6.2)
a
Thus to find the definite integral of f we first find its anti-derivative and substitute the endpoints.
Example 6.2.
Z 1
Evaluate x dx.
0
Solution.
The anti-derivative of f ( x ) = x is given by
x2
F(x) = +C
2
Consequently,
F (0) = 0 + C = C
100 calculus
and
12 1
F (1) = +C = +C
2 2
Therefore by the fundamental Theorem of Calculus we have
Z 1
x dx = F (1) − F (0)
0
1
= +C −C
2
1
=
2
Example 6.3.
Z 5
Evaluate x2 dx.
2
Solution.
The anti-derivative of f ( x ) = x2 is given by
x3
F(x) = +C
3
Therefore by the fundamental Theorem of Calculus we have
5 x3 5
Z
x dx =
2 3 2
53 23
= −
3 3
125 8
= −
3 3
= 39
1. Integration by parts
2. Substitution
integration 101
4. Reduction formulae
5. etc. etc.
7 Introduction to functions of several variables
So far we have dealt only with functions of single (independent) variables. Many familiar quantities,
however, are functions of two or more variables. For instance, the work done by a force (W = FD) and
the volume of a right circular cylinder (V = πr2 h) are both functions of two variables. The volume of a
rectangular solid (V = lwh) is a function of three variables. The notation for a function of two or three
variables is as follows
z = f ( x, y) = x2 + xy
| {z }
2 variables
and
w = f ( x, y, z) = x + 2y − 3z.
| {z }
3 variables
Let D be a set of ordered pairs of real numbers. If to each ordered pair ( x, y) in D there corresponds a
unique real number f ( x, y), then f is called a function of x and y. The set D is the domain of f and
the corresponding set of values for f ( x, y) is the range of f . For the function given by z = f ( x, y), we
call x and y the independent variables and z the dependent variable.
As with functions of one variable, the most common way to describe a function of several variables is
with an equation, and unless otherwise restricted, we can assume that the domain is the set of all points
for which the equation is defined. for example, the domain of the function given by
f ( x, y) = x2 + y2
Example 7.1.
Determine the domains of the following functions.
p
x 2 + y2 − 9
1. f ( x, y) =
x
104 calculus
x
2. g( x, y, z) = p .
9 − x 2 − y2 − z2
In the application of functions of several variables, the question often arises, “How will a function
be affected by a change in one of its independent variables?”. You can answer by considering the
independent variables one at a time. The process is called partial differentiation, and the result is
referred to as the partial derivative of f with respect to the chosen independent variable 1 . 1
The intro-
duction of
partial deriva-
tives followed
Newton’s and
7.2.2 Definition of Partial Derivatives of a Function of Two Variables Leibniz’s work
in calculus
by several
Definition 7.1. years. Between
1760, Leon-
If z = f ( x, y), then the first partial derivatives of f with respect to x and y denoted by f x and f y are the hard Euler and
functions defined by Jean Le Rond
d’Alembert
f ( x + h, y) − f ( x, y) (1717-1783)
f x ( x, y) = lim (7.1) separately
h →0 h
published sev-
f ( x, y + h) − f ( x, y) eral papers on
f y ( x, y) = lim (7.2)
h →0 h dynamics, in
which they es-
provided the limits exist. tablished much
of the the-
ory of partial
derivatives
∂ ∂z
f ( x, y) = f x ( x, y) = z x =
∂x ∂x
and
∂ ∂z
f ( x, y) = f y ( x, y) = zy =
∂y ∂y
∂z
= f x ( a, b)
∂x
( a,b)
introduction to functions of several variables 105
and
∂z
= f y ( a, b)
∂y
( a,b)
∂f
The notation ∂x is read “partial f – partial x”.
Example 7.2.
Let f be defined by f ( x, y) = x2 + 3xy + 2y2 . Evaluate
∂f
Example 7.3.1. ∂x
∂f
2. ∂y
∂f
3. ∂y
(5,−2)
Solution.1.
f ( x + h, y) − f ( x, y)
∂f
= lim
∂x h →0 h
[( x + h)2 + 3( x + h)y + 2y2 ] − [ x2 + 3xy + 2y2 ]
= lim
h →0 h
2
x + 2xh + h + 3xy + 3hy + 2y2 − x2 − 3xy − 2y2
2
= lim
h →0 h
2
2xh + h + 3hy
= lim
h →0 h
= lim(2x + 3y + h)
h →0
= 2x + 3y
2.
f ( x, y + h) − f ( x, y)
∂f
= lim
∂y h →0 h
[ x2 + 3x (y + h) + 2(y + h)2 ] − [ x2 + 3xy + 2y2 ]
= lim
h →0 h
2
x + 3xy + 3xh + 2(y2 + 2yh + h2 )2 − x2 − 3xy − 2y2
= lim
h →0 h
3xh + 4hy + 2h2
= lim
h →0 h
= lim(3x + 4y + 2h)
h →0
= 3x + 4y
3.
