Advanced Calculus Study Guide 2025
Advanced Calculus Study Guide 2025
SMTA 021
University of Limpopo (Turfloop Campus)
DR L. Rundora
School of Mathematical and Computer Sciences
Department of Mathematics and Applied Mathematics
University of Limpopo (Turfloop Campus)
Contents
1 Limits: A Review 1
1.1 What this Unit is all About . . . . . . . . . . . . . . . . . . . . . 1
1.2 Limits: A Review . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
3 Sequences 14
3.1 What this Unit is all About . . . . . . . . . . . . . . . . . . . . . 14
3.2 Properties of convergent sequences . . . . . . . . . . . . . . . . . 16
3.2.1 Boundedness . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.3 The Bolzano-Weierstrass Theorem . . . . . . . . . . . . . . . . . . 25
5 Power Series 48
5.1 What this Unit is all About . . . . . . . . . . . . . . . . . . . . . 48
5.2 Power Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.3 Taylor and Maclaurin Series . . . . . . . . . . . . . . . . . . . . . 51
i
6.2.1 Characterization of limits of functions in terms of conver-
gence of sequences . . . . . . . . . . . . . . . . . . . . . . 56
6.3 Continuous functions . . . . . . . . . . . . . . . . . . . . . . . . . 57
6.3.1 Some properties of continuous functions . . . . . . . . . . 59
8 Integration 81
8.1 What this Unit is all About . . . . . . . . . . . . . . . . . . . . . 81
8.2 The Riemann Integral . . . . . . . . . . . . . . . . . . . . . . . . 81
8.2.1 Riemann Integrable Functions . . . . . . . . . . . . . . . . 84
8.2.2 Properties of the Riemann integral . . . . . . . . . . . . . 85
8.2.3 The fundamental theorem of integral calculus . . . . . . . 87
8.3 Multiple integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
8.3.1 The Double integral . . . . . . . . . . . . . . . . . . . . . 88
8.3.2 Triple Integrals . . . . . . . . . . . . . . . . . . . . . . . . 89
8.3.3 Transformations of multiple integrals . . . . . . . . . . . . 92
8.4 Line Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
8.5 Surface integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
8.5.1 Gradient, Divergence and Curl . . . . . . . . . . . . . . . . 100
ii
9.3.2 Equations reducible to variables separable type . . . . . . 112
9.3.3 Homogeneous equations . . . . . . . . . . . . . . . . . . . 114
9.3.4 Linear differential equations . . . . . . . . . . . . . . . . . 117
9.4 Second order linear differential equations with constant coefficients 120
9.4.1 Homogeneous equations of first order . . . . . . . . . . . . 120
9.4.2 Homogeneous equations of second order . . . . . . . . . . . 120
iii
Preface
In this module, we study sequences, infinite series, power series, limits and con-
tinuity, partial derivatives, Riemann integrals, multiple integrals, Line integrals
and surface integrals. These major areas are subdivided into units and subunits.
We start off with a review of limits and limits of indeterminate forms. These
first 2 units are included only for revision purposes and will not be discussed in
class. They were covered in SMTH011. The last unit is on ordinary differential
equations. It is most unlikely that we will have time to discuss this last unit in
class.
In each unit we incorporate several worked examples, individual class activities
and tutorial activities. A thorough study of the worked examples should enable
a student to work through the tutorial questions with little difficulty. Students
are encouraged to make sure that they read and understand their lectures before
attempting tutorial questions. Students are further encouraged to make use of
their lecturer should they have any queries arising from their studies.
It should be noted that the learning of Mathematics, unlike other subjects, is in
the ’doing of the Mathematics’ rather than just reading through.
This study guide, as the name suggests, is just a guide through the lectures. It is
still important to attend all the lectures. Again, the study guide is not meant to
replace any textbook. Thus, the study guide should be used in conjunction with
the relevant prescribed textbooks. A habit of visiting the library is encouraged.
It is my hope that you will enjoy going through this module.
iv
Recommended Textbooks
1. Murray R. Spiegel, ”Advanced Calculus, Schaum’s Outline Series”, McGraw-
Hill, 1981.
2. Murray R. Spiegel, ”Advanced Mathematics for Engineers and Scientists”,
McGraw-Hill, 1971.
3. Buck R.C, ”Advanced Calculus”, McGraw-Hill.
4. Finney Thomas, ”Calculus”, 2nd edition, Addison Wesley.
5. Burkill J.C., ”A first Course in Mathematical Analysis”, Cambridge Univer-
sity Press,1970.
6. Rudin W., ”Principles of Mathematical Analysis”, McGraw-Hill, 1974.
7. Leonard F. Richardson, ”ADVANCED CALCULUS: An Introduction to Lin-
ear Analysis”, John Wiley & Sons, Inc., New Jersey, 2008.
v
Unit 1
Limits: A Review
Notation:
Write l = lim f (x) or f (x) −→ l as x −→ x0 .
x−→x0
Remark 1.1 If lim f (x) = l1 and lim f (x) = l2 , then we call l1 and l2
x−→x+
0 x−→x−
0
respectively the right hand limit and the left hand limit of f (x) at x0 . In this way,
1
2 Limits: A Review
we say that lim f (x) exists if lim + f (x) = lim− f (x), i.e., if l1 = l2 .
x−→x0 x−→x0 x−→x0
Solution:
Use the graph of f (x) = |x| to observe that lim+ |x| = 0 = lim |x|.
x−→0 x−→0−
Therefore lim |x| = 0.
x−→0
|x|
2. If f (x) = , x 6= 0, what is lim f (x)?
x x−→0
Solution:
x, if x > 0, 1, if x > 0,
Recall that |x| = Therefore f (x) =
−x, if x < 0 −1, if x < 0
Now, use a graph to observe that lim− f (x) = −1 and lim+ f (x) = 1. So
x−→0 x−→0
|x|
we have that lim− f (x) 6= lim+ f (x). Therefore lim , x 6= 0 does not
x−→0 x−→0 x−→0 x
exist.
3 − x, if x < 2,
3. f (x) = 2, if x = 2 Does lim f (x) exist?
x x−→2
2
, if x>2
Solution
Exercise!!
4. Prove that
(a) lim (7x − 1) = 13.
x−→2
x2 − 4
(b) lim = 4, x 6= 2
x−→2 x − 2
Solution:
(a). Here f (x) = 7x − 1, x0 = 2 and l = 13. So, given ε > 0, we need to
find δ(ε) > 0 so that ∀ x, 0 < |x − 2| < δ =⇒ |f (x) − 13| < ε. Now,
So 0 < |x − 2| < δ =⇒
ε
where we have chosen δ = . So 0 < |x − 2| < δ =⇒ |f (x) − 13| < ε, i.e.
7
lim (7x − 1) = 13.
x−→2
x2 −4
(b). Let ε > 0 be given. We have that x−2
− 4 = |x + 2 − 4| = |x − 2|,
x2 −4
so that if 0 < |x − 2| < δ, then x−2
− 4 = |x + 2 − 4| = |x − 2| < δ. So
x2 − 4
we choose δ = ε to conclude that lim = 4, x 6= 2.
x−→2 x − 2
Solution
Here f (x) = x2 , l = 4 and x0 = 2. Now,
The desired factor is |x−2|. We need to bound |x+2|. For the factor |x+2|,
if x is close to 2, x + 2 is close to 4. So we should be able to replace |x + 2|
by a constant close to 4. More precisely, we observe that if |x − 2| < 1, then
|x + 2| < 5.
Thus, we have
so that given ε > 0, let δ = min 1, 5ε . This ensures that 0 < |x−2| < δ =⇒
1 1
6. Prove that lim = .
x−→2 x + 1 3
Solution
1 1
Here f (x) = , x0 = 2 and l = . So
x+1 3
1 1 3 − (x + 1) 1 2−x 1 |x − 2|
|f (x) − l| = − = = = .
x+1 3 3(x + 1) 3 x+1 3 |x + 1|
1 |x − 2| 1 |x − 2| 1
|f (x) − l| = < = |x − 2|. Therefore, given ε > 0,
3 |x + 1| 3 2 6
we can take δ = min{6ε, 1} to ensure that if 0 < |x − 2| < δ we have
1 1 |x − 2| 1 1
x+1
− 31 = 1
< |x − 2| ≤ · 6ε = ε. So x+1 − 13 < ε whenever
3 |x + 1| 6 6
1 1
0 < |x − 2| < δ and so lim = .
x−→2 x + 1 3
lim f (x)
f (x) x−→x0 l
3. lim = = , m 6= 0.
x−→x0 g(x) lim g(x) m
x−→x0
1 − cos x 0
Example 2.1 1. lim 2
takes the form but can be solved as:
x−→0 x 0
1 − cos x 1 − (1 − 2 sin2 x2 )
lim = lim
x−→0 x2 x−→0 x2
2 x
2 sin 2
= lim
x−→0 x2
sin2 x2
= 2 lim
x−→0 x2
2 sin2 x2
= lim
4 x−→0 ( x2 )2
1
=
2
6
Limits of Indeterminate Type 7
x2 − 9 0
2. lim again takes the form but can be solved as:
x−→3 x − 3 0
x2 − 9 (x − 3)(x + 3)
lim = lim = lim x + 3 = 6.
x−→3 x − 3 x−→3 x−3 x−→3
Problem!! These (such) techniques do not work for some types of functions.
For such functions, we will need differential techniques in order to evaluate the
associated limits. This is done via L’Hospital’s Rules, which we are now set
to explore.
Recall that
f (x) − f (c)
Proof 2.1 f 0 (c) = lim > 0 =⇒ f (x) > f (c) for x > c and f (x) <
x−→c x−c
f (c) for x < c. ♠
f (b), then f (a) and f (b) are less than M and ∃ a point c ∈ (a, b) such that
f (c) = M . Since f (c) is the least upper bound of the set f ([a, b]), it implies that
f (x) ≤ f (c) ∀ x ∈ [a, b]. This means that f has a local maximum at c. Hence
f 0 (c) = 0.
Similarly, if m 6= f (a) = f (b), then f (a) and f (b) are greater than m and ∃ a
point c ∈ (a, b) such that f (c) = m. Since f (c) is the greatest lower bound of the
set f ([a, b]), it implies that f (x) ≥ f (c) ∀ x ∈ [a, b]. This means that f has a
local minimum at c. Hence f 0 (c) = 0. ♠
Remark 2.1 If g is chosen to be the identity function g(x) = x, then the Gen-
eralized Mean Value Theorem becomes the Mean Value Theorem above.
Remark 2.2 Suppose g(b) − g(a) = 0. Then we cannot solve for k. However,
we do not need to since we see that the theorem will hold if we can find a c with
g 0 (c) = 0 so that both sides will be 0. This we can do by applying Rolle’s theorem
to g.
Limits of Indeterminate Type 9
Proof 2.4 Suppose f and g are differentiable in (c, x) and g 0 6= 0 in (c, x). It
follows from the Generalized Mean Value Theorem that there is a point y ∈ (c, x)
f 0 (y) f (x) − f (c) f 0 (y)
such that 0 = . Since f (c) = g(c) = 0, this reduces to 0 =
g (y) g(x) − g(c) g (y)
f (x)
. Since y ∈ (c, x), x −→ c implies y −→ c which in turn ensures that
g(x)
f (x) f 0 (y)
lim = lim 0 = l which proves. ♠
x−→c g(x) x−→c g (y)
Case 2
Suppose, instead, that x −→ ∞. We show that the same result as in Case 1 is
1
obtained. We introduce the substitution x(t) = and apply the chain rule of
t
differentiation. We see that as x −→ ∞, t −→ 0+ . Now,
f (x) f ( 1t )
lim = lim+ 1
x−→∞ g(x) t−→0 g( )
t
(f ox)(t)
= lim+
t−→0 (gox)(t)
(f ox)0 (t)
= lim+
t−→0 (gox)0 (t)
f 0 (x)x0 (t)
= lim+ 0
t−→0 g (x)x0 (t)
f 0 ( 1t )
− t2
= lim g 0 ( 1t )
t−→0+
− t2
f 0 ( 1t )
= lim
t−→0+ g 0 ( 1t )
0
f (x)
= lim
x−→∞ g 0 (x)
= l
L’Hospital’s Rule 2
Suppose f and g are functions differentiable in (a, b) with
10 Limits of Indeterminate Type
g 0 (x) 6= 0 ∀ x ∈ (a, b). Let c ∈ (a, b). If lim f (x) = lim g(x) = ∞ but
x−→c x−→c
f 0 (x) f (x)
lim 0 = l. Then lim = l.
x−→c g (x) x−→c g(x)
Proof 2.5 We shall prove the theorem for the case x → +∞.
From lim f (x) = lim g(x) = ∞, pick a sufficiently large real number y such that
x−→c x−→c
the functions f, g and g 0 are non-degenerate on (y, ∞). Then for any x ≥ y, the
Generalized Mean Value Theorem ensures the existence of a point λ ∈ (y, x) such
that " f (y) #
0
f (λ) f (x) − f (y) f (x) 1 − f (x)
0
= = .
g (λ) g(x) − g(y) g(x) 1 − g(y)
g(x)
But then this is so since we assumed that lim f (x) = lim g(x) = ∞ and y is
x−→c x−→c
f (x) f 0 (λ)
fixed. Hence lim = lim 0 = l. ♠
x−→c g(x) λ−→c g (λ)
sin x cos x
Example 2.2 1. lim = lim = 1.
x−→0 x x−→0 1
1 − cos x sin x cos x 1
2. lim 2
= lim = lim = .
x−→0 x x−→0 2x x−→0 2 2
3.
xn nxn−1
lim = lim
x−→∞ ex x−→∞ ex
n(n − 1)xn−2
= lim
x−→∞ ex
n(n − 1)(n − 2)xn−3
= lim
x−→∞ ex
Limits of Indeterminate Type 11
.
= ..
n!
= lim
x−→∞ ex
= 0
x2 − 9 2x
4. lim = lim = 6.
x−→3 x − 3 x−→3 1
ln g(x)
ln h(x) = f (x) ln g(x) = 1 .
f (x)
So " #
ln g(x)
lim ln h(x) = lim 1
x−→c x−→c
f (x)
∞
which is a limit of the type . So L’Hospital’s Rule 2 is applicable. Consequently,
∞
" #
ln g(x)
lim 1
x−→c
lim h(x) = e f (x) .
x−→c
ln x
y = xx =⇒ ln y = x ln x = 1
x
and
1
ln x x
lim+ ln y = lim+ 1 = lim+ = lim+ (−x) = 0.
x−→0 x−→0
x
x−→0 − x12 x−→0
So
lim ln y
lim+ xx = lim+ eln y = ex−→0+ = e0 = 1.
x−→0 x−→0
12 Limits of Indeterminate Type
1
2. Find lim (cos x) x2
x−→0
Solution
1
Since lim cos x = 1 and lim 2 = ∞, the limit in question takes the inde-
x−→0 x−→0 x
1
terminate form 1 . Now, let y = (cos x) x2 . Then ln y = ln(cos
∞
x2
x)
. We now
apply L’Hospital’s rule to this. We have
ln cos x
lim ln y = lim
x−→0 x−→0 x2
sin x
−
= lim cos x
x−→0 2x
− sin x
= lim
x−→0 2x cos x
− cos x
= lim
x−→0 −2x sin x + 2 cos x
= − 12
1 1
Hence lim (cos x) x2 = e− 2 .
x−→0
x
1
3. Find lim 1 + .
x−→∞ x
Solution x
Observe that we have a limit of the form 1∞ . Let y = 1 + x1 . Then
1
ln y = x ln 1 +
x
ln 1 + x1
= 1
x
ln 1+x
x
= 1
x
ln 1+x
x
∴ lim ln y = lim 1
x−→∞ x−→∞
x
x −1
x+1 x2
= lim −1
x−→∞
x2
x
= lim
x−→∞ x + 1
= 1
x
1
∴ lim 1+ = e.
x−→∞ x
Activity 2.2 1. For each of the following functions find, if possible, a number
η satisfying Rolle’s theorem in the given interval. If not possible explain why
Limits of Indeterminate Type 13
3. Suppose the functions f and g are differentiable with f (a) = g(a) and
f 0 (x) < g 0 (x) ∀ x > a. Show that f (x) < g(x) ∀ x > a.
[Hint: Let h(x) = f (x) − g(x) and apply the Mean Value theorem.]
4. Show that ln(1+x) < x ∀ x > 0. Apply the exercise above with appropriate
f (x) and g(x).
3. 0, 1, 0, 21 , 0, 31 , 0, 14 , · · ·
4. 2, 32 , 43 , 54 , 65 , · · ·
One can deduce a formula or an algorithm or a defining rule for any given se-
quence. In the above example
1. f (n) = un = 2n−1
14
Sequences 15
2. Exercise
3. Exercise
n+1
4. f (n) = un = .
n
We can also define sequences by giving a relation or a formula that connects
successive terms of the sequence and specifying the value or values of the first
term or the first and the second terms etc. The formula or relation linking the
terms is called a recursive formula or a recurrence relation. Thus, in this case the
sequence is said to be defined recursively.
Example 3.2
2
un = 1 + , u1 = 1
un−1
un = un−1 + un−2 , n ≥ 3, u1 = 1, u2 = 1
3
un+1 = un + 1, u1 = 1
4
Activity 3.1 Obtain the first five terms for each of the sequences defined above.
Write lim un = l or un −→ l as n −→ ∞.
n−→∞
4 − 2n
Example 3.3 1. Find the number l to which the sequence {un } =
3n + 2
converges and prove, using the definition, that this number is indeed the
limit.
Solution
2
Here, it is clear that l = − . To prove this: Given ε > 0, we look for
3
16 Sequences
12 − 6n + 6n + 4
=⇒ < ε
9n + 6
16
=⇒ < ε
9n + 6
16
=⇒ < ε
9n + 6
=⇒ 16 < 9nε + 6ε
16 − 6ε
=⇒ < n
9ε
16 − 6ε
So N = will do. e.g. if ε = 0.1, N = 18. What is N if ε = 0.001?
9ε
2n − 7
2. Let {un } = Find how large n must be so that
3n + 2
(a) |un − l| < 0.1
(b) |un − l| < 0.001
(c) |un − l| < ε.
Solution
If x = 0, then clearly lim xn = 0. Now, assume that x 6= 0. Then we
n−→∞
must show that given ε > 0 ∃ N ∈ N such that |xn − 0| < ε ∀ n > N .
log ε
Now, |xn − 0| < ε =⇒ |xn | < ε =⇒ n log |x| < log ε =⇒ n > since
log |x|
log ε
log |x| < 0 as |x| < 1. Thus, if N is the smallest integer greater than
log |x|
then |xn − 0| < ε ∀ n > N and so lim xn = 0.
n−→∞
|l − s| = |l − un + un − s|
≤ |l − un | + |un − s| by triangle inequality
ε ε
< + by definitions above
2 2
= ε
∴ |l − s| < ε This implies that l = s for all n > N = max{N1 , N2 }. i.e. l and
s differ by at most ε which is arbitrary, i.e., can be chosen to be as small as we
like. ♠
3.2.1 Boundedness
Definition 3.3 A sequence {un } is said to be bounded above if ∃ a number α
such that un ≤ α ∀ n.
Remark 3.2 • The number α in the definition above is often called an upper
bound for the sequence {un }.
• Whenever {un } is bounded above, there are uncountably many upper bounds to
the sequence.
• If {un } is bounded above, then the graph of the sequence would show, the points
of the sequence all lying below a specific horizontal line.
Remark 3.3 • The number β in the definition above is often called a lower
bound for the sequence {un }.
• Whenever {un } is bounded below, there are uncountably many lower bounds to
the sequence.
• If {un } is bounded below, then the graph of the sequence would show, the points
of the sequence all lying above a specific horizontal line.
18 Sequences
Definition 3.6 Let {un } be a sequence. Then a number α is said to be the the
least upper bound or supremum or simply lub of {un } (a) if α is an upper
bound of {un }; and (b) if η is an upper bound of {un }, then α ≤ η.
Definition 3.7 Let {un } be a sequence. Then a number β is said to be the the
greatest lower bound or infinimum or simply glb of {un } (a) if β is a lower
bound of {un }; and (b) if γ is a lower bound of {un }, then γ ≤ β.
Theorem 3.3 Suppose {un } and {vn } are sequences of real numbers such that
lim un = u and lim vn = v. Then
n−→∞ n−→∞
1. lim {un ± vn } = u ± v.
n−→∞
Proof 3.3 1.
ε
{un } −→ u =⇒ ∀ ε > 0 ∃ N1 (ε) : |un − u| <
2
for all n > N1 and
ε
{vn } −→ v =⇒ ∀ ε > 0 ∃ N2 (ε) : |vn − v| <
2
Sequences 19
for all n > N2 . Now, choosing N : N = max{N1 , N2 } we have that for all
n>N
=⇒ un + vn −→ u + v. ♠
2.
{un } −→ u =⇒ ∀ ε > 0 ∃ N (ε) : |un − u| < ε
for all n > N . Now
3. We need to show that ∀ ε > 0 ∃ N ∈ N such that |un vn −uv| < ε ∀ n > N .
Now,
by Theorem 3.2.
Since un −→ u and vn −→ v, then ∀ ε > 0 we can find N1 and N2 such
that
ε
|un − u| < ∀ n > N1
2(|v| + 1)
and
ε
|vn − v| < ∀ n > N2
2M
Hence, from (3.2)we have
ε ε
|un vn − uv| < M · + (|v| + 1) ·
2M 2(|v| + 1)
ε ε
= +
2 2
= ε
1 1
Lemma 3.1 Suppose un −→ u where u 6= 0, then −→ .
un u
20 Sequences
Remark 3.4 The results in the previous theorem can be extended by Mathemat-
ical Induction to a finite number of convergent sequences.
• If an , bn , cn , · · · , zn are convergent sequences, then their sum (an + bn + · · · + zn )
is also a convergent sequence and lim(an + bn + · · · + zn ) = lim an + lim bn +
lim cn + · · · + lim zn .
• Their product an bn · · · zn is also a convergent sequence and lim(an bn · · · zn ) =
(lim an )(lim bn )(lim cn ) · · · (lim zn ).
• Hence, if k ∈ N and if {un } is a convergent sequence then lim(un )k = (lim un )k .
Theorem 3.4 and Theorem 3.5 taken together give the following theorem which is
referred to as ”The Principle of Monotone Bounded Convergence (PMBC)”:
n
Example 3.6 We want to show that the sequence an = 1 + n1 is monotonic
increasing.
n
1
an = 1+
n
2 n
1 n(n − 1) 1 1
= 1+n + + ··· +
n 2! n n
1 1 1 1 2 r−1
= 1+1+ 1− + ··· + 1− 1− ··· 1 −
2! n r! n n n
1 1 n−1
+ ··· + 1− ··· 1 −
n! n n
n+1
1
an+1 = 1+
n+1
1 1 1 1 2 r−1
= 1+1+ 1− + ··· + 1− 1− ··· 1 −
2! n+1 r! n+1 n+1 n+1
1 1 n−1 1 1 n
+ ··· + 1− ··· 1 − + 1− ··· 1 −
n! n+1 n+1 (n + 1)! n+1 n+1
Observe that the first two terms of an and an+1 are the same: the rth term of
an+1 is greater than the rth term of an for 3 ≤ r ≤ n + 1. Also, an+1 has an
1 1 n
extra positive term (n+1)! 1 − n+1 · · · 1 − n+1 at the end. Hence an < an+1
1
n
and an = 1 + n is a strictly increasing sequence.
NB: We have already (in the tutorial) shown that an −→ e.
2 < an < 3 ∀ n.
5
Proof 3.5 a1 = , and so the result is true for a1 . Now, suppose the result is
2
true for n = k − 1, where k is some integer such that k ≥ 2. Then
We have
5ak = a2k−1 + 6 < 32 + 6 = 15
by above. Also
5ak = a2k−1 + 6 > 22 + 6 = 10
again by above. Hence 10 < 5ak < 15 =⇒ 2 < ak < 3.
Sequences 23
1 2
an+1 − an = (a + 6) − an
5 n
1 2
= (a − 5an + 6)
5 n
1
= (an − 3)(an − 2)
5
< 0
and
|vn − l| < ε ∀ n > N.
This implies that
−ε < un − l < ε ∀ n > N
and
−ε < vn − l < ε ∀ n > N.
Now, un ≤ an ≤ vn ∀ n ∈ N =⇒
un − l ≤ an − l ≤ vn − l ∀ n > N.
sin n
Example 3.8 To show that lim = 0, we observe that
n−→∞ n
−1 ≤ sin n ≤ 1 ∀ n ∈ N.
This implies
−1 sin n 1
≤ ≤ ∀ n ∈ N.
n n n
−1 1 sin n
Since lim = 0 = lim we conclude that lim = 0 by the squeeze
n−→∞ n n−→∞ n n−→∞ n
theorem.
1
Activity 3.4 Show that lim (2n + 3n ) n = 3.
n−→∞
n
1
3. Prove that lim 1 + = e.
n→∞ n
[Hint: Use the Binomial expansion.]
4. For {sn } given by the following formulae, establish either the convergence
or divergence of {sn }.
(−1)n n nn 1 + (−1)n
(i) (ii) (iii)
n+1 (n + 1)n+1 n√
1 n
(iv) ln n − ln(n + 1) (v) 1 − n2 (vi) n n + 1.
5. Given that {sn } = {(−1)n }, show that {sn } is divergent but the sequence
s1 + · · · + sn
{tn } defined by tn = is convergent.
n
6. (a) Suppose {sn } is a sequence converging to 0, and {tn } another sequence
such that 0 ≤ tn ≤ sn for every index n. Show that {tn } converges to
0.
(b) Investigate
n the behaviour (by sketching a graph) of the sequence
n+1
, when n is odd,
sn = 1
1 − n , when n is even
Definition 3.12 Let {an } be a sequence of real numbers and let n1 < n2 <
n3 < · · · < nk < · · · be a strictly increasing sequence of positive integers. Then
the sequence an1 , an2 , an3 , · · · , ank , · · · is called a subsequence of {an } and is
denoted by {ank }
Example 3.10 If {an } = {a1 , a2 , a3 , a4 , · · ·}. Then two subsequences which im-
mediately come out are
26 Sequences
• {a2n−1 } = {a1 , a3 , a5 , · · ·}
Example 3.11 {an } = (−1)n . Let nk = 2k. Then the constant sequence
{a2k } = {1, 1, 1, · · ·}
1 3n 1 2n 1 n
Example 3.12 1 + 3n
, 1+ 2n
are both examples of subsequences of 1 + n
.
Theorem 3.9 A sequence {an } converges to a limit l if and only if every subse-
quence of {an } converges to l.
Definition 3.13 A sequence {un } is called a Cauchy sequence if, given ε > 0
∃ N ∈ N such that ∀ m, n > N |um − un | < ε.
|um − un | = |um − l + l − un |
≤ |um − l| + |l − un |
ε ε
< +
2 2
= ε
Proof 3.8 Let {xn } be a Cauchy sequence. Then given ε > 0 ∃ N ∈ N such that
∀ m, n > N |xn − xm | < ε. In particular, if we take ε = 1 and m = N + 1 we
obtain that ∀ n > N, |xn − xN +1 | < 1. In other words, for n > N ,
xN +1 − 1 < xn < xN +1 + 1.
Sequences 27
Now, let
M = max{x1 , x2 , · · · , xN , xN +1 + 1}
and
m = min{x1 , x2 , · · · , xN , xN +1 − 1}.
Then, clearly m < xn < M . We have proved that {xn } is bounded. It remains to
prove that {xn } is convergent.
Now, since {xn } is bounded, it has at least one convergent subsequence {xnk }.(This
is by which result?) Let xnk −→ l. Then given ε > 0 ∃ N1 ∈ N such that
∀ n > N1 |xnk − l| < 2ε . But, since {xn } is Cauchy, ∃ N2 ∈ N such that
ε
∀ m, n > N2 and ∀ nk > N1 , |xm − xn | < .
2
Let N = max{N1 , N2 }. Then for n > N, we have
and xn −→ l. ♠
4. If √
un+1 = 3un , u1 = 1,
show that un converges and find the limit.
n
5. By using the fact that 1 + n1 → e as n → ∞, show that the following
sequences
are convergent and find their limits in terms
of e:
1 n 1 n 1 n 1 1 n
1 + 3n , 1 − n , 1 + 2n , 1 − 2n − 2n2 .
6. The sequence {an } is defined by
r
a2n
a1 = 1, an+1 = 1+ , n = 1, 2, 3, · · ·
2
Show that (a) a2n − 2 < 0, (b) a2n+1 − a2n > 0. Deduce that {an }
converges and find its limit.
Theorem 3.13 To every sequence of nested intervals there corresponds one and
only one real number in common to all of them.