∂f
= 3(5) + 4(−2) = 15 − 8 = 7
∂y
(2,3)
106 calculus
Let’s consider the function f ( x, y) = x2 + 3xy + 2y2 again. Suppose we treat y as constant and find the
derivative of f with respect to x we get the derivative as 2x + 3y + 0. Which is similar to the partial
derivative of f with respect to x. Thus we arrive at the following
Remark 7.1.
Consider the function of 2 variable, f ,
f x ( x, y) = 3 − 2xy2 + 6x2 y.
f y ( x, y) = −2x2 y + 2x3 .
We can use this approach for all types of derivatives using all the techniques and properties we have
learned from differentiation.
Example 7.5.
Evaluate the following
ln(3xy2 + x2 )
∂
1. ∂x
The concept of a partial derivative can be extended naturally to functions of three or more variables.
For instance, if w = f ( x, y, z), then there are three partial derivatives, each of which is formed by
holding two of three variables constant.
∂w f ( x + h, y, z) − f ( x, y, z)
= f x ( x, y, z) = lim
∂x h →0 h
∂w f ( x, y + h, z) − f ( x, y, z)
= f y ( x, y, z) = lim
∂y h →0 h
∂w f ( x, y, z + h) − f ( x, y, z)
= f z ( x, y, z) = lim .
∂z h →0 h
introduction to functions of several variables 107
Example 7.6.
Let f ( x, y, z) = xy + yz2 + xz. Evaluate
1. f x
2. f y
3. f z
Solution.
1. To find the partial derivative of f with respect to x, we keep y and z constant to obtain
∂
xy + yz2 + xz = y + z
∂x
2. To find the partial derivative of f with respect to y, we keep x and z constant to obtain
∂
xy + yz2 + xz = x + z2 .
∂y
3. To find the partial derivative of f with respect to z, consider x and y to be constant and obtain
∂
xy + yz2 + xz = 2yz + x.
∂z
It is possible to take second, third and higher partial derivatives of a function of several variables,
provided such derivatives exist. Higher-order derivatives are denoted by the order in which the differ-
entiation occurs. For instance, the function z = f ( x, y) has the following second partial derivatives.
∂2 f
∂ ∂f
= = f xx .
∂x ∂x ∂x2
∂2 f
∂ ∂f
= = f yy .
∂y ∂y ∂y2
∂2 f
∂ ∂f
= = f xy .
∂y ∂x ∂y∂x
108 calculus
4. Differentiate first with respect to y and then with respect to x:
∂2 f
∂ ∂f
= = f yx .
∂x ∂y ∂x∂y
The third and fourth cases are called mixed partial derivatives.
Example 7.7.
Find the second partial derivatives of f ( x, y) = 3xy2 − 2y + 5x2 y2 and determine the value of f xy (−1, 2).
Solution.
We begin by finding the first partial derivatives with respect to x and y.
f x ( x, y) = 3y2 + 10xy2
and
f y ( x, y) = 6xy − 2 + 10x2 y
Notice that the two mixed partials are equal. Sufficient conditions for this occurrence are given in the
next theorem.
f xy ( a, b) = f yx ( a, b).
Example 7.8.
Show that f xz = f zx and f xzz = f zxz = f zzx for the function given by
f ( x, y, z) = ye x + x ln z.
Solution.
First partials:
x
f x ( x, y, z) = ye x + ln z, f z ( x, y, z) = .
z
introduction to functions of several variables 109
Recall the study of differential equations from high school. Differential equations of one variable are
called ordinary differential equations. Examples of ordinary differential equation are given below
If a physical phenomena depends on two variables then differential equations governing behaviour
of that phenomena are called partial differential equations (PDEs). Some examples of PDEs are given
below.
1. [Heat equation]
Let u( x, t) be the temperature of a rod at position x at time t. Then u is a function of two variables
x and the time t.
∂u ∂2 u
= c2 2
∂t ∂t
where c is a constant.
2. [Wave equation]
Let u( x, t) denote the vertical displacement experienced by the string at the point x at time t. If
damping effects, such as air resistance, are neglected, and if the amplitude of the motion is not too
large, then u( x, t) satisfies the partial differential equation
∂2 u ∂2 u
2
= c2 2
∂t ∂t
110 calculus
where c is a constant.
3. [Black-Scholes pde]
Then the price of a European call option V (t, S) on a non-dividend paying stock is a function of
stock price S and time t. V given by
∂V 1 ∂2 V ∂V
+ σ2 S2 2 + rS − rV = 0
∂t 2 ∂S ∂S
where r is the risk-free interest rate, and σ is the volatility of the stock.
4. [Laplace Equation]
Let u be a function of x and y. The following a pde has many applications in electricity and fluid
flow.
∂2 u ∂2 u
+ 2 =0
∂x2 ∂y
Example 7.9.
Determine which of the following are solution of the Laplace equation.
1. u( x, y) = ex sin y
2. u( x, y) = x2 − y2
p
3. u( x, y) = − ln x2 + y2