Proof 3.9 Let {In } be a sequence of nested intervals where In = [an , bn ]. Then
an ≤ an+1 and bn+1 ≤ bn and a1 ≤ a2 ≤ · · · ≤ an ≤ an+1 ≤ · · · ≤ bn+1 ≤
bn ≤ · · · ≤ b2 ≤ b1 . Thus, {an } is monotonic increasing bounded by a1 and b1
and {bn } is monotonic decreasing bounded by b1 and a1 . Hence the sequences
{an } and {bn } converge to a and b respectively. So ∀ ε > 0 ∃ N1 , N2 ∈
N s.t. |an − a| < 3ε ∀ n > N1 and |bn − b| < 3ε ∀ n > N2 . Further, since
ε
lim (bn − an ) = 0, ∃ N3 ∈ N s.t. |bn − an | < ∀ n > N3 . Now, let
n−→∞ 3
N = max{N1 , N2 , N3 }. Then ∀ n > N ;
|a − b| = |a − an + an − bn + bn − b|
= |(a − an ) + (an − bn ) + (bn − b)|
≤ |a − an | + |an − bn | + |bn − b|
ε ε ε
< + +
3 3 3
= ε
⇒a=b ♠
Unit 4
Infinite series: Convergence and
Divergence
4.2 Preliminaries
Let {un } be a sequence of numbers. Then
∞
X
u1 + u2 + · · · = un
n=1
Let
s1 = u1 , s2 = u1 + u2 , s3 = u1 + u2 + u3 ,
sn = u1 + u2 + · · · + un .
Then
{s1 , s2 , s3 , · · ·}
is referred to as the sequence of partial sums.
29
30 Infinite series: Convergence and Divergence
Note:
a(xn − 1)
lim sn = lim
n−→∞ n−→∞ x−1
a
= lim (xn − 1)
x−1 n−→∞
a
1−x
, if |x| < 1,
=
∞, if |x| > 1
∞ ∞
X
n−1 a X
Hence ax = if |x| < 1, i.e., the series converges and axn−1
n=1
1 − x n=1
diverges if |x| > 1.
∞
X
Activity 4.1 Determine whether or not the series (−1)n+1 converges or di-
n=1
verges.
1
Here x = < 1 =⇒ the series converges and
10
1 n
a(xn − 1) 3( 10 ) −1 10
lim sn = lim = lim 1 = .
n−→∞ n−→∞ x−1 n−→∞
10
−1 3
Hence ∞
X 3 10
n
= .
n=0
10 3
∞
X ∞
X
Let an and bn be convergent series. Write
n=1 n=1
s n = a1 + a2 + · · · + an ,
tn = b1 + b2 + · · · + bn .
Then
(a1 + b1 ) + (a2 + b2 ) + · · · + (an + bn ) = sn + tn
∞
X ∞
X
and sn + tn −→ s + t as n −→ ∞, where s = an and t = bn . Hence
n=1 n=1
∞
X ∞
X ∞
X ∞
X
(an + bn ) converges and (an + bn ) = an + bn . Thus, there is no
n=1 n=1 n=1 n=1
difficulty about adding convergent series.
∞
X 3n + 4n
Example 4.3 Consider the series . Observe that
n=0
5n
∞ ∞ ∞ ∞ n ∞ n
X 3n + 4n X 3n X 4n X 3 X 4
= + = + .
n=0
5n n=0
5n n=0
5n n=0
5 n=0
5
3 4
Both series are geometric with r = < 1 and r = < 1 respectively. This
5 5
implies that both series converge and hence the original series converges also.
Now,
( 3 )n − 1 ( 4 )n − 1 15
lim sn = lim 53 + lim 54 = .
n−→∞ n−→∞
5
−1 n−→∞
5
−1 2
∞
X 3n + 4n 15
Hence = .
n=0
5n 2
1
3
− 14 , · · · , an−1 = 1
n−1
− n1 , an = 1
n
− 1
n+1
. So
s n = a1 + a2 + a3 + · · · + an
1 1 1 1 1 1 1 1 1
= 1− + − + − + ··· + − + −
2 2 3 3 4 n−1 n n n+1
1
= 1−
n+1
and sn −→ 1 as n −→ ∞, so that the series
∞ ∞
X X 1
an =
n=1 n=1
n(n + 1)
∞
X 1
converges and = 1.
n=1
n(n + 1)
∞
1 X
Activity 4.2 Find an explicit expression for the partial sums of log 1 + .
n=1
n
∞
X 1
Deduce that log 1 + diverges.
n=1
n
∞
X 1 1 1 1
Consider the series = 1+ + + · · · + + · · ·. This series is called the
r r=1
2 3 n
Harmonic series. Now,
1
s2 = 1 +
2
1 1 1
s4 = 1+ + +
2 3 4
1 1 1
= 1+ + +
2 3 4
1 1 1
> 1+ + +
2 4 4
1 1
= 1+ +
2 2
1 1 1 1 1
s8 = 1 + + + + + ··· +
2 3 4 5 8
1 1 1 1 1
= 1+ + + + + ··· +
2 3 4 5 8
1 1 1 1 1 1 1
> 1+ + + + + + +
2 4 4 8 8 8 8
1 1 1
= 1+ + +
2 2 2
1 1 1 1 1 1 1
s16 = 1 + + + + + ··· + + + ··· +
2 3 4 5 9 10 16
1 1 1 1 1 1 1 1 1
> 1+ + + + + + + + + ··· +
2 4 4 8 8 8 8 16 16
1 1 1 1
= 1+ + + +
2 2 2 2
Let us have a summary of these observations:
s21 > 1
1
s21 ≥ 1 +
2
1 1
s22 > 1+ +
2 2
1 1 1
s23 > 1+ + +
2 2 2
1 1 1 1
s24 > 1+ + + +
2 2 2 2
..
.
1 1 1
s2 n > 1+ + + ··· + n − times
2 2 2
1
= 1+n
2
34 Infinite series: Convergence and Divergence
n
= 1+
2
1
= (n + 2) −→ ∞
2
∞
X 1
as n −→ ∞. Thus, lim s2n does not exist and hence the harmonic series
n−→∞
r=1
r
diverges.
where p is a constant. It will be shown later that this series converges for p > 1
and diverges for p ≤ 1. This series is called the p-series. Observe that when
p = 1, this series becomes the harmonic series.
∞
X 1
Example 4.6 The series is a p-series with p = 2.
n=1
(2n)2
∞
X
Proof 4.1 (a) The partial sums of the series un are bounded above by
n=1
∞
X
M = u1 + u2 + · · · + uN + vn .
n=N +1
∞
X
They therefore form a non-decreasing sequence with a limit L ≤ M . Thus un
n=1
converges.
∞
X
(b) Clearly, the partial sums of un are not bounded from above. If they were,
n=1
∞
X
the partial sums for vn would be bounded by
n=1
∞
X
0
M = v1 + v2 + · · · + vN + un
n=N +1
∞
X ∞
X
and vn would have to converge instead of diverge. Thus, un diverges. ♠
n=1 n=1
1 1 1 1
Example 4.7 Let un = and vn = . Then we see that > for all
∞
ln n n ∞
ln n n
X1 X 1
n ≥ 2. Now, since diverges (harmonic series), diverges also by
n=2
n n=2
ln n
the comparison test.
∞
X 1
Example 4.8 Test the series for convergence.
n=1
n(n + 2)
∞
1 1 1 1 X 1
Here un = . Let vn = 2 . Then < 2 ∀ n and is
n(n + 2) n n(n + 2) n n=1
n2
∞
X 1
convergent (p-series). =⇒ converges by comparison test.
n=1
n(n + 2)
n o
an
Proof 4.2 Let bn
−→ l 6= 0. Note that l > 0.
an 1
=⇒ ∃ N ∈ N : −l < l ∀ n>N
bn 2
i.e.,
1 an 1
=⇒ − l < −l < l
2 bn 2
1 an 3
=⇒ l < < l
2 bn 2
1 3
lbn < an < lbn ∀ n > N.
2 2
∞
X ∞
X
Now, if bn converges, an must also converge by the comparison test since
n=1 n=1
∞ ∞
3 X X
0 < an < lbn ∀ n > N . Moreover, if an converges then bn must also
2 n=1 n=1
2an
converge by the comparison test since 0 < bn < ∀ n > N . Thus if one of
l ∞ ∞
X X
the series converges, then the other also converges, i.e., either an and bn
n=1 n=1
∞
X ∞
X
both converge or an and bn both diverge. ♠
n=1 n=1
1 1
Example 4.9 Let an = and bn = . Then for all n > 1, an > 0, bn > 0
ln n n
and
an n
=
bn ln n
and
an n
lim = lim = lim n = ∞.
n−→∞ bn n−→∞ ln n n−→∞
∞
X1 ∞
X 1
Now, since diverges (harmonic series), it follows that diverges also
n=2
n n=2
ln n
by the limit form of the comparison test.
1 1
Example 4.10 Let an = , and bn = 2 . Then For all n ≥ 1, an >
n(n + 2) n
0, bn > 0 and
an n2
= 2 −→ 1
bn n + 2n
∞
X 1
as n −→ ∞. Now is a convergent p-series, so by the limit form of com-
n=1
n2
∞
X 1
parison test, also converges.
n=1
n(n + 2)
Infinite series: Convergence and Divergence 37
X n X ln n
Activity 4.3 Test for convergence: (a). √ , (b). √ .
3n5 − 8 n+1
Example 4.11 (p-series)
1 1
1. Let an = p
where p ≥ 2 and bn = 2 . For all n ∈ N, an > 0, bn > 0 and
n n
1 1
0 < an = p ≤ 2 = bn .
n n
∞
X ∞
X
Now, since bn converges, it follows by the comparison test that an
n=1 n=1
converges. So
∞
X 1
p
converges ∀ p ≥ 2.
n=1
n
1 1
2. Let an = , bn = p where p ≤ 1. For all n ∈ N, an > 0, bn > 0 and
n n
1 1
0 < an = ≤ p = bn .
n n
∞
X ∞
X
Since p ≤ 1, by the comparison test bn diverges, since an diverges
n=1 n=1
(harmonic series). Thus
∞
X 1
diverges ∀ p ≤ 1.
n=1
np
∞
X1
Activity 4.4 Test for convergence: sin
n=1
n
∞
X an+1
Corollary 4.1 Let an be a series of positive terms such that −→ ∞ as
n=1
an
∞
X
n −→ ∞. Then an diverges.
n=1
[(2n)!]2
Example 4.12 Let an = for all n ∈ N. Then an > 0 and
(4n)!
an+1 [(2n + 2)!]2 (4n)!
=
an (4n + 4)![(2n)!]2
(2n + 2)(2n + 1)(2n + 2)(2n + 1)
=
(4n + 4)(4n + 3)(4n + 2)(4n + 1)
(2n + 2)(2n + 1)
=
2(4n + 3)2(4n + 1)
4n2 + 6n + 2 1
= −→
4(16n2 + 16n + 3) 16
∞
X [(2n)!]2
as n −→ ∞. Since the limit is less than 1, converges.
n=1
(4n)!
∞ ∞
X 1 X n7
Activity 4.5 Test for convergence using the ratio test: (a). (b)
n=1
n2 n=1
2n
∞
X 2n
(c) 3 (n + 4)
.
n=1
n
If we include u1 , we have
Z n
u1 + u2 + u3 + · · · + un ≤ u1 + f (x)dx.
1
Infinite series: Convergence and Divergence 41
1
• If p > 1, lim = 0 and series converges.
M −→∞ M p−1
1
• If p < 1, lim diverges and the series diverges.
M −→∞ M p−1
• If p = 1, lim ln M diverges and series diverges.
M −→∞
∞
X 1
Activity 4.6 Use the integral test to test the series for convergence.
n=1
n2
Proof 4.5
sn = a1 − a2 + a3 − a4 + · · · + (−1)n+1 an first n terms
s2n = a1 − a2 + a3 − a4 + · · · + a2n−1 − a2n first 2n terms
s2n+2 = a1 − a2 + a3 − a4 + · · · + a2n−1 − a2n + a2n+1 − a2n+2
∴ s2n+2 − s2n = a2n+1 − a2n+2 ≥ 0,
since {an } is decreasing and, therefore, a2n+1 ≥ a2n+2 . Thus s2n+2 ≥ s2n and the
sequence {s2n } is increasing. Moreover,
s2n = a1 − a2 + a3 − a4 + a5 − · · · − a2n−2 + a2n−1 − a2n
= a1 − (a2 − a3 ) − (a4 − a5 ) − · · · − (a2n−2 − a2n−1 ) − a2n
≤ a1
because a2 ≥ a3 ≥ a4 ≥ · · · ≥ a2n−2 ≥ a2n−1 and a2n > 0. Thus s2n ≤ a1 for all n
and the sequence {s2n } is bounded above. Since it is also increasing, by PMBC,
it must converge to some number s, i.e., s2n −→ s as n −→ ∞. Now, if we
consider the first 2n + 1 number of terms, s2n+1 , we have
s2n+1 = a1 − a2 + a3 − a4 + · · · + a2n−1 − a2n + a2n+1
= s2n + a2n+1
As n −→ ∞, a2n+1 −→ 0 and s2n −→ s. It follows that s2n+1 −→ s as n −→ ∞.
Since s2n −→ s and s2n+1 −→ s as n −→ ∞, we see that sn −→ s as n −→ ∞
∞
X
and the series therefore converges and (−1)n+1 an = s. ♠
n=1
1
Example 4.15 Let an = sin n
, n = 1, 2, 3, · · ·. For
1 1 π
n ≥ 1, 0 < < ≤1< ,
n+1 n 2
and therefore
1 1
an+1 = sin < sin = an ,
n+1 n
i.e., {an } is a decreasing sequence of positive real numbers. Furthermore,
an = sin n1 −→ 0 as n −→ ∞. By the alternating series test,
∞
X
n+1 1
(−1) sin
n=1
n
converges.
Infinite series: Convergence and Divergence 43
Activity 4.7 What happens to the series in example 4.15 if we remove the factor
(−1)n+1 ?
∞
X
n+1 1
Activity 4.8 Test for convergence: (−1) cos .
n=1
n
Theorem 4.4 Every absolutely convergent series of real numbers is also conver-
gent.
∞
X
Proof 4.6 Let an be an absolutely convergent series of real numbers, i.e., it
n=1
∞
X ∞
X
is a series such that |an | converges. Now, some of the terms of an may
n=1 n=1
be positive and some may be negative. We therefore
separate the positive and
an , if an ≥ 0,
negative terms in the following way: Let bn = ,
0, if an < 0
0, if an ≥ 0,
cn = Then bn ≥ 0, cn ≥ 0 for all n, and an = bn − cn .
−an , if an < 0
Write
Definition 4.2 A series of real numbers which is convergent, but not absolutely
convergent, is called a conditionally convergent series.
∞
X
n+11 2
. We notice that (−1)n+1 sin2 1
Example 4.16 Consider (−1) sin n
=
n=1
n
1
sin2 n1 . Write an = sin2 n1 , bn = 2 . Then an > 0, bn > 0 and
n
! !
sin2 n1 sin n1 sin n1
an
= 1 = 1 1 −→ 1 · 1 = 1
bn n2 n n
∞
X 1
as n −→ ∞. Now 2
converges, and therefore by the limit form of the compar-
n=1
n
∞ ∞
X
2 1 X
n+1 2 1
ison test, sin is convergent, i.e., (−1) sin is convergent.
n=1
n n=1
n
∞
X
2 1
Thus sin is absolutely convergent and therefore convergent.
n=1
n
The notion of absolute convergence allows us to extend the ratio test in the
following way:
∞
X an+1
Theorem 4.5 Let an be a series of non-zero real numbers. If an
−→ l as
n=1
n −→∞∞ then
X
(a). an is absolutely convergent (and so convergent) if l < 1.
n=1
∞
X
(b). an is divergent if l > 1.
n=1
converges.
(b). l > 1.
√
For all indices beyond some integer M , we have n un > 1, so that un > 1 for
n > M . The terms of the series do not converge to 0. The series diverges. ♠
∞
th
X n2
Example 4.17 Use the n root test to test the series for convergence.
n=1
2n
r √
n √
n n2 n2 ( n n)2 1
n
= √
n
= −→ < 1.
2 2n 2 2
Thus, the series converges by the nth root test.
∞
X 2n
Activity 4.9 Test for convergence: .
n=1
n2
1
Activity 4.10 1. Let an = , n = 1, 2, 3, · · ·. Let sn = a1 +
n(n + 1)(n + 2)
a2 + · · · + an . Express an in partial fractions and hence, or otherwise, show
that
1 1 1
sn = − + .
4 2(n + 1) 2(n + 2)
∞
X 1
Deduce that converges and find its sum.
n=1
n(n + 1)(n + 2)
Deduce that
∞
X 1
log 1 +
n=1
n
diverges.
∞
X 1
3. Determine the sum, if it exists, of the series .
n=1
(5n − 2)(5n + 3)
∞ ∞ ∞
n2 + 2
X X 1 X
n 1
(d) (e) (f ) (−1) sin √
n=1
3n3 + 4n n=2
n ln n n=1
n
∞ ∞ ∞
X (5n)! X (n2 + 1)4 X 2 + (−1)n
(g) (h) (i) 3
n=1
(30)n (3n)!(2n)! n=1
(n3 + 1)3 n=1 n2
∞ ∞ ∞
X n2 + 1 X (2n)! X sin(n!)
(j) 4 2
(k) 2
(l)
n=1
4n + n − n − 1 n=1
(n!) n=1
n2 +n+1
∞ ∞ ∞
X (−1)n n2 X 1 X 9n (3n)!n!
(m) (n) (o)
n=1
n3 + 1 n=2
n(ln n)3 n=1
(4n)!
∞ ∞ ∞
2 + (−1)n n2 + n + 1 (−1)n
X X X
(p) √ (q) √ (r) sin
n=1
n n=1
n7 + n4 + 1 n=1
n
Infinite series: Convergence and Divergence 47
∞ ∞ ∞
X 1 X (−1)n+1 X (−1)n
(s) 3 (t) (u)
n=2 n(ln n) 2
n=1
2n n=1
n2 + 2n + 2
∞ ∞ ∞
X (−1)n−1 X (−1)n−1 n X (−1)n
(a) (b) (c)
n=1
n2 + 1 n=1
n2 + 1 n=2
n ln n
∞ ∞ ∞
X (−1)n n3 X (−1)n−1 1 X (−1)n−1 n3
(d) 4 (e) sin √ . (f )
n=1 (n2 + 1) 3 n=2
2n − 1 n n=1
2n − 1
Unit 5
Power Series
or
∞
X
an (x − c)n = a0 + a1 (x − c) + a2 (x − c)2 + · · · + an (x − c)n + · · ·
n=0
A power series converges for |x| < R and diverges for |x| > R, where the constant
R is called the radius of convergence of the series. For |x| = R, the series may
or may not converge.
The interval |x| < R or −R < x < R, with possible inclusion of end points, is
called the interval of convergence of the series.
Theorem 5.1 A power series converges uniformly and absolutely in any interval
which lies entirely within its interval of convergence.
48
Power Series 49
|x|
=
3
|x| |x|
and by the ratio test the series converges if < 1, and diverges if > 1. If
3 3
|x|
= 1, i.e., x = ±3, the test fails.
3 ∞ ∞
X 1 1X1
• If x = 3, the series becomes = which diverges (harmonic se-
n=1
3n 3 n=1 n
ries).
∞ ∞
X (−1)n−1 1 X (−1)n−1
• If x = −3, the series becomes = which converges
n=1
3n 3 n=1 n
by the alternating series test. Thus the interval of convergence is −3 ≤ x < 3.
50 Power Series
(−1)n−1 x2n−1
(b). Here fn = . Then
(2n − 1)!
fn+1 (−1)n x2n+1 (2n − 1)!
lim = lim ·
n−→∞ fn n−→∞ (2n + 1)! (−1)n−1 x2n−1
(−1)n x2n x(2n − 1)!
= lim
n−→∞ (2n + 1)(2n)(2n − 1)!(−1)n (−1)−1 x2n x−1
−x2
= lim
n−→∞ 2n(2n + 1)
= 0
and the series converges (absolutely) for all x. So the interval of (absolute) con-
vergence is −∞ < x < ∞.
∞
X
Proof 5.1 Let R > 0 be the radius of convergence of an xn . Let 0 < |x0 | < R.
n=0
Then we can choose N so that
1
|an | <
|x0 |n
for n > N . Thus the terms of the series
X X
|nan xn−1 | = n|an ||x|n−1
can, for n > N , be made less than corresponding terms of the series
X n|x|n−1
|x0 |n
X
which converges, by the ratio test, for |x| < |x0 | < R. Hence nan xn−1 con-
verges absolutely for all points x0 (no matter how close |x0 | is to R). If, however,
Power Series 51
X
|x| > R, lim an xn 6= 0 and thus lim nan xn−1 6= 0, so that nan xn−1 does
n−→∞ n−→∞ X
not converge. Thus R is the radius of convergence of nan xn−1 . ♠
∞
X xn
Example 5.3 Illustrate Example 5.2 by using the series 2 3n
.
n=1
n
Here
xn+1 n2 3n n2 |x| |x|
lim · = lim = .
n−→∞ (n + 1)2 3n+1 xn n−→∞ (n + 1)2 3 3
Thus, by an earlier example, the radius of convergence is −3 ≤ x < 3. Now,
∞
X xn−1
consider the series of derivatives . This has already been found to have
n=1
n3n
radius of convergence −3 ≤ x < 3. This proves.
a0 = lim f (a + h)
h−→0
2.
en (h) f (a + h) − f (a)
= − (a1 + a2 h + · · · + an hn−1 )
h h
and so en (h) = (h) if and only if f 0 (a) exists and a1 = f 0 (a).
52 Power Series
1 (m)
am = f (a).
m!
Definition 5.2 For any f such that f (n) (a) exists, the nth Taylor polynomial
of f at a is
∞
X f (k) (a)
Tn (f ; a) = (x − a)k
k=0
k!
f 00 (a) f (n) (a)
= f (a) + f 0 (a)(x − a) + (x − a)2 + · · · + (x − a)n + · · ·
2! n!
The coefficients are the Taylor coefficients of f at a. The error when f is
approximated near a by Tn satisfies
e(k)
n (h) = (h), k = 0, 1, 2, · · · , n − 1
f (a) + f 0 (a)(x − a)
Remark 5.3 If we take the case n = 0, then Taylor’s Theorem is just the Mean-
Value Theorem. Thus, Taylor’s Theorem gives generalizations of the Mean-Value
Theorem.
Z 1 2
1 − e−x
Example 5.4 Evaluate dx.
0 x2
u2 u3 u4 u5
eu = 1 + u + + + + + · · · , −∞ < u < ∞
2! 3! 4! 5!
Then if u = −x2 , we have
2 x4 x6 x8 x1 0
e−x = 1 − x2 + − + − + · · · , −∞ < x < ∞
2! 3! 4! 5!
Thus
2 x4 6 8 x10
1 − e−x 1 − (1 − x2 + − x3! + x4! −
2! 5!
+ · · ·)
=
x2 x2
2 4 6
x x x x8
= 1− + − + + ···
2! 3! 4! 5!
54 Power Series
Since the series converges for all x and so, in particular, converges uniformly for
0 ≤ x ≤ 1, we can integrate term by term to obtain
1 2
1 − e−x
Z
1 1 1 1
2
dx = 1 − + − + − ···
0 x 3 · 2! 5 · 3! 7 · 4! 9 · 5!
x3 x5 x7
tan−1 x = x − + − + ···
3 5 7
where the series is uniformly convergent in −1 ≤ x ≤ 1.
2. Prove that
π 1 1 1
= 1 − + − + ···
4 3 5 7
Unit 6
Limits and Continuity
Definition 6.1 A point x0 is called a cluster point of the set D if and only if
∀ δ > 0 ∃ x ∈ D such that 0 < |x − x0 | < δ.
Thus, a cluster point x0 of a set D has the property that it is always possible to
find points x ∈ D for which x 6= x0 and yet x is as close to x0 as we like.
Example 6.1 Let D = [0, 1). The set of cluster points of D is the interval [0, 1].
Example 6.2 Let D = n1 , n ∈ N . Then D has only one cluster point; namely
55
56 Limits of Indeterminate Type
We need δ > 0 such that 0 < |x − x0 | < δ ⇒ |(mx + b) − (mx0 + b)| < ε ⇔
ε ε
|m(x − x0 )| = |m||x − x0 | < ε ⇒ |x − x0 | < |m| , provided m 6= 0. So δ = |m| .
(Notice that if m = 0, then all x would satisfy the required inequality).
x2 − 1 2 −1
• ∀ x 6= 1, lim = 2, since xx−1 = x + 1 −→ 2 as x −→ 1.
x−→1 x − 1
0, if x 6= 0,
• Let f (x) = Then lim f (x) = 0 6= f (0).
1, if x = 0 x−→0
1, if x ≥ 0,
• f (x) = Then lim f (x) does not exist. In fact, no matter
0, if x < 0 x−→0
how small we make δ > 0, the inequality 0 < |x − 0| < δ will be satisfied by
points x for which f (x) can be either 1 or 0, and we cannot force 1 and 0
to be within arbitrarily small ε > 0 of an one number l.
T
Theorem 6.1 Suppose x0 is a cluster point of the domain Df Dg , lim f (x) =
x−→x0
l and lim g(x) = M . Then
x−→x0
• lim (f ± g)(x) = l ± M
x−→x0
• lim (f g)(x) = lM
x−→x0
f l
• lim (x) = , provided M 6= 0 and that x0 is a cluster point of D f .
x−→x0 g M g
• lim f (x) = l.
x−→x0
(⇐), i.e.(b) ⇒ (a). Suppose (b) is true. We need to show that lim f (x) = l. Now
x−→x0
suppose this conclusion were false, i.e. suppose that lim f (x) 6= l. Then ∃ ε > 0
x−→x0
such that ∀ δ > 0 ∃ x ∈ Df such that 0 < |x − x0 | < δ and yet |f (x) − l| ≥ ε.
In particular, if we let δn = n1 , then we get xn such that 0 < |xn − x0 | < n1 and
yet |f (xn ) − l| ≥ ε. Now, xn ∈ Df \{x0 } and xn −→ x0 , yet f (xn ) 9 l. This
contradicts (b) which was assumed to be true.♠
sin x1 , if x 6= 0,
Example 6.4 Let f (x) = Show that lim f (x) does not ex-
0, if x = 0 x−→0
ist.
2
Let xn = , n = 1, 2, 3, · · ·. Observe that xn → 0. Now f (xn ) =
(2n
+ 1)π
sin (2n+1)π
2
= (−1)n which diverges. We conclude that f has no limit at x = 0.
xn − xn0
Activity 6.1 • Find lim , n ∈ N.
x−→x0 x − x0
1
• Prove: lim f (x) = ∞ ⇔ lim = 0.
x−→∞ x−→∞ f (x)
Remark
If f is continuous at every point x0 ∈ Df , we say f ∈ C(Df ), the family of all
continuous functions.
T
Theorem 6.3 Suppose f and g are each continuous at x0 ∈ Df Dg . Then
• f ± g is continuous at x0 .
• f g is continuous at x0 .
f
• is continuous at x0 provided g(x0 ) 6= 0.
g
58 Limits of Indeterminate Type
Activity 6.2 1. Prove part (c) of Theorem 6.3. The domain D f is that subset
T g
of Df Dg consisting of points x for which g(x) 6= 0.
2. Prove that if a(x) = |x|, the absolute value function, then a ∈ C(R).
√
3. Suppose Q(x) = x is defined ∀ x ≥ 0. Prove that Q ∈ C([0, ∞)).
xn −xn
x−x0
0
, if x 6= x0 ,
6. Let f (x) = Find the value of c which makes
c, if x = x0 , n ∈ N
f ∈ C(R)
f (p)
Proof 6.4 Let ε = , which is positive. Then ∃ δ > 0 such that x ∈
2
T f (p) 3f (p)
Df (p − δ, p + δ) ⇒ |f (x) − f (p)| < ε ⇒ 0 < < f (x) < , which
2 2
proves the lemma.
Proof 6.5 We will suppose that f (a) < k < f (b). We will let a1 = a and b1 = b
a1 + b 1
and we will use the method of interval halving. Let m = . If f (m) > k, let
2
b2 = m and a2 = a1 . Otherwise let a2 = m and b2 = b1 . Then halve the interval
and proceed as above in that fashion. We generate two sequences of endpoints
both of which are guaranteed to be Cauchy since they are sequences of left and
right endpoints respectively, generated by interval halving. Thus an → Ca and
bn → Cb . But {an } is monotone increasing bounded above by Ca and {bn } is
b−a
monotone decreasing bounded below by Cb . So |Ca − Cb | ≤ |bn − an | = n−1 → 0.
2
so Ca = Cb = c. Since c ∈ [a, b], f is continuous at c. Thus f (c) = lim an ≤ k
n−→∞
and f (c) = lim bn ≥ k of which the two inequalities can be satisfied if and only
n−→∞
if f (c) = k ♠
Remark
If f ∈ C([a, b]), for a closed finite interval [a, b], we can conclude more than just
the continuity of f at each point c ∈ [a, b]. The next theorem shows that f will
be uniformly continuous on [a, b].
1
Example 6.6 Let f (x) = on (0, 1). We will show that f is continuous on
x
(0, 1). Let x0 be any point in (0, 1). We want to show that given ε > 0 ∃ δ > 0
such that |f (x) − f (x0 )| < ε whenever 0 < |x − x0 | < δ. Now
1 1 |x − x0 |
|f (x) − f (x0 )| = − =
x x0 |x||x0 |
Our problem is how to control the denominator. We see that as x becomes small,
the fraction becomes large. However, if x is reasonably close to x0 , for example,
x0
if we take x > , we have
2
1 1 |x − x0 | 2|x − x0 | 2δ
|f (x) − f (x0 )| = − = < <
x x0 |x||x0 | |x0 |2 |x0 |2
2δ ε|x0 |2
In other words we want = ε, i.e. δ = . Thus if |x − x0 | < δ, then
|x0 |2 2
|f (x) − f (x0 )| < ε.
We see that δ depends on both the point x0 and ε. Then, given ε > 0, the same
δ DOES NOT work for all points.
Example 6.7 Let f (x) = x2 . We will prove that f (x) is uniformly continuous
on [0, 4].
We want to show that ∀ ε > 0 ∃ δ > 0 such that ∀ x1 ∈ [0, 4] and ∀ x ∈ [0, 4]
|f (x) − f (x1 )| < ε whenever |x − x1 | < δ. Now,
Theorem 6.6 If f ∈ C([a, b]), i.e. if f is continuous on [a, b], then f is uniformly
continuous on [a, b].
Proof 6.6 We suppose that the theorem were false, and we will deduce a contra-
diction. Thus we suppose f is not uniformly continuous on [a, b]. Then ∃ ε > 0
such that no δ > 0 will suffice to meet the uniform continuity condition. Hence
1 1
if δn = , ∃ xn , yn ∈ [a, b] such that |xn − yn | < but |f (xn ) − f (yn )| ≥ ε. Now,
n n
Limits of Indeterminate Type 61
1
Example 6.8 Let f (x) = on (0, 1]. We show that f is not uniformly contin-
x
1
uous on (0, 1] using the above theorem. Let xn = . Then clearly xn ∈ (0, 1] and
n
∀ m, n > N we have
1 1 1 1 1
− < max , < <ε
m n m n N
Activity 6.3 1. Suppose f ∈ C([a, b]) and f (a) > k > f (b). Prove: ∃ c ∈
(a, b) such that f (c) = k. [Hint: Consider g(x) = −f (x).]
3. Let p(x) = x4 +x3 −2x2 +x+1. Prove that the polynomial equation p(x) = 0
has a root in (−1, 0).
62 Limits of Indeterminate Type
5. Prove the following fixed point theorem: Suppose f ∈ C([0, 1]) and suppose
0 ≤ f (x) ≤ 1 ∀ x ∈ [0, 1]. Then ∃ c ∈ [0, 1] such that f (c) = c. (The
point c is then called a fixed point for the function f .)[Hint: Consider
g(x) = f (x) − x].
6. Suppose f is uniformly continuous on (a, m] and also on [m, b). Prove that
f is uniformly continuous on (a, b).
1
7. Let f (x) = , ∀ x ∈ (0, 1). Is f uniformly continuous on (0, 1)? Justify
x
your conclusions.
7.2 Differentiation
The concept of differentiation was introduced in SMTH011 and there you dealt
with problems of finding derivatives of various functions. In this unit we deal
with differentiation as a continuation from where it was left there in SMTH011.
Suffice to say that the techniques and differentiation formulae that you have
learnt remain the same, the only new thing is that this time around we will be
applying these to different type of functions. In particular we focus on inverse
trigonometric functions, hyperbolic functions, inverse hyperbolic functions and
parametric equations.
Perhaps it will be worthwhile to start by recalling the techniques/formulae that
you have used before.
If y = f (x), then
dy f (x + h) − f (x)
= lim
dx h→o h
63
64 Differentiation and Partial Differentiation
Derivative of a constant
d
If k is a constant, (k) = 0
dx
dy
You are also urged to recall implicit differentiation for finding when y is an
dx
implicit function of x. This, as well as logarithmic differentiation, will also be
encountered in the foregoing.
Now, since the inverse sine function sin−1 x is continuous in the interval in which it
is defined, it is differentiable. We use implicit differentiation to find its derivative.
dy
Let y = sin−1 x, then sin y = x =⇒ cos y = 1 and
dx
dy 1 p √
= . Now, since cos y = 1 − sin2 y = 1 − x2 then
dx cos y
dy 1 1
= =√
dx cos y 1 − x2
and so
d 1
(sin−1 x) = √ −1<x<1
dx 1 − x2
dy
In general if y = sin−1 (f (x)), then sin y = f (x) and cos y = f 0 (x), so that
dx
dy f 0 (x) f 0 (x)
= =p
dx cos y 1 − [f (x)]2
3x2 −1
dy
Example 7.1 If y = arcsin 2
find and simplify the answer.
dx
Solution
3x2 − 1
Here f (x) = =⇒ f 0 (x) = 3x and therefore
2
dy 3x 3x 6x
=q =q =√
dx 2 2 4−(3x2 −1)2 3 − 9x4 + 6x2
1 − 3x 2−1
4
66 Differentiation and Partial Differentiation
Solution
2 1 1 0 1
First note that arcsec(2x ) = arccos . Thus f (x) = =⇒ f (x) = −
2x2 2x2 x3
and
− − x13 1
dy x3 2
=q = √ = √
dx 4x4 −1 x 4x4 − 1
1 − 4x14 2x2
d
Example 7.3 Find (cot−1 (cosec x))
dx
Solution
d −1 d −1 1 d cos x
(cot (cosec x)) = tan = (tan−1 (sin x)) =
dx dx cosec x dx 1 + sin2 x
Activity 7.2 1. Differentiate the following and simplify where possible
(a) y = arctan x+1
2
(b) y = (tan−1 x)2
(c) y = tan−1 x2
1
(d) y =
tan−1 x
(e) y = cos−1 (sin−1 x)
q
(f ) y = arctan 1−x
1+x
b+a cos x
(g) y = arccos a+b cos x
1+x2
(h) y = arccos 1−x2
x
(i) y = arctan 16−x2
(j) y = cos−1 (e ) 2x
(k) y = earccot 2x
d2 y dy
2. If y = earctan x , show that (1 + x2 ) 2
+ (2x − 1) =0
dx dx
2arcsin x d2 y2 dy
3. If y = e , prove that (1 − x ) 2 − x − 4y = 0
dx dx
ex + e−x ex − e−x
Definition 7.1 1. cosh x = 2. sinh x =
2 2
sinh x ex − e−x
From these two, we can define tanh x = = x ,
cosh x e + e−x
cosh x ex + e−x 1 2
coth x = = x −x
, sech x = = x ,
sinh x e −e cosh x e + e−x
1 2
cosech x = = x , etc,etc.
sinh x e − e−x
Hyperbolic functions also play an important role in integral calculus.
d ex + e−x ex − e−x
d
Now, (cosh x) = = = sinh x
dx dx 2 2
68 Differentiation and Partial Differentiation
ex − e−x ex + e−x
d d
and (sinh x) = = = cosh x.
dx dx 2 2
From these two derivatives, and in conjunction with the laws or techniques of
differentiation, we can obtain the derivative of any given hyperbolic function.
dy
Example 7.4 If y = cosh3 x, find .
dx
Solution
dy
The chain rule gives = 3 cosh2 x sinh x
dx
dy
Example 7.5 If y = sinh2 x cosh x, find .
dx
Solution
The chain rule and product rule give
dy
= 2 sinh x cosh x cosh x + sinh2 x sinh x = 2 sinh x cosh2 x + sinh3 x
dx
dy
Example 7.6 If y = x3 etanh x , find .
dx
Solution
dy
= 3x2 etanh x + x3 sech2 xetanh x
dx
Solution
dy sinh x 1 sinh x sinh(sin−1 x)
=p + sinh(sin−1 x) · √ =p + √
dx 1 − cosh2 x 1 − x2 1 − cosh2 x 1 − x2
2
1. y = cosh
x
1
2. y = sinh2
x
√
3
3. y = tanh x2
1+tanh x
4. y = ln 1−tanh x
8. y = xsinh 3x
9. y = sech3 (2x )
ey + e−y
=x
2
=⇒ ey + e−y = 2x
=⇒ e2y − 2xey + 1 = 0
=⇒ z 2 − 2xz + 1 = 0, z = ey
√
2x ± 4x2 − 4
=⇒ z =
√2
=⇒ z = x ± x2 − 1
√
=⇒ ey = x ± x2 − 1
√
=⇒ y = ln(x ± x2 − 1)
√
=⇒ cosh−1 x = ln(x + x2 − 1), x ≥ 1
and
1 1+x
arctanh x = ln −1<x<1
2 1−x
70 Differentiation and Partial Differentiation
Solution
y = cosech−1 (sech x3 ) = sinh−1 1
= sinh−1 (cosh x3 ). Thus,
sech x3
dy 3x2 sinh x3
=p
dx 1 + cosh2 x3
Differentiation and Partial Differentiation 71
d2 y dy
3. If ln y = cosh−1 x, find the value of (1 + x2 ) 2
+x −y
dx dx
x = x(θ), y = y(θ)
are called parametric equations. Now, given such parametric equations, our ob-
dy d2 y
jective here is to find and 2 .
dx dx
dy dy dθ dθ 1
Let x = x(θ), y = y(θ). Chain rule gives = · . Since = dx , we
dx dθ dx dx dθ
have that
dy
dy dθ
= dx .
dx dθ
d2 y
d dy d dy dθ
Further 2 = = · , which gives
dx dx dx dθ dx dx
d dy
d2 y dθ dx
= dx .
dx2 dθ
72 Differentiation and Partial Differentiation
dy d2 y
Example 7.11 Given x = 2 cos 2t, y = 2 sin t, determine and .
dx dx2
Solution
dx dy
= −4 sin 2t, = 2 cos t and
dt dt
dy
dy dt 2 cos t cos t cosec t
= dx
= =− =−
dx dt
−4 sin 2t 2 sin 2t 4
d dy cosec t cot t
Now, = and
dt dx 4
d dy
d2 y dt dx cosec t cot t cosec3 t
= dx = − =−
dx2 dt
16 sin 2t 32
Example 7.12 Obtain an equation of the tangent line to the parametric curve
Solution
dx dy
= 4 cos 2t, = 2 cos t. The slope is
dt dt
dy
dy dt 2 cos t cos t
= dx
= =
dx dt
4 cos 2t 2 cos 2t
√ π
The point ( 3, 1) corresponds to the parameter value t = , so that the slope of
6
the tangent at this point is
√
cos π6
dy 3
|t= π6 = π
=
dx 2 cos 3 2
√ √
3 √ 3 1
Thus, the equation of the tangent line is y − 1 = (x − 3) =⇒ y = x− .
2 2 2
Activity 7.7 1. Determine an equation of the tangent line to the curve
x = t + 1, y = t3 + t at the point t = −1.
4
dy d2 y
4. Determine and for each of the following:
dx dx2
(a) y = sinh t, x = t + cosh t
(b) x = cos−1 (ln t), y = −2 sin−1 (3t )
(c) y = sin θ + θ cos θ, x = cos θ − θ sin θ
π
(d) y = 3 sinh θ, x = cosh3 θ, at θ =
6
t t
(e) x = e cos t, y = e sin t
(f ) x = cos3 2θ, y = sin3 2θ
(g) x = a cos 2θ, y = a sin θ, a-constant.
f (x + h, y) − f (x, y)
fx (x, y) = lim
h→o h
and
f (x, y + h) − f (x, y)
fy (x, y) = lim
h→o h
∂f ∂ ∂z
fx (x, y) = fx = = f (x, y) =
∂x ∂x ∂x
and
∂f ∂ ∂z
fy (x, y) = fy = = f (x, y) =
∂y ∂y ∂y
74 Differentiation and Partial Differentiation
Solution
fx (x, y) = 3x2 − 2xy 3 and therefore fx (2, −1) = 3 · 22 − 2 · 2 · (−1)3 = 16
fy (x, y) = −3x2 y 2 + 4y and so f − y(2, −1) = −3 · 22 · (−1)2 + 4 · (−1) = −16
x
∂f ∂f
Example 7.14 If f (x, y) = cos 1+y
, calculate and .
∂x ∂y
Solution
∂f x ∂ x x 1 1 x
= − sin = − sin · =− sin
∂x 1+y ∂x 1+y 1+y 1+y 1+y 1+y
and
∂f x ∂ x x −x x x
= − sin = − sin · = sin
∂y 1+y ∂y 1+y 1 + y (1 + y)2 (1 + y)2 1+y
∂z ∂z
Example 7.15 Find and if z is defined implicitly as a function of x and
∂x ∂y
y by the equation x2 + y 2 + z 2 + 3xyz = 1
Solution
∂z ∂z
2x + 2z
+ 3yz + 3xy =0
∂x ∂x
∂z
=⇒ (2z + 3xy) = −2x − 3yz
∂x
∂z 2x + 3yz
=⇒ =−
∂x 2z + 3xy
Similarly,
∂z ∂z
2y + 2z + 3xz + 3xy =0
∂y ∂y
∂z 2y + 3xz
=⇒ =−
∂y 2z + 3xy
Differentiation and Partial Differentiation 75
Example 7.16 If f (x1 , x2 , x3 ) = x1 sin(2x2 −x3 ), find the first partial derivatives
∂f ∂f ∂f
, and .
∂x1 ∂x2 ∂x3
Solution
∂f
= sin(2x2 − x3 )
∂x1
∂f
= 2x1 cos(2x2 − x3 )
∂x2
and
∂f
= −x1 cos(2x2 − x3 )
∂x3
∂ 2f
∂ ∂f
(fy )y = fyy = 2
=
∂y ∂y ∂y
xy
Example 7.17 Given z = , find all the second partial derivatives.
x−y
Solution
∂z y(x − y) − xy y2
= = −
∂x (x − y)2 (x − y)2
∂z x(x − y) + xy x2
= =
∂y (x − y)2 (x − y)2
∂ 2z y2 2y 2
∂ ∂z ∂
= = − = −
∂x2 ∂x ∂x ∂x (x − y)2 (x − y)3
∂ 2z y2 −2y(x − y)2 − 2y 2 (x − y) 2xy 2 − 2x2 y
∂ ∂z ∂
= = − = =
∂y∂x ∂y ∂x ∂y (x − y)2 (x − y)4 (x − y)4
∂ 2z x2 2x(x − y)2 − 2x2 (x − y) 2xy 2 − 2x2 y
∂ ∂z ∂
= = = =
∂x∂y ∂x ∂y ∂x (x − y)2 (x − y)4 (x − y)4
∂ 2z x2 2x2
∂ ∂z ∂
= = = −
∂y 2 ∂y ∂y ∂y (x − y)2 (x − y)3
Activity 7.8 1. Determine the first partial derivatives of the following func-
tions:
(f ) z = f (x + y)
(g) z = f (x)g(y)
(h) z = f (xy)
(i) z = f ( xy )
3. Obtain all the second partial derivatives
−1 x+y
(a) z = tan
1 − xy
1
(b) z = p
x2 + y 2
(c) z = cos xy
p
(d) z = ln x2 + y 2
x+y+z ∂u ∂u ∂u
4. If u = p , show that x +y +z =0
x2 + y 2 + z 2 ∂x ∂y ∂z
∂z ∂z
5. If z = ln (ex + ey ) show that + =1
∂x ∂y
∂ 2z ∂ 2z
6. Show that the equation + = 0 is satisfied by
∂x2 ∂y 2
1 y
z = ln x2 + y 2 + tan−1
p
2 x
∂z ∂z
7. If z = f yz , show that x
+y =0
∂x ∂y
8. If z = z(u, v) where u = u(x, y), v = v(x, y), prove that
∂z ∂z ∂u ∂z ∂v
(a) = +
∂x ∂u ∂x ∂v ∂x
∂z ∂z ∂u ∂z ∂v
(b) = +
∂y ∂u ∂y ∂v ∂y
9. If x = ρ cos φ, y = ρ sin φ, prove that if U is a twice differentiable function
of x and y,
∂ 2U ∂ 2U ∂ 2U 1 ∂U 1 ∂ 2U
+ = + +
∂x2 ∂y 2 ∂ρ2 ρ ∂ρ ρ2 ∂φ2
This expression is the basis for solving problems involving small increments and
rates of change. From
∂z ∂z
dz = dx + dy
∂x ∂y
we get
∂z ∂z
∆z ≈ ∆x + ∆y
∂x ∂y
which gives an approximation of the small change in z, ∆z as a result of small
changes in x and y, ∆x and ∆y respectively. We also have
dz ∂z dx ∂z dy
= + ,
dt ∂x dt ∂y dt
dz
where t is time. This expression gives a calculation of the rate of change of z,
dt
dx dy
given the rates of changes and of x and y respectively. If z is a function
dt dt
of only one variable or a function of more than two variables, the formulae are
adapted accordingly.
Small increments
Example 7.18 The dimensions of a rectangular box are measured to be 70cm, 60cm
and 50cm, and each measurement is correct to within 0.2cm. Estimate the largest
possible error when the volume of the box is calculated from these measurements.
Solution
If the dimensions of the box are x, y and h, then its volume is V = xyh
∂V ∂V ∂V
∆V ≈ ∆x + ∆y + ∆h = yh∆x + xh∆y + xy∆h
∂x ∂y ∂h
Here |∆x| ≤ 0.2, |∆y| ≤ 0.2 and |∆h| ≤ 0.2. Thus
Solution
∂L ∂L ∂L ∂L
∆L ≈
∆C + ∆ρ + ∆V + ∆A
∂C ∂ρ ∂V ∂A
1 1 1
= ρV 2 A(±0.01C) + CV 2 A(±0.005ρ) + CρV A(±0.006V ) + CρV 2 (±0.001A)
2 2 2
1 1 1 1
= CρV 2 A(±0.01) + CρV 2 A(±0.005) + CρV 2 A(±0.012) + CρV 2 A(±0.001)
2 2 2 2
1
= CρV 2 A(±0.01 ± 0.005 ± 0.012 ± 0.001)
2
= L(±0.028)
and the largest percentage error in the value of L is 2.8%
Rates of change
Example 7.20 The altitude of a right circular cone is 20cm and is increasing
at 0.1cms−1 . The radius of the base is 10cm and is decreasing at 0.2cms−1 . How
fast is the volume of the cone changing?
Solution
1
The volume of the right circular cone is given by V = πr2 h where r is the radius
3
dh dr
and h is the altitude. = 0.1 and = −0.2
dt dt
dV ∂V dr ∂V dh
= +
dt ∂r dt ∂h dt
2 dr 1 2 dh
= πrh + πr
3 dt 2 dt
2 1
= π(10)(20)(−0.2) + π(100)(0.1)
3 3
= −73.3
and the volume is decreasing at a rate of 73.3cm3 s−1 .
3. A square pyramid reservoir has a base of side 12m and height 18m. It
contains oil that is flowing out at a rate of 2m3 s−1 . At the moment when
the depth of oil is 8m, find the rate at which the depth is decreasing.
Hint:The volume of a square pyramid with a base of side a and height h is
1
given by V = a2 h.
3
4. The perimeter of a rectangle is 25m and one side decreases at a rate of
0.02m/s. At the moment when the side is 25cm, at what rate does the area
change if the perimeter remains constant?
In this notation, the (closed) ith subinterval is [xi−1 , xi ] and ∆xi = xi − xi−1 is its
length, i = 1, 2, · · · , n. We want to define integrability for a bounded function f
on a closed interval [a, b]. Using the boundedness of f , we introduce, as heights
of rectangles, mi = inf {f (x); xi−1 ≤ x ≤ xi } and Mi = sup{f (x); xi−1 ≤ x ≤ xi }
and form n n
X X
the lower sum L(f, P ) = mi (xi − xi−1 ) = mi ∆xi and
i=1 i=1
Xn Xn
the upper sum U (f, P ) = Mi (xi − xi−1 ) = Mi ∆xi .
i=1 i=1
Clearly L(f, P ) ≤ U (f, P ) for any f and P . Any partition of [a, b] that contains
P is called
S a refinement of P . Two partitions P an Q have common refinements;
i.e. P Q is one.
Example 8.1 Suppose [a, b] = [0, 3]. Let P = {0, 1, 2, 3} and Q = {0, 12 , 1, 2, 2 12 , 3}.
Then Q is a refinement of P .
81
82 Integration
Now, observe that since f is bounded on [a, b], ∃ numbers m and M such that
m ≤ f (x) ≤ M ∀ x ∈ [a, b]. Thus m ≤ mi (f ) ≤ Mi (f ) ≤ M, ∀ i = 1, 2, · · · , n.
So
n
X n
X n
X n
X
m(b − a) = m∆xi ≤ mi (f )∆xi ≤ Mi (f )∆xi ≤ M ∆xi = M (b − a)
i=1 i=1 i=1 i=1
In other words
Proof 8.1 We shall show that under refinement, lower sums increase (wide
sense) and upper sums decrease. This gives the conclusion since then
[ [
L(f, P ) ≤ L(f, P Q) ≤ U (f, P Q) ≤ U (f, Q)
Definition 8.2 The bounded function f is Riemann integrable on [a, b] if, for
every ε > 0, ∃ a partition P such that U (f, P ) − L(f, P ) < ε. Now, Let
Z b
f dx = supP ∈P L(f, P )
a
Integration 83
and Z b
f dx = infP ∈P U (f, P )
a
Z b Z b
Then f dx is called the Lower Riemann integral and f dx is called the
a a
upper Riemann integral.
Z b Z b
Theorem 8.1 f dx ≤ f dx and equality holds if and only if f is Riemann
a a
integrable on [a, b].
S
Proof 8.2 Let P1 and P2 be any two partitions of [a, b]. Let P = P1 P2 and so
P is a refinement of both P1 and P2 . So
Thus, for any partition P2 of [a, b], U (f, P2 ) is an upper bound of {L(f, P1 ) : P1 ∈
P} where P is the set of all partitions of [a, b]. Hence supP1 ∈P L(f, P1 ) ≤ U (f, P2 ).
Z b Z b
Thus f dx ≤ U (f, P2 ), since f dx = supP1 ∈P L(f, P1 ).
a a
Z b
Since P2 is an arbitrary partition of [a, b] in P, this implies that f dx is a
a
lower bound of {U (f, P2 ) : P2 ∈ P}. Hence
Z b Z b
f dx ≤ infP2 ∈P U (f, P2 ) = f dx
a a
Z b Z b
Thus f dx ≤ f dx.♠
a a
and 2
i−1 i
2 i
Mi (f ) = sup{x ; x ∈ , = }
n n n
and
n 2 n
X i−1 1 1 X 2 1 1 1 1 1
L(f, Pn ) = · = 3 (i−1) = 3 · (n−1)(n)(2n−1) = 1− 2−
i=1
n n n i=1
n 6 6 n n
Z 1
1 1 1 1
and f dx = lim 1− 2− = . Moreover,
0 n−→∞ 6 n n 3
n 2 n
X i 1 1 X 2 1 1 1 1 1
U (f, Pn ) = · = 3 i = 3 · (n)(n+1)(2n+1) = 1+ 2+
i=1
n n n i=1
n 6 6 n n
Z 1 Z 1 Z 1
1 1 1 1 1
and f dx = lim 1+ 2+ = . Hence f dx = f dx =
0 n−→∞ 6 n n 3 0 0 3
and so f is Riemann integrable over [0, 1].
1, if x is rational,
Example 8.4 Let f (x) = Show that for any a < b,
0, if x is not rational
f is not Riemann integrable on [a, b].
Let Pn = {x0 , x1 , · · · , xn } be any partition of a, b. Any subinterval [xi−1 , xi ] con-
tains both rationals and irrationals. Thus mi (f ) = 0 and Mi (f ) = 1 ∀ i =
1, 2, · · · , n.
Xn Xn
L(f, Pn ) = mi (f )∆xi = 0 · (b − a) = 0
i=1 i=1
and n n
X X
U (f, Pn ) = Mi (f )∆xi = 1 · (b − a) = b − a
i=1 i=1
Z b Z b
Thus f dx = 0 and f dx = b − a and f is not Riemann integrable over
a a
[a, b].
i.e. U (f, P )−L(f, P ) < ε which is condition for Riemann integrability. Therefore
f is Riemann integrable.
Proof 8.4 Assume f is increasing and let ε > 0 be given. Let P be a partition
b−a
such that ∆xi = . Since f is monotonic increasing, we have Mi − mi =
n
f (xi ) − f (xi−1 ), so that
n
X
U (f, P ) − L(f, P ) = (Mi − mi )∆xi
i=1
n
X b−a
= (f (xi ) − f (xi−1 )) ·
i=1
n
n
b−aX
= (f (xi ) − f (xi−1 ))
n i=1
b−a
= (f (b) − f (a))
n
< ε
Z b Z b
2. αf is Riemann integrable for every real number α and (αf )dx = α f dx.
Z b Z b Z b a a
f dx.
c
Z b
Theorem 8.6 If f is Riemann integrable on [a, b] and f ≥ 0, then f dx ≥ 0.
a
Corollary 8.1 If f and g are Riemann integrable on [a, b] and f (x) ≤ g(x) ∀ x ∈
Z b Z b
[a, b], then f (x)dx ≤ g(x)dx.
a a
Proof 8.5 Since f is continuous on [a, b], it is integrable on [a, b]. Moreover,
since [a, b] is closed and bounded, it means that f attains its maximum and min-
imum values M and m, respectively on [a, b]. In other words, there are points x0
and y0 in [a, b] such that m = f (x0 ) and M = f (y0 ). Thus
Z b
m(b − a) ≤ f dx ≤ M (b − a)
a
1
Rb
by the corollary above. Therefore m ≤ b−a a
f dx ≤ M . Since f is continuous
on [a, b], it attains all values between m and M on [a, b] and in particular the value
1
Rb 1
Rb
f dx. Hence there is a point ξ on [a, b] such that f (ξ) = f dx ⇒
Zb−a
b
a b−a a
f dx = f (ξ)(b − a). ♠
a
Integration 87
Theorem Z x8.9 Let f be a bounded Riemann integrable function on [a, b] and let
F (x) = f (t)dt ∀ x ∈ [a, b]. Then F is continuous on [a, b].
a
Proof 8.6 Let M = sup{f (x) : x ∈ [a, b]} and let x0 be any point of [a, b]. Now,
let ε > 0 be given. Suppose x ∈ [a, b]. Then
Z x Z x0
|F (x) − F (x0 )| = | f (t)dt − f (t)dt|
a a
Z x
= | f (t)dt|
x0
Z x
≤ |f (t)|dt
x0
≤ M |x − x0 | < ε
ε
provided |x − x0 | < δ = and F is continuous at x0 .♠
M
Z x
Theorem 8.10 Let f be a continuous function on [a, b], and let F (x) = f (t)dt
a
for x ∈ (a, b). Then F is differentiable at x and F 0 (x) = f (x).
Z x
Proof 8.8 Let F be any function whose derivative is f . Then F (x) = f (t)dt+
a
C, C a constant. This is so since all indefinite integrals
Z x of f differ by a
constant and we know from the previous theorem that f (t)dt is an indef-
Z a a
Z b Z b Z d
A(x)dx = f (x, y)dy dx
a a c
Integration 89
The integral on the right side is called an iterated integral. In this way, we
have the two iterated integrals:
Z bZ d Z b Z d
f (x, y)dydx = f (x, y)dy dx
a c a c
and Z d Z b Z d Z b
f (x, y)dxdy = f (x, y)dx dy
c a c a
where the former means that we first integrate with respect to y from c to d and
then with respect to x from a to b and the later means that we first integrate
with respect to x from a to b and then with respect to y from c to d. The two
iterated integrals are always equal.
This theorem gives a practical method for evaluating a double integral, as ex-
plained earlier, by expressing it as an iterated integral.
Now, if f is positive, then the double integral
Z Z
f (x, y)dA
R
can be interpreted as the volume V of the solid S that lies above R and under
the surface Z = f (x, y).
where the innermost integral is to be evaluated first. The integration can also be
performed in any other order to give an equivalent result. This can be extended
to higher dimensions.
90 Integration
√
(x2 + y 2 )dxdy = √
(x2 + y 2 )dx dy
y=1 x= y y=1 x= y
4 2
x3
Z
= + xy 2 dy
y=1 3 √
x= y
3
!
Z 4
8 y2 5
= + 2y 2 − − y2 dy
y=1 3 3
4
8 2 2 5 2 7
= x + y3 − y 2 − y 2
3 3 15 7 1
1006
=
105
Example 8.6 Find the volume of the region R bounded by the parabolic cylinder
z = 4 − x2 and the planes x = 0, y = 0, y = 6, z = 0.
Z Z Z Z 2 Z 6 Z 4−x2
dxdydz = dzdydx
R x=0 y=0 z=0
Z 2 Z 6
= (4 − x2 )dydx
x=0 y=0
Z 2
= (4 − x2 )y|6y=0 dx
Zx=0
2
= (24 − 6x2 )dx
x=0
2
= 24x − 2x3 0
= 32
Integration 91
∂x ∂x
∂(x,y)
where G(u, v) = F {f (u, v), g(u, v)} and ∂(u,v) ≡ ∂u ∂v
∂y ∂y is the Jacobian
∂u ∂v
of x and y with respect to u and v. Similarly if (u, v, w) are curvilinear co-
ordinates in three dimensions, there will be a set of transformation equations
Integration 93
Z Z Z Z Z Z
Thus F (x, y, z)dxdydz = G(r, θ, z)rdrdθdz where G(r, θ, z) =
R R0
F (r cos θ, r sin θ, z)
3. Use a double integral to find the area of one loop of the rose r = cos 3θ.
Z Z p
4. Evaluate x2 + y 2 dxdy where R is the region x2 + y 2 ≤ a2 .
R
Z Z
2 +y 2 )
5. Evaluate e−(x dxdy where R is the region x2 + y 2 ≤ a2 .
R
Z 1 Z 1−z y
6. By using the transformation x+y = u, y = uv, show that e x+y dydx =
x=0 y=0
e−1
2
n
X
Definition 8.4 The limit lim {P (ξk , ηk )∆xk + Q(ξk , ηk )∆yk }, if it exists,
n−→∞
k=1 Z
is called a line integral along C and is denoted by [P (x, y)dx + Q(x, y)dy] or
Z (a2 ,b2 ) C
[P dx + Qdy].
(a1 ,b1 )
The limit in the definition above will exist if the functions P and Q are continuous
(or piecewise continuous) at all points of C. The definition above can be extended
to the definition of a line integral along a curve C in three dimensional space:
Xn Z
lim {A1 (ξk , ηk , ζk )∆xk +A2 (ξk , ηk , ζk )∆yk +A3 (ξk , ηk , ζk )∆zk } = [A1 dx+A2 dy+A3 dz]
n−→∞ C
k=1
represents physically the total work done in moving the object along the curve
C.
How do we evaluate line integrals?
If the equation of a curve C in the plane z = 0 is given as y = f (x), the line
integral in definition 8.4 is evaluated byZ placing y = f (x), dy = f 0 (x)dx in the
a2
integrand to obtain the definite integral [P {x, f (x)}dx + Q{x, f (x)}f 0 (x)dx].
a1
Similarly if C is given as x = g(y), then dx = g 0 (y)dy and the line integral
Z b2
becomes [P {g(y), y}g 0 (y)dy + Q{g(y), y}dy]. If C is given in parametric form
b1
x = φ(t), y = ψ(t), the line integral becomes
Z t2
[P {φ(t), ψ(t)}φ0 (t)dt + Q{φ(t)ψ(t)}ψ 0 (t)dt]
t1
Definition 8.6 If a plane region has the property that any closed curve in it can
be continuously shrunk to a point without leaving the region, then the region is
called simply-connected, otherwise it is called multiply-connected.
Proof 8.9 Refer to the figure. Let the equations of the curves AEB and AF B
be y = Y1 (x) and y = Y2 (x) respectively. If R is the region bounded by C, we
have
Z Z Z b "Z Y2 (x) #
∂P ∂P
dxdy = dy dx
R ∂y x=a y=Y1 (x) ∂y
Z b Z b
Y2 (x)
= P (x, y)|y=Y1 (x) dx = [P (x, Y2 (x)) − P (x, Y1 (x))]dx
x=a a
Z b Z a I
= − P (x, Y1 (x))dx − P (x, Y2 (x))dx = − P dx
a b C
Then I Z Z
∂P
P dx = − dxdy.
C R ∂y
Similarly let the equations of curves EAF and EBF be x = X1 (y) and x = X2 (y)
respectively. Then
Z Z Z f "Z X2 (y) #
∂Q ∂Q
dxdy = dx dy
R ∂x y=e x=X1 (y) ∂x
Z f Z f
X2 (y)
= Q(x, y)|x=X1 (y) dx = [Q(X2 (y), y) − Q(X1 (y), y)]dy
y=e e
Z e Z f I
= Q(X1 (y), y)dy + Q(X2 (y), y)dy = Qdy
f e C
Then I Z Z
∂Q
Qdy = dxdy.
C R ∂x
Adding the two line integrals gives
I Z Z
∂Q ∂P
[P dx + Qdy] = − dxdy
C R ∂x ∂y
Integration 97
Z (1,2)
Example 8.9 Evaluate [(x2 − y)dx + (y 2 + x)dy] along (a). a straight line
(0,1)
from (0, 1) to (1, 2), (b). straight lines from (0, 1) to (1, 1) and then from (1, 1)
to (1, 2), (c). the parabola x = t, y = t2 + 1.
(a). An equation for the line joining (0, 1) and (1, 2) in the xy plane isZy = x + 1.
1
In this way, dy = dx and the line integral becomes, after substituting, [(x2 −
1 x=0
Z 1
2 2 5
(x + 1))dx + ((x + 1)2 + x)dx] = (2x2 + 2x)dx = x3 + x2 = + 1 =
0 3 0 3 3
(b). Along the straight line from (0, 1) to (1, 1), y = 1, dy = 0 and the line integral
Z 1 1
2 1 3 1 2
equals (x − 1)dx = x − x = − 1 = − .
x=0 3 0 3 3
Along the straight line from (1, 1) to (1, 2), x = 1, dx = 0 and the line integral
Z 2 2
2 1 3 14 4 10
equals (y + 1)dy = y + y = − = .
y 3 1 3 3 3
Z 1
(c). Since t = 0 at (0, 1) and t = 1 at (1, 2), the line integral equals [(t2 −
Z 1 t=0
2 2 2 5 3 2
(t + 1))dt + ((t + 1) + t)2tdt] = (2t + 4t + 2t + 2t − 1)dt = 2.
0
Z
2 2
Activity 8.3 If A = (3x − 6yz)i + (2y + 3xz)j + (1 − 4xyz )k, evaluate A · dr
C
from (0, 0, 0) to (1, 1, 1) along the following paths C:
1. x = t, y = t2 , z = t3 .
2. the straight lines from (0, 0, 0) to (0, 0, 1) then to (0, 1, 1), and then to
(1, 1, 1).
3. the straight line joining (0, 0, 0) and (1, 1, 1).
I
Example 8.10 Verify Green’s theorem in the plane for [(2xy − x2 )dx + (x +
C
y 2 )dy] where C is the closed curve of the region bounded by y = x2 and y 2 = x.
The plane curves y = x2 and y 2Z = x intersect at (0, 0) and (1, 1). Along
1
y = x2 , the line integral equals [(2x(x2 ) − x2 )dx + (x + (x2 )2 )d(x2 )] =
Z 1 x=0 Z 0
3 2 5 7 2
(2x + x + 2x )dx = . Along y = x, the line integral equals [(2(y 2 )y −
0 6 y=1
Z 0
17
(y 2 )2 )d(y 2 ) + (y 2 + y 2 )dy] = (4y 4 − 2y 5 + 2y 2 )dy = − The required line
1 15
integral is thus 76 − 17 15
1
= 30
Z Z Z Z
∂Q ∂P ∂ 2 ∂ 2
− dxdy = (x + y ) − (2xy − x ) dxdy
R ∂x ∂y R ∂x ∂y
98 Integration
√
Z Z Z 1 Z x
= (1 − 2x)dxdy = (1 − 2x)dydx
R x=0 y=x2
Z 1 √
Z 1 √
x 3 1
= (y − 2xy)|y=x2 dx = ( x − 2x 2 − x2 + 2x3 )dx =
x=0 0 30
Hence Green’s theorem is verified.
Z
Theorem 8.14 A necessary and sufficient condition for [P dx + Qdy] to be
C
independent of the path C joining any tow given points in a region R is that in
R
∂P ∂Q
=
∂y ∂x
where it is supposed that these partial derivatives are continuous in R.
The condition in the theorem above is also the condition that P dx + Qdy is an
exact differential, i.e. that there exists a function φ(x, y) such that P dx + Qdy =
dφ. In this way if the end points of curve C are (x1 , y1 ) and (x2 , y2 ), the value of
the line integral is given by
Z (x2 ,y2 ) Z (x2 ,y2 )
[P dx + Qdy] = dφ = φ(x2 , y2 ) − φ(x1 , y1 )
(x1 ,y1 ) (x1 ,y1 )
Activity 8.4 Let P (x, y) and Q(x, y) be continuous and have continuous first
partial derivatives at each point of a simply connected region R. Prove that a
necessary and sufficient condition that
I
[P dx + Qdy] = 0
C
Z (4,2)
Activity 8.5 1. Evaluate [(x + y)dx + (y − x)dy] along (a). the parabola
(1,1)
y 2 = x, (b). a straight line, (c). straight lines from (1, 1) to (1, 2) and then
to (4, 2), (d). the curve x = 2t2 + t + 1, y = t2 + 1.
Z
2 2
2. If F = (x − y )i + 2xyj, evaluate F · dr along the curve C in the xy
C
plane given by y = x2 − x from the point (1, 0) to (2, 2).
100 Integration
I
3. Verify Green’s theorem in the plane for [(x2 − xy 3 )dx + (y 2 − 2xy)dy]
C
where C is a square with vertices at (0, 0), (2, 0), (2, 2), (0, 2).
Z (2,1)
4. (a). Prove that [(2xy − y 4 + 3)dx + (x2 − 4xy 3 )dy] is independent of
(1,0)
the path joining (1, 0) and (2, 1). (b). Evaluate the integral in (a).
I
5. (a) Evaluate F · dr where F = yi + (x + z)2 j + (x − z)2 k from (0, 0, 0)
C
to (2, 4, 0) along the straight line y = 2x and along the parabola y =
x2 , z = 0.
(b) Show that the area A bounded by a simple closed curve C is given by
I
A= xdy − ydx.
C
Activity 8.6 If φ = x2 yz 3 and A = xzi − y 2 j + 2x2 yk, find (a). ∇φ, (b). ∇ · A,
(c). ∇ × A, (d). div(φA), (e). curl(φA)
where F = (x2 − z)i + 2xyj − 2xk and V is the closed region bounded by the planes
x = 0, y = 0, z = 0, 2x + 2y + z = 6
Solution
∇ · F = 2x + 2x + 0 = 4x. Thus
Z Z Z Z 3 Z 3−x Z 6−2x−2y
∇ · FdV = 4xdzdydx
V x=0 y=0 z=0
Z 3 Z 3−x
= 4x(6 − 2x − 2y)dydx
x=0 y=0
Z 3
3−x
24xy − 8x2 y − 4xy 2
= 0
dx
Zx=0
3
24x(3 − x) − 8x2 (3 − x) − 4x(3 − x)2 dx
=
Zx=0
3 3
(36x − 24x2 + 4x3 )dx = 18x2 − 8x3 + x4 0 = 27
=
0
i j k
∂ ∂ ∂
∇×F =
∂x ∂y ∂z
x2 − z 2xy −2x
∂(x2 − z) ∂(−2x) ∂(2xy) ∂(x2 − z)
∂(−2x) ∂(2xy)
= − i+ − j+ − k
∂y ∂z ∂z ∂x ∂x ∂y
= (0 − 0)i + (−1 + 2)j + (2y − 0)k = j + 2yk
Integration 103
and therefore
Z Z Z Z 3 Z 3−x Z 6−2x−2y Z 3 Z 3−x Z 6−2x−2y
∇×FdV = j dzdydx +k 2ydzdydx
V x=0 y=0 z=0
| {z } | x=0 y=0 z=0
{z }
I1 I2
Z 3 Z 3−x Z 3 3−x
6y − 2xy − y 2 0 dx
I1 = j (6 − 2x − 2y)dydx = j
x=0 y=0 0
3 3
x3
Z
2 2
= j (9 − 6x + x )dx = j 9x − 3x +
0 3 0
= j(27 − 27 + 9) = 9j
Z 3 Z Z 3 Z 3−x
3−x
I2 = k y = 0 2y(6 − 2x − 2y)dydx = k (12y − 4xy − 4y 2 )dydx
x=0 0 0
Z 3 3−x Z 3
4 2
= k 6y 2 − 2xy − y 3 dx = k (18 − 18x + 6x2 − x3 )dx
0 3 0 0 3
3
2 81 27
= k 18x − 9x2 + 2x3 − x4 = (54 − 81 + 54 − )k = k
12 0 6 2
Z Z Z
27
Hence ∇ × FdV = 9j + k
V 2
Z Z
Activity 8.7 State the divergence theorem. Hence evaluate r · ndS where
S
S is a closed surface.
Z Z Z
∂A3 ∂A2
[A1 dx + A2 dy + A3 dz] = − cos αdS
C S ∂y ∂z
Z Z Z Z
∂A1 ∂A3 ∂A2 ∂A1
+ − cos βdS + − cos γdS
S ∂z ∂x S ∂x ∂y
Stoke’s theorem states that the line integral of the tangential component of a
vector A taken around a simple closed curve C is equal to the surface integral of
the normal component of the curl of A taken over any surface S having C as a
boundary. The orientation of C is determined from the orientation of S according
to the right hand rule.
Example 8.13 Show that Green’s theorem is a special case of Stoke’s theorem.
Let A = P i + Qj be a vector point function which is continuously differentiable
in a simply connected region S in the xy-plane whose boundary C is a piecewise
smooth simple closed curve. Now,
j i k
∂ ∂ ∂
∇×A =
∂x
∂y ∂z
P
Q 0
∂Q ∂P ∂Q ∂P
= − i+ j+ − k
∂z ∂z ∂x ∂y
∂Q ∂P
and n · (∇ × A) = k · (∇ × A) = − since the unit normal vector to the
∂x ∂y
xy-plane is k. Also
Example 8.14 Verify Stoke’s theorem for F = (2x − y)i − yz 2 j − y 2 zk, where S
is the upper half of the sphere x2 + y 2 + z 2 = 1 and C is its boundary.
The boundary C of S is a circle in the xy plane of radius 1 and centre at the
origin. Let x = cos t, y = sin t, z = 0, 0 ≤ t ≤ 2π be parametric equations of C.
Then
I I
F · dr = [(2x − y)dx − yz 2 dy − y 2 zdz]
C
ZC2π
= (2 cos t − sin t)(− sin t)dt = π
0
Also,
i j k
∂ ∂ ∂
∇×F= =k
∂x ∂y ∂z
2x − y −yz 2 −y 2 z
Integration 105
Then Z Z Z Z Z Z
(∇ × F ) · ndS = k · ndS = dxdy
S S R
since k · ndS = dxdy and R is the projection of S on the xy plane. In this way,
√ √
Z Z Z 1 Z 1−x2 Z 1 Z 1−x2
dxdy = √
dydx = 4
R x=−1 y=− 1−x2 0 0
Z 1√
= 4 1 − x2 dx =π
0
Introduction
You have already acquired knowledge of various methods of integration. You are
urged to review these methods and get ready to use them to solve simple ordinary
differential equations.
In this unit, we introduce the standard form of a first order differential equation:
P (x, y)dx + Q(x, y)dy = 0 (9.1)
In the domain in which Q(x, y) is not zero, equation 9.1 can equivalently be
written as;
dy
= f (x, y) (9.2)
dx
where
P (x, y)
f (x, y) = − (9.3)
Q(x, y)
106
Ordinary differential equations 107
We will discuss some methods which allow us to solve some of the equations of
this form and those equations which reduce to the first order form.
F (x, y, y 0 , y 00 , · · · , y (n) ) = 0,
dy
Example 9.1 1. = ky, k is a constant.
dt
dy xy − y 2
2. = 2 , x2 − xy 6= 0
dx x − xy
p
3. x 1 + (y 0 )2 = 0
dy d5 y
4. x4 + = y + x7
dx dx5
dy
= f (x, y).
dx
Definition 9.2 The order of a differential equation is that of the highest deriva-
tive occurring in the differential equation.
108 Ordinary differential equations
Example 9.2
y 0 + ay 2 = bx2
is a first order differential equation, while
(y 000 )2 − 2xy 000 + y(y 0 )3 = 0
is a third order equation.
Exercises
4 dy d5 y
(a) x + 5 = y + x7
dx dx
d3 y dy
(b) 3
+ x − y = x6
dx dx
0 2 32
(c) [x + (y ) ] = y 00
(d) y 000 + cos y 00 + x2 y 0 = x3
(e) (x0 )2 x000 = x4 x00 + t5 x0
2. Check whether the given function y = φ(x) is a solution of the given o.d.e.
1−x
(a) φ(x) = ; (x2 + 1)y 0 + y 2 + 1 = 0
1+x
d3 y d2 y
(b) φ(x) = e3x ; − 9 =0
dx3 dx2
d2 y dy
(c) φ(x) = xe3x ; 2
−9 = 6e3x
dx dx
(d) φ(x) = ln(−x), x < 0; xy 0 = 1
solely depends on the variable x, g(x), and another function which solely depends
on the variable y, h(y), such that the differential equation then assumes the form:
dy
= g(x) · h(y) (9.4)
dx
dy
= g(x)dx, h(y) 6= 0 (9.5)
h(y)
If the differential equation can be presented in the form 9.1 where
P (x, y) = P1 (x) · P2 (y)
and
Q(x, y) = Q1 (x) · Q2 (y)
then we also consider such a differential equation to be of variables separable
type. At all points (x, y) where the functions P (x, y) and Q(x, y) are defined, the
equation can still be written in the form
M (x)dx + N (y)dy = 0 (9.6)
with
P1 (x) Q2 (y)
M (x) = , N (y) = , Q1 (x) 6= 0, P2 (y) 6= 0
Q1 (x) P2 (y)
The variables in equations 9.5 and 9.6 are said to be separable where 9.5 is the
separated form, hence the name separable variables. Now, suppose that the
function g(x) is continuous in an interval I1 and that the function h(y) is also
continuous and never zero in the interval I2 . Then the solution of 9.5 is given by
Z Z
dy
= g(x)dx + C
h(y)
where C is an arbitrary constant. Note that it is not necessary to introduce a
constant of integration for the first integral since this will be combined with the
one from theZsecond integral to produceZ one constant C.
dy
Let H(y) = and let G(x) = g(x)dx, then
h(y)
H(y) = G(x) + C (9.7)
is the general solution of 9.5. The differential equation of the form
dy
= g(x) (9.8)
dx
dy
is a particular case of = g(x)h(y) where h(y) = 1 and g is continuous in the
dx
dy
interval I. Similarly if g(x) = 1 in = g(x)h(y), we obtain
dx
dy
= h(y) (9.9)
dx
110 Ordinary differential equations
The class of equations 9.8 and 9.9 is called autonomous differential equations.
Autonomous equations are such that the right side of the equations does not
depend explicitly on the independent variable x. In Physics, the independent
variable is usually the time component t.
dy x2
Example 9.3 Show that the equation = is separable and then find an
dx 1 − y2
equation for the integral curves.
Solution
If we rewrite the equation as
dy
−x2 + (1 − y 2 ) = 0,
dx
then it has the form
dy
M (x) + N (y) =0
dx
and therefore separable. Solving:
(1 − y 2 )dy = x2 dx
y3 x3
y− = +C
3 3
Example 9.4 Show that the following differential equation is separable and find
a solution that satisfies the given initial condition.
xy − y
y0 = , y(2) = 1
1+y
Solution
y 0 (y + 1) = xy − y
y 0 (y + 1) = y(x − 1)
0 1
1−x+y 1+ = 0
y
1
(1 − x)dx + 1 + dy = 0
y
Z Z
1
(1 − x)dx + 1+ dy = C
y
1
x − x2 + y + ln |y| = C
2
Now, y(2) = 1, =⇒ 2 − 2 + 1 + 0 = C, =⇒ C = 1 Hence
1
x − x2 + y + ln |y| = 1
2
Ordinary differential equations 111
y 0 = xey−x
number who have not. Suppose that 100 people initiate the rumour and that a
total of 500 people know the rumour after two days. How long will it take for half
the people to hear the rumour?
Solution
100(5000) 5000
Here m = 5000 and R = y(0) = 100, so y(t) = −5000kt
= .
100 + 4900e 1 + 49e−5000kt
Now, we are given that y(2) = 500. We determine k using this.
5000
500 = , =⇒ k = 0.00017
1 + 49e−10000k
Therefore
5000
y(t) =
1 + 49e−0.85t
Now, to determine how long it will take for half the population to hear the
rumour, we solve
5000
2500 =
1 + 49e−0.85t
which gives t = 4.58. Thus, it will take slightly more than 4 12 days for half the
population to hear the rumour.
Exercises
dy
Example 9.8 Prove that the differential equation = (x+y)2 can be reduced to
dx
one of variables separable type by the substitution v = x + y. Obtain the solution
if y(0) = 0.
Solution
The substitution v = x + y allows us to obtain
dv dy
=1+
dx dx
or
dy dv
= −1
dx dx
after we differentiate with respect to x, and the differential equation takes the
form
dv
− 1 = v2
dx
or
dv
= v2 + 1
dx
which is separable. Its general solution is
v = tan(x + C)
which becomes
y = tan(x + C) − x
after back substitution. Now, y(0) = 0 gives the particular solution
y = tan x − x.
y 0 y y y−x
; y = log x−log y; y 0 = +sin and y 0 =
Differential equations like y 0 =
x x x y+x
2
y x
are all homogeneous. However, the equation y 0 = + is not homogeneous.
x y
The form (4.10) suggests the introduction of a new variable by the substitution
Ordinary differential equations 115
y
v = that is, y = vx. This substitution, first used by the Mathematician Leibniz
x
in 1961, transforms the differential equation 9.10 into one of variables separable.
dy dv
= x + v = F (v),
dx dx
or
dv F (v) − v
= , x 6= 0.
dx x
We separate variables and then integrate;
Z Z
dv dx
= , x 6= 0, F (v) 6= v, =⇒ G(v) = ln |x| + C.
F (v) − v x
for which y
G = ln |x| + C.
x
Definition 9.5 A function f (x, y) is said to be a homogeneous function of
degree r with respect to the variables x and y if
where P (x, y) and Q(x, y) are homogeneous functions of the same degree, then
it is a homogeneous differential equation.
y−x
Example 9.9 Solve y 0 = .
y+x
Solution
The differential equation is homogeneous given that:
y
y−x 1− x
y
f (x, y) = = y =F
y+x 1+ x
x
is a homogeneous function of degree r = 0. To determine its solution we use the
substitution y = vx. Thus
dy dv 1−v
=v+x =
dx dx 1+v
which is variable separable. Separating the variables we obtain:
(1 + v)dv dx
2
= , with x 6= 0
1 − 2v − v x
The solution to this equation is x2 − 2xy − y 2 = C
116 Ordinary differential equations
y 2 + x2
Example 9.10 Solve y 0 = , xy 6= 0.
yx
Solution
dy x y
Rewrite the equation in the form = + , with x, y 6= 0. It is clear that this
dx y x
equation is homogeneous. Let y = vx to obtain
dy dv
=v+x
dx dx
which reduces to
dv 1
v+x = + v, v 6= 0,
dx v
or
dv 1
= , vx 6= 0.
dx vx
Separating variables and integrating we obtain
1 2
v − ln |x| = C, x 6= 0.
2
Since y = vx, the general solution of this equation is given by
y 2 = Cx2 + x2 ln x2 .
y
Activity 9.3 Solve (xe x + y)dx − xdy = 0, x 6= 0.
Exercises
The characteristic element for linearity for this equation is in y and y 0 . Linear
differential equations are important given the frequency in which they occur in
applications.
The linear
R equation 9.11 requires an integrating factor of the form
v(x) = e P (x)dx where
a0 (x)
P (x) =
a1 (x)
for which the general solution becomes
Z
R
− P (x)dx
R R f (x)
y=e Q(x)e P (x)dx dx + Ce− P (x)dx ; Q(x) =
a1 (x)
1
Example 9.11 Solve y 0 + y = 2, x 6= 0
x
Solution
1 R 1
From the general form, P (x) = so that v(x) = e x dx = eln x = x. We multiply
x
the differential equation by the integrating factor to get:
0 1
x y + y = 2x
x
or
xy 0 + y = 2x
which is then written as
d
(xy) = 2x
dx
=⇒ yx = x2 + C
or
y = x + Cx−1 , x 6= 0.
Solution
Here P (x) = 1, so that v(x) = ex and multiplying the differential equation by
this integrating factor and integrating we obtain
ex y = (x − 1)ex + C
which simplifies to
y = Ce−x + x − 1.
118 Ordinary differential equations
Activity 9.4 1. Obtain the general solution of the linear differential equation
y 0 + ay = b
y 0 + 2xy = x
xy 0 − 2y = 3x4 , y(−1) = 2.
Example 9.13 Newton’s Law of Cooling states that the rate of change of the
temperature T of an object is proportional to the difference between T and the
(constant) temperature τ of the surrounding medium, called the ambient temper-
ature. Assuming that the temperature of the object is a differentiable function,
the mathematical formulation of Newton’s Law is
dT
∝T −τ
dt
dT
=⇒ = −k(T − τ ), k > 0 constant
dt
dT
=⇒ + kT = kτ
dt
which is a first order linear equation with P (x) = k and f (x) = kτ . We solve
this equation.
R
v(t) = e kdt
= ekt
dT
ekt + ekt kT = kτ ekt
dt
d
=⇒ (ekt T ) = kτ ekt
dt
=⇒ ekt T = τ ekt + C
=⇒ T = τ + Ce−kt
Example 9.14 A cup of coffee is served to you at 185◦ F in a room where the
temperature is 65◦ F . Two minutes later, the temperature of the coffee has dropped
to 155◦ F . How many more minutes would you expect to wait for the coffee to
cool to 105◦ F ?
Ordinary differential equations 119
Solution
Here τ = 65 and T (0) = 185. So by Newton’s Law of cooling
T = τ + Ce−kt
we find C:
185 = 65 + C, =⇒ C = 120.
Thus
T = 65 + 120e−kt
Now, using T (2) = 155, we determine k.
155 = 65 + 120e−2k
This gives k = 0.144. Hence
T = 65 + 120e−0.144t
Finally
105 = 65 + 120e−0.144t
gives t ≈ 7.63 minutes. Thus you should wait for another 7.63 minutes.
Exercises
dB
4. Suppose that a certain population P has a known birth rate and a
dt
dD
known death rate . Then the rate of change of P is given by
dt
dP dB dD
= − .
dt dt dt
dB dD
(a) Assume that = aP and = bP , where a and b are constants.
dt dt
Find P (t) if P (0) = P0 > 0.
(b) Analyze the cases (i) a > b, (ii) a = b, (iii) a < b.
Find lim P (t) in each case.
t→∞
The last equation is called the auxiliary equation or the characteristic equa-
tion for the second order ode. Solving this equation leads to 3 cases of solution:
Case 1: Distinct real roots:
If the roots are m1 and m2 , m1 6= m2 , then we get two solutions y1 = em1 x and
y2 = em2 x . The general solution is thus
y = c1 em1 x + c2 em2 x .
y = c1 e(α+jβ)x + c2 e(α−jβ)x .
and
ejβx − e−jβx = 2j sin βx
Since y = c1 e(α+jβ)x + c2 e(α−jβ)x is a solution, the choices c1 = c2 = 1 and
c1 = 1, c2 = −1 give, in turn, 2 solutions:
y1 = e(α+jβ)x + e(α−jβ)x = eαx ejβx + e−jβx = 2e αx
cos βx
−jβx
(α+jβ)x (α−jβ)x αx jβx
and y2 = e −e =e e −e = 2jeαx sin βx
We have thus, without loss of generality, shown that the real functions eαx cos βx
and eαx sin βx are solutions of the second order equation under this case 3. Con-
sequently, the general solution is
d2 y dy
1. 2 2
− 5 − 3y = 0
dx dx
d2 y dy
2. 2
− 10 + 25y = 0
dx dx
122 Ordinary differential equations
3. y 00 + y 0 + y = 0
Solution
The solution process involve giving the auxiliary equations, the roots and the
corresponding general solutions.
1
1. 2m2 − 5m − 3 = 0 =⇒ (2m + 1)(m − 3) = 0, =⇒ m1 = − , m2 = 3 and
x
2
y = c1 e− 2 + c2 e3x
Activity 9.5 1. Determine the general solution of the given differential equa-
tion:
(a) y 00 − y 0 − 6y = 0
d2 y dy
(b) 2
+ 8 + 16y = 0
dx dx
00 0
(c) y + 3y − 5y = 0
(d) 12y 00 − 5y 0 − 2y = 0
d2 y dy
(e) 2
− 3 + 2y = 0
dx dx
00
(f ) y + 9y = 0
(g) 4y 00 + y 0 = 0
2. Solve the given differential equation subject to the indicated initial condi-
tions: