Multivariable Functions Fields and Vector Calculus Notes 2020 PDF
Multivariable Functions Fields and Vector Calculus Notes 2020 PDF
1 Engineering Mathematics 3
1.1 What is engineering mathematics? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Why study the engineering mathematics? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 How hard are the engineering mathematical modules? . . . . . . . . . . . . . . . . . . . . . . 4
1.4 About this lecture note . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.5 Useful maths websites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3 Multiple integrations 27
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.2 Double integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.3 Triple integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4 Vector calculus 37
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.2 Partial differentiation of vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.3 The gradient of a scalar field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.3.1 Physical interpretation of ∇φ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
4.4 The divergence of a vector field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.4.1 Physical interpretation of ∇ · v . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.5 The curl of a vector field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.5.1 Physical interpretation of ∇ × v . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.5.2 Conservative fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.5.3 The potential function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.6 Examples involving grad, div and curl . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.7 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
1
2 CONTENTS
6 Fourier Series 63
6.1 Joseph Fourier (1768-1830) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
6.2 Fourier series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
6.3 Periodic functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
6.4 Fourier coefficients for f (x) with period 2π . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
6.4.1 Determination of the constant term a0 . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
6.4.2 Determination of the coefficients an of the cosine terms . . . . . . . . . . . . . . . . . 66
6.4.3 Determination of the coefficients bn of the sine terms . . . . . . . . . . . . . . . . . . . 66
6.4.4 Summary of Fourier coefficients and Fourier series . . . . . . . . . . . . . . . . . . . . 67
6.5 Fourier coefficients for f (x) with any period p = 2L . . . . . . . . . . . . . . . . . . . . . . . . 68
6.6 Complex Fourier series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
6.7 Sine and Cosine series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
6.8 Differentiation and integration of Fourier series . . . . . . . . . . . . . . . . . . . . . . . . . . 73
6.8.1 Integration of a Fourier series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
6.8.2 Differentiation of a Fourier series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
6.9 Some applications for the Fourier series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
6.9.1 Frequency response of a linear system . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
6.9.2 Parseval’s theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
6.9.3 Fourier series and Ordinary differential equations . . . . . . . . . . . . . . . . . . . . . 77
6.10 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
8 Appendixes 95
A Formulas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
Chapter 1
Engineering Mathematics
3
4 CHAPTER 1. ENGINEERING MATHEMATICS
an approximation formula ⇒ a lecture is a kind of communication between the lecturer and students
Of course, I will alter things in the lecture whenever I think it necessary if a change will improve communi-
cation and you may choose to write down things from the lecture (or tutorial) that I have not included in
the note.
Experience shows that very few people are able to use lecture notes as a substitute for lectures. If it
were, otherwise, lecturing, as a profession, would have died out by now. So, you should bear in mind that
this note is intended to complement the experience you get from attending the lectures and is available to
supplement the note that you take in the lectures. There is significant value in taking your own notes, you
are much more likely to see what is going on as you do so. I hope that by having this version of the note as
well you can quickly correct any errors that occur as you take them.
At the end of each chapter is a set of questions. You should make sure you have a proper attempt at each
question before looking at the solution, which I hope will be helpful if you are having difficulties. However,
you should be aware that exam solutions are, of course, not made available always. Many things can cause
the solution to be wrong. One such is when the question is changed but the answer is not updated. In the
end you are responsible for ensuring that you understand the solutions and believe them to be correct. If
you have problems in doing so, please seek help.
A PDF (Portable Document Format) version is available from the Blackboard. The file can then be
viewed using Acrobat reader (freely available) or may well display directly in the browser. In either case,
the whole document or selected pages can be printed.
2. http://en.wikipedia.org/wiki/Catergory:Mathematics
1.5. USEFUL MATHS WEBSITES 5
3. http://helm.lboro.ac.uk/
(a very good website to help engineers learn mathematics)
4. http://www.mathcentre.ac.uk/
(which offers students quick reference guides, practice and revision materials, and also online practice
exercises on many branches of mathematics)
There are, of course, hundreds of other websites to explore . . . , but, please don’t spend too much time
on surfing the internet.
6 CHAPTER 1. ENGINEERING MATHEMATICS
Chapter 2
2.1 Introduction
Functions with two or more independent variables appear more often in science and engineering than func-
tions of a single variable, and their calculus is even richer. Their derivatives are more varied and more
interesting because of the different ways in which the variables can interact. Their integrals lead to a greater
variety of applications. The studies of mechanical and electronic engineering, fluid dynamics, to mention only
a few, all lead in natural ways to functions of more than one variable. The mathematics of these functions
is one of the finest achievements in science.
As we see in this chapter, the rule of calculus remain essentially the same as we move into higher
dimensions. We need to keep track of multiple directions of change at the same time, necessitating some new
notations that use the vector notations, but fortunately we do not need to reinvent the theories. Indeed, the
calculus of several variables is really single variable calculus applied to several variables at once.
z = f (x, y)
7
8 CHAPTER 2. MULTIVARIABLE FUNCTIONS AND THEIR DERIVATIVES
where z is called the image of the ordered pair (x, y). In this case, x and y are called the independent
variables of f , and z is the dependent variable.
It is easy to extended to n-variables (x1 , x2 , · · · , xn ), thereby defining a function of n variables in the
obvious way
z = f (x1 , x2 , · · · , xn )
Suppose we consider a function with two independent input variables x and y, for example
f (x, y) = x2 + 2y + 3
If we specify values for x and y then we have a single value f (x, y). For example, if x = 2 and y = 3 then
f (x, y) = 22 + 2 × 3 + 3 = 13. We write f (2, 3) = 13.
Example 2.2.1 Find the values or formula of f (2, 1), f (−1, −3) and f (0, 0), for the following functions:
(a) f (x, y) = x + y
(b) f (x, y) = x2 + y 2 + 1
(c) f (x, y) = 2x + xy + y 3
Solution:
f (x, y, z) = x2 + y 2 + z 2
p
Figure 2.1: Surface of the function f (x, y) = x2 + y 2 with some values of k.
p
Figure 2.2: A contour plot of the function f (x, y) = x2 + y 2 with some values of k.
and
p
sin( x2 + y 2 )
f (x, y) = p , −10 ≤ x ≤ 10, −10 ≤ y ≤ 10
x2 + y 2
Definition 2.4.1 A partial derivative of a function of several variables is its derivative with respect to one
of these variables, with others held constant. For example, the partial derivative of f (x, y) with respect to x
(or y) is the (ordinary) derivative of the partial function with respect to x (or y). That is
∂f (x, y) f (x + h, y) − f (x, y)
fx (x, y) = = lim
∂x h→0 h
10 CHAPTER 2. MULTIVARIABLE FUNCTIONS AND THEIR DERIVATIVES
Figure 2.3: A surface plot for the function f (x, y) = sin(x + y) cos(xy) on the domain 0 ≤ (x, y) ≤ 4.
0.8
0.6
0.4
0.2
−0.2
−0.4
10
5 10
5
0
0
−5
−5
−10 −10
p p
Figure 2.4: A surface plot for the function f (x, y) = sin( x2 + y 2 )/ x2 + y 2 .
z = f (x, y) = x2 + xy + y 2 .
The graph of this function defines a surface as shown in Fig.2.5. To every point on this surface, there are an
infinite number of tangent lines. Partial differentiation is the act of choosing one of these lines and finding
its slope. Usually, the lines of most interest are those that are parallel to the xz-plane, and those that are
parallel to the yz-plane.
To find the slope of the line tangent to the function at (1, 1, 3) that is parallel to the xz-plane, the y
variable is treated as constant. The graph and this plane are shown in Fig.2.5. On the graph follows it, i.e,
Fig.2.6, we see the way the function looks on the plane y = 1. By finding the derivative of the equation
while assuming that y is a constant, the slope of f at the point (x, y, z) is found to be
∂z
= 2x + y
∂x
2.4. THE PARTIAL DERIVATIVE 11
Figure 2.5: A graph of z = x2 + xy + y 2 . For the partial derivative at (1, 1, 3) that leaves y constant, the
corresponding tangent line is parallel to the xz-plane.
∂z
So at (1, 1, 3), by substitution, the slope is 3. Therefore, = 3 at the point (1, 1, 3). That is, the partial
∂x
derivative of z with respect to x at (1, 1, 3) is 3.
∂f
In general, measures the rate of change of f (x, y) with respect to x when y is held constant. Similarly,
∂x
∂f
measures the rate of change of f (x, y) with respect to y when x is held constant.
∂y
Consider f (x, y) = x3 + 2x2 y + y 2 + 2x + 1. Suppose we keep y constant and vary x; then what is the
rate of change of the function f ? Suppose we hold y at the value 3, for example, then
In effect, we now have a function of x only. If we differentiate it with respect to x we obtain the expression:
3x2 + 12x + 2
We say that f has been partially differentiated with respect to x when y = 3. We denote the partial derivative
of f with respect to x by
∂f (x, 3)
= 3x2 + 12x + 2
∂x
In the same way if y is held at the value 4 then f (x, 4) = x3 + 8x2 + 2x + 17 and so, for this value of y = 4,
we have
∂f (x, 4)
= 3x2 + 16x + 2
∂x
Now if we return to the original formulation
f (x, y) = x3 + 2x2 y + y 2 + 2x + 1
and treat y as a constant then the process of partial differentiation with respect to x gives
∂f (x, y)
= 3x2 + 4xy + 0 + 2 + 0 = 3x2 + 4xy + 2
∂x
12 CHAPTER 2. MULTIVARIABLE FUNCTIONS AND THEIR DERIVATIVES
The rules for partial differentiation are essentially the same as for one variable differentiation, but it
must always be remembered which variables are being held constant.
Example 2.4.1 Find the first partial derivatives for the following functions:
(i) f (x, y, z) = x + 2y + z 3
(ii) f (x, y, z) = x2 (y + 2z)
x+y
(iii) f (x, y, z) = 3
z +x
Solution:
We have
∂f ∂f ∂f
(i) fx = = 1, fy = = 2, fz = = 3z 2 .
∂x ∂y ∂z
∂f ∂f ∂f
(ii) fx = = 2x(y + 2z), fy = = x 2 , fz = = 2x2
∂x ∂y ∂z
∂f z3 − y ∂f 1 ∂f −3z 2 (x + y)
(iii) fx = = 3 , f y = = , f z = = .
∂x (z + x)2 ∂y z3 + x ∂z (z 3 + x)2
∂f (1, 2) ∂f (1, 2)
Example 2.4.2 If f (x, y) = exy + xy, find and .
∂x ∂y
Solution:
Since,
∂f ∂f
= yexy + y, = xexy + x
∂x ∂y
Hence,
∂f ∂f
= 2e1(2) + 2 = 2e2 + 2, = 1e1(2) + 1 = e2 + 1
∂x (1,2) ∂y (1,2)
Example 2.4.3 The pressure, P , for one mole of an ideal gas is related to its absolute temperature, T , and
specific volume, v, by the equation
P v = RT
where R is the gas constant. Obtain simple expressions for
1. the coefficient of thermal expansion, α defined by
1 ∂v
α=
v ∂T P
Solution:
1. Since
RT ∂v R
v= =⇒ =
P ∂T P P
hence, we have
1 ∂v R 1
α= = =
v ∂T P Pv T
2. Since
RT ∂v RT
v= =⇒ =−
P ∂P T P2
hence, we have
1 ∂v RT 1
κT = − = 2
=
v ∂P T vP P
2.5. CHAIN RULE FOR PARTIAL DERIVATIVES 13
For the most part, partial derivatives are quite easy to compute, once you become adept at treating
variables like constants. The following example shows how to use the quotient rule in the multivariable
functions.
Example 2.4.4 Find the first order partial derivatives for the function
xy
f (x, y) =
x2 + y2
Solution:
From the quotient rule of ordinary calculus (one variable derivative), we have
∂f ∂f ∂u ∂f ∂v ∂f ∂w
= + +
∂y ∂u ∂y ∂v ∂y ∂w ∂y
∂f ∂f ∂u ∂f ∂v ∂f ∂w
= + +
∂z ∂u ∂z ∂v ∂z ∂w ∂z
So, basically what we are doing here is differentiating f with respect to each variable in it and then multiplying
each of these by the derivative of that variable with respect to x, y, z, respectively. The final step is to then
add all this up.
∂f
Example 2.5.1 Find when
∂t
f (x, y) = x2 y
where
x = t3 , y = t4
Solution:
By the chain rule
∂f ∂f ∂x ∂f ∂y
= +
∂t ∂x ∂t ∂y ∂t
Since
∂f ∂x
= 2xy, = 3t2
∂x ∂t
∂f ∂y
= x2 , = t4
∂y ∂t
Hence
∂f
= 2xy(3t2 ) + x2 (4t3 ) = 10t9
∂t
Actually, f (x, y) can be written as an explicit function with the variable t,
∂f ∂
f (x, y) = x2 y = (t3 )2 (t4 ) = t10 = f (t), = (t10 ) = 10t9
∂t ∂t
which gives the same result when the chain rule is used.
14 CHAPTER 2. MULTIVARIABLE FUNCTIONS AND THEIR DERIVATIVES
∂T ∂T
Example 2.5.2 Find and when
∂r ∂θ
T (x, y) = x3 − xy + y 3
and
x = r cos θ, y = r sin θ
Solution:
By the chain rule
∂T ∂T ∂x ∂T ∂y
= +
∂r ∂x ∂r ∂y ∂r
In this example
∂T ∂T
= 3x2 − y, = −x + 3y 2
∂x ∂y
and
∂x ∂y
= cos θ, = sin θ
∂r ∂r
so that
∂T
= (3x2 − y) cos θ + (−x + 3y 2 ) sin θ
∂r
Substituting for x and y in terms of r and θ gives
∂T
= 3r2 (cos3 θ + sin3 θ) − 2r cos θ sin θ
∂r
Similarly,
∂T
= 3r3 (sin θ − cos θ) cos θ sin θ + r2 (sin2 θ − cos2 θ)
∂θ
∂2f
∂ ∂f
= = fxx (x, y)
∂x ∂x ∂x2
∂2f
∂ ∂f
= = fyy (x, y)
∂y ∂y ∂y 2
and
∂2f
∂ ∂f
= = fyx (x, y)
∂y ∂x ∂y∂x
∂2f
We must be careful with the notation here. We use to mean differentiate first with respect to y and
∂x∂y
∂2f
then with respect to x and we use to mean differentiate first with respect to x and then with respect to
∂y∂x
y. However, for most common functions, the mixed derivatives are equal,
∂2f ∂2f
=
∂x∂y ∂y∂x
Example 2.6.1 Given f (x, y) = 3xy 3 − 2xy + sin x, find all second order partial derivatives of f (x, y).
2.7. PARTIAL DIFFERENTIAL EQUATIONS 15
Solution:
∂f
= 3y 3 − 2y + cos x
∂x
∂2f
∂ ∂f ∂
= = (3y 3 − 2y + cos x) = − sin x
∂x2 ∂x ∂x ∂x
∂2f
∂ ∂f ∂
= = (3y 3 − 2y + cos x) = 9y 2 − 2
∂y∂x ∂y ∂x ∂y
∂f
= 9xy 2 − 2x
∂y
∂2f
∂ ∂f ∂
= = (9xy 2 − 2x) = 9y 2 − 2
∂x∂y ∂x ∂y ∂x
∂2f
∂ ∂f ∂
= = (9xy 2 − 2x) = 18xy
∂y 2 ∂y ∂y ∂y
We found that
∂2f ∂2f
= 9y 2 − 2 =
∂y∂x ∂x∂y
Third, fourth and higher order derivatives are found in a similar way to finding second order derivatives.
It is easy to image computing partial derivatives by iterating the process of differentiating with respect to
one variable, while all others are held constant.
Solution:
∂f ∂f
= 2 cos(2r) − 12r3 s2 = −6r4 s
∂r ∂s
∂2f ∂2f
= −4 sin(2r) − 36r2 s2 = −6r4
∂r2 ∂s2
∂2f
= −24r3 s
∂r∂s
The third order derivatives are found by differentiating the second derivatives
∂3f ∂ ∂2f
= = −8 cos(2r) − 72rs2
∂r3 ∂r ∂r2
∂3f
2
∂ ∂ f
= = −72r2 s
∂r2 ∂s ∂r ∂r∂s
∂3f ∂ ∂2f
= = −24r3
∂r∂s2 ∂r ∂s2
∂3f ∂ ∂2f
= =0
∂s3 ∂s ∂s2
∂3f
2
∂ ∂2f
∂ ∂ f
= =
∂r2 ∂s ∂r ∂r∂s ∂s ∂r2
3 2
2
∂ f ∂ ∂ f ∂ ∂ f
2
= 2
=
∂r∂s ∂r ∂s ∂s ∂r∂s
One example is the wave equation. The displacement, u of the wave depends upon time and position.
Under certain assumptions, the displacement of a wave travelling in one dimension satisfies
∂2u ∂2u
2
= c2 2
∂t ∂x
where c is the speed of the wave. This PED is called the one dimensional wave equation. Under prescribed
conditions, the subsequent displacement of the wave can be calculated as a function of position and time.
Example 2.7.1 Verify that
u(x, t) = sin(x + 2t)
is a solution of the one dimensional wave equation
∂2u ∂2u
= 4
∂t2 ∂x2
Solution:
The first order partial derivatives are calculated
∂u ∂u
= cos(x + 2t), = 2 cos(x + 2t)
∂x ∂t
The second partial derivatives are
∂2u ∂2u
= − sin(x + 2t), = −4 sin(x + 2t)
∂x2 ∂t2
Now,
∂2u ∂2u
= −4 sin(x + 2t) = 4(− sin(x + 2t)) = 4
∂t2 ∂x2
Hence, u(x, t) = sin(x + 2t) is a solution of the given wave equation.
Example 2.7.2 Verify that
x
f (x, y) = xy +
y
is a solution of the partial differential equation
∂2f ∂2f
y + 2x = 2x
∂y 2 ∂x∂y
Solution:
The first partial derivatives are
∂f 1 ∂f x
=y+ and =x− 2
∂x y ∂y y
and the second order partial derivatives are
∂2f
∂ ∂f ∂ x
= = x − 2 = 2xy −3
∂y 2 ∂y ∂y ∂y y
∂2f
∂ ∂f ∂ x 1
= = x− 2 =1− 2
∂x∂y ∂x ∂y ∂x y y
Therefore,
1 2x
y(2xy −3 ) + 2x 1 − 2 = 2xy 2 + 2x − 2 = 2x
y y
x
Hence, f (x, y) = f (x, y) = xy + is a solution of the given equation.
y
Another important PDE is Laplace equation, which is used extensively in electrostatics. Under certain
conditions the electrostatic potential in a region is described by a function φ(x, y), which is satisfies Laplace
equation in two dimensions.
∂2φ ∂2φ
+ 2 =0
∂x2 ∂y
This equation is so important that a whole area of applied mathematics, called potential theory, is devoted
to the study of its solution.
2.8. THE STATIONARY POINTS OF A FUNCTION OF TWO VARIABLES 17
the surface. At point A there is a local maximum, at B there is a local minimum, and at C and D there are
what are known as saddle points.
18 CHAPTER 2. MULTIVARIABLE FUNCTIONS AND THEIR DERIVATIVES
At A the surface is at its greatest height in the immediate neighbourhood. If we move on the surface
from A we immediately lose height no matter in which direction we travel. At B the surface is at its least
height in the neighbourhood. If we move on the surface from B we immediately gain height, no matter in
which direction we travel.
The features at C and D are quite different. In some directions as we move away from these points along
the surface we lose height whilst in others we gain height. The similarity in shape to a horse’s saddle is
evident.
In Fig.2.7 at each of the points A, B, C, D the tangent plane to the surface is horizontal at the point of
interest. Such points are thus known as stationary points of the function. In the next section we show how
to locate stationary points and how to determine their nature using partial differentiation of the function
f (x, y).
∂f (x, y) ∂f (x, y)
= 0, =0
∂x ∂y
Note that this is similar to locating and identifying turning points of a function of one variable. Equating
the first order partial derivatives to zero locates the stationary points, however, it does not identify them as
a maximum or minimum point. To distinguish between these various points a test involving second partial
derivative must be made.
Let f (x, y) and its first and second partial derivatives be continuous in a neighbourhood of x0 , y0 ) and
let fx (x0 , y0 ) = fy (x0 , y0 ) = 0, then
• f (x, y) has a local maximum at (x0 , y0 ) if
Actually, there is a pattern for the above description. Consider the array of second derivatives (remember
fxy = fyx ):
fxx fxy
fxy fyy
Then notice that the conditions we have just derived depend on the sign of fxx and
fxx fxy
fxy fyy
the determinant of the array. This is the key. It is a classic result of linear algebra that the sign of the
quadratic form in the above is dependent on such determinant conditions. This is the way in which the idea
is generalized to functions of more than two variables, and although things do become more complicated,
the determinant notation greatly simplifies things.
2.9. MAXIMUM AND MINIMUM POINTS OF A FUNCTION OF TWO VARIABLES 19
f (x, y) = xy − x2 − y 2
Solution:
The first order partial derivatives are found
∂f ∂f
= y − 2x = 0, = x − 2y = 0
∂x ∂y
Solving the above equations, it yields x = 0 and y = 0. The second partial derivatives are found
The second derivative test to identify the stationary point is used. Now
0
f-20 4
-40 2
-4 0
Y
-2
0 -2
X 2
4 -4
f (x, y) = x2 + xy + 3x + 2y + 5
Solution:
The first partial derivative are found
fx = 2x + y + 3 = 0, fy = x + 2 = 0
So
fxx fxy 2 1
fyx = = −1 < 0
fyy 1 0
Example 2.9.3 Consider a closed rectangular box with dimension x, y and z. If the fixed volume of the box
is 1000 cm3 , find the dimensions so that it has minimum total surface area. What conclusion can you draw
from your result?
20 CHAPTER 2. MULTIVARIABLE FUNCTIONS AND THEIR DERIVATIVES
Solution:
The volume
Volume = xyz = 1000
1000
z=
xy
The surface area, A, is given by
A = 2xy + 2yz + 2xz
We need to replace the z term and get
1000 1000
A = 2xy + 2yz + 2xz = 2xy + 2y + 2x
xy xy
2000 2000
=2xy + +
x y
∂A ∂A
=2y − 2000x−2 , = 2x − 2000y −2
∂x ∂y
2y − 2000x−2 =0
2x − 2000y −2 =0
2000
2y = =⇒ x2 y = 1000
x2
From the second equation we have
2000
2x = =⇒ xy 2 = 1000
y2
• if fxx fyy − (fxy )2 > 0 and fxx > 0 then the stationary point is a minimum.
• if fxx fyy − (fxy )2 > 0 and fxx < 0 then the stationary point is a maximum.
1000
z= = 10
xy
So, we can conclude that the minimum surface area of a rectangular box containing a fixed volume of 1000
cm3 is a cube of 10cm × 10cm × 10cm.
2.10. TAYLOR SERIES FOR FUNCTIONS OF TWO VARIABLES 21
h2 ′′ h3 hn (n)
f (x + h) = f (x) + hf ′ (x) + f (x) + f ′′′ (x) + · · · + f (x) + Rn
2! 3! n!
where the remainder term Rn may be written in the Lagrange form
hn+1 (n+1)
Rn = f (x + θh), 0<θ<1
(n + 1)!
h2 ′′ h3 hn (n)
f (h) = f (0) + hf ′ (0) + f (0) + f ′′′ (0) + · · · + f (0) + Rn
2! 3! n!
where
hn+1 (n+1)
Rn = f (θh), 0<θ<1
(n + 1)!
which is also referred to as expansion about the origin.
treated as a function of t alone, for x, y, h, k all regarded as parameters. Clearly, F (0) = f (x, y). Also, by
the conditions imposed on f (x, y), F (t) is a continuous function of t whenever (x + ht, y + kt) lies in the
region D. Also, F ′ (t) exists and is continuous on D.
Example 2.10.1 Show that
∂f ∂f
F (t) = h +k
∂x ∂y
and obtain an expression for F (n) (t).
Solution:
We only have to apply the expression for the total derivative
∂f dα ∂f dβ ∂f ∂f ∂f ∂f
F ′ (t) = + =h +k =h +k
∂α dt ∂β dt ∂α ∂β ∂x ∂y
We can think of this in the operator form
∂ ∂
F ′ (t) = h +k f (x + ht, y + kt)
∂x ∂y
Similarly,
2
∂ ∂
F ′′ (t) = h +k f (x + ht, y + kt)
∂x ∂y
∂2f ∂2f ∂2f
= h2 2 + 2hk + k2 2
∂x ∂x∂y ∂y
22 CHAPTER 2. MULTIVARIABLE FUNCTIONS AND THEIR DERIVATIVES
We can therefore apply Maclaurin theorem for a single variable to F (t) and get
t2 ′′ t3 tn
F (t) = F (0) + tF ′ (0) + F (0) + F ′′′ (0) + · · · + F (n) (0) + Rn
2! 3! n!
If we now return to the original (x, y) variables and put t = 1 then this finally gives us:
For simplicity, we state the theorem for a rectangular region. If f (x, y) and its partial derivatives up to order
n + 1 are continuous throughout an open rectangular region D centred on (x, y), then, throughout D,
∂ ∂
f (x + h, y + k) = f (x, y) + h +k f (x, y)
∂x ∂y
2 n
1 ∂ ∂ 1 ∂ ∂
+ h +k f (x, y) + · · · + h +k f (x, y) + Rn
2! ∂x ∂y n! ∂x ∂y
where n+1
1 ∂ ∂
Rn = h +k f (x + θh, y + θk), 0<θ<1
(n + 1)! ∂x ∂y
Note that this is very similar to Taylor theorem for single variable with an operator (h∂/∂x + k∂/∂y)n .
Example 2.10.2 Let f (x, y) = cos x cos y and (x, y) = (0, 0). Find the second order Taylor series.
Solution:
We have
f (x, y) = cos x cos y
f (0, 0) = 1
fx (0, 0) = − sin x cos y|(0,0) = 0
fy (0, 0) = − cos x sin y|(0,0) = 0
fxx (0, 0) = − cos x cos y|(0,0) = −1
fxy (0, 0) = sin x sin y|(0,0) = 0
fyy (0, 0) = − cos x cos y|(0,0) = −1
Hence,
1 1 1
f (x, y) ≈ 1 + (−1 · x2 − 1 · y 2 ) = 1 − x2 − y 2
2 2 2
Example 2.10.3 Measuring the height of a building
The height h of a building is estimated from (i) the known horizontal distance x between the point of ob-
servation M and the foot of the building and (ii) the elevation angle θ between the horizontal and the line
joining the point of observation to the top of the building (see Fig.2.9). If the measured horizontal distance
is x = 150 m and the elevation angle is θ = 40o , estimate the error in measured building height due to an
error of 0.1o degree in the measurement of the angle of elevation.
Solution:
The variables x, θ and h are related by
h
tan θ = , x tan θ = h (2.10.1)
x
The error in h resulting from a measurement error in θ can be deduced by differentiating equation (2.10.1):
δh
≈ x sec2 θ
δθ
40π 2π
It is given that x = 150 m, the incidence angle θ = 40o can converted to radians, θ = = rad. Then
180 9
the error in angle δθ = 0.1o need to be expressed in radians for consistency of the units in (2.10.3).
0.1π π
So δθ = = rad. Hence, from equation (2.10.3)
180 1800
π
δh = 150 ≈ 0.45
1800 × cos2 (2π/9)
So the error in building height resulting from an error in elevation angle of 0.1o is about 0.45 m.
24 CHAPTER 2. MULTIVARIABLE FUNCTIONS AND THEIR DERIVATIVES
2.11 Exercises
1. Find the first partial derivatives of the following functions
√
2 2 x 3x2 y
(a) z = 2x + 3y (b) z = x y + xy (c) z= (d) z= −
y y x
2 2
+y 2 ) 2
(e) f = xy 2 + ex y
(f ) f = e−(x (g) f = x2 e−y (h) f = x4 y 2 + xy
y
p x
(i) z = x2 + y 2 (j) z = x ln (k) z = ln(2x − 3y) (l) z = sin
x y
∂2φ ∂2φ
+ 2 = −x sin y.
∂x2 ∂y
6. Let f (x, y) = x sin(xy 2 ), find all second order partial derivatives of f (x, y).
7. Show that if w = f (u, v) satisfies the Laplace equation fuu + fvv = 0 and if u = (x2 − y 2 )/2 and v = xy,
then w satisfies the Laplace equation wxx + wyy = 0.
8. In physics, one dimensional wave equation is given by
∂2w 2
2∂ w
= c
∂t2 ∂x2
Show that the following functions are all solution of the wave equation:
(a) w = sin(x + ct)
(b) w = 5 cos(3x + 3ct) + ex+ct
(c) w = f (u), where f is a differentiable function of u, and u = a(x + ct), where a is a constant.
9. Find the local extreme values of the function
f (x, y) = xy − x2 − y 2 − 2x − 2y + 4
f (x, y) = 4 − x2 − xy − y 2
11. Given
H(t) = sin(3x − y)
and
1 2
x = 2t2 − 3, y= t − 5t + 1
2
use the chain rule to show that
dH(t) 11 2
= (11t + 5) cos t + 5t − 10
dt 2
2.11. EXERCISES 25
12. Given p
φ= x2 + y 2
(a) show
∂2φ 3
2
= y 2 (x2 + y 2 )− 2
∂x
(b) verify that φ is a solution of
∂2φ ∂2φ 1
+ 2 =p
∂x2 ∂y x + y2
2
find
∂4Ω ∂4Ω ∂4Ω
∇4 Ω = + 2 +
∂x4 ∂x2 ∂y 2 ∂y 4
If ∇4 Ω = 0 then Ω(x, y) is called the Airy stress function. Test whether the above Ω(x, y) is an Airy
stress function for x 6= 0.
14. The fluid velocity, v, around a sphere of radius r is given by
r3
v = v0 1 + 3
x
Solution:
∂z
1. (a) ∂x = 2, ∂z
∂y
= 3. (b) ∂z
∂x
= 2xy + y 2 , ∂z
∂y
= x2 + 2xy. (c) ∂z
∂x
= 1
y
∂z
,
∂y
= − yx2 . (d) ∂z
∂x
= 6x
y
+
√ 2 2 2 2
y 2 −(x +y 2 )
2
x2
, ∂z
∂y
= −3x
y2
− 1
√ .
2x y
(e) fx = y 2 + 2xyex y , fy = 2xy + x2 ex y , (f) fx = −2xe−(x +y ) , fy = −2ye ,
−y 2 2 2
3 2 4
(g) fx = 2xe , fy = −2x ye , (h) 4x y + y, fy = 2yx + x
−y
∂z ∂z ∂z ∂z ∂z ∂z
2. (a) ∂x = 2 cos x, ∂y = −3 sin y. (b) ∂x = y sin y, ∂y = x sin y + xy cos y, (c) ∂x = cos(x + y), ∂y = cos(x + y).
sin y cos y
∂z
(d) ∂x = − 2 , ∂y = ∂z ∂y x ∂y x ∂y −t ∂y 2 −t ∂y
, (e) ∂x = te , ∂t = e . (f) ∂x = 2xe , ∂t = −x e , (g) ∂x = 6xe − t3 e−x ∂y 2t
∂t
=
x x
2 2t 2 −x ∂y 2x+3t ∂y 2x+3t x y y x
6x e − 3t e , (h) ∂x = 2e , ∂t = 3e , (i) zx = √ , zy = √ , (j) zx = ln( x ) − 1, zy = y , (k)
2 2 2 2 x +y x +y
2 3
zx = 2x−3y , zy = 2x−3y , (l) zx y1 cos( xy ), zy = − yx2 cos( xy ).
4 2
3. (a) − 9 , − 9 , (b) 2, 1, (c) 5.4366, 5.4366 (d) 25, −12
6. fx = sin(xy 2 ) + xy 2 cos(xy 2 ), fxx = 2y 2 cos(xy 2 ) − xy 4 sin(xy 2 ), fxy = fyx = 4xy cos(xy 2 ) − 2x2 y 3 sin(xy 2 ),
fy = 2x2 y cos(xy 2 ), fyy = 2x2 cos(xy 2 ) − 4x3 y 2 sin(xy 2 )
9. a local maximum at (−2, −2), the value of f at this point is f (−2, −2) = 8.
x(y+z)−y 2 −z 2 −x2 +xy+z(y−z)
10. (0, 0) is local maximum. 15. dz dt
= 2tet + t2 et . 16. (a) fx = 2
√
2 2 2
,fy = 2
√
2 2 2
,
(x+y+z) x +y +z (x+y+z) x +y +z
x2 +y 2 +z 2 x2 +y 2 +z 2 x2 +y 2 +z 2
2x cos 2y cos 2(x2 +y 2 ) cos
−x2 +xz+y(z−y) z2 z2 z2
fz = √ . (b) fx = z2
, fy = z2
,fz = − z3
(x+y+z)2 x2 +y 2 +z 2
26 CHAPTER 2. MULTIVARIABLE FUNCTIONS AND THEIR DERIVATIVES
Chapter 3
Multiple integrations
3.1 Introduction
Before starting on multiple integrals let us do a quick review of the definition of a definite integral for
functions of a single variable. First, when working with the integral,
Z b
f (x)dx
a
we think of x as coming from the interval a ≤ x ≤ b. For this integral we can say that we are integrating
over the interval a ≤ x ≤ b. When we derived the definition of the definite integral we first thought of this
as an area problem. We first asked what the area under the curve was, and to do this we broke up the
interval a ≤ x ≤ b into n subintervals of width ∆x and choose a point, x∗i , from each interval as shown in
Fig. 3.1. Each of the rectangles has height of f (x∗i ) and we then use the area of each of these rectangles to
To get the exact area we must take the limit as n goes to infinity and this was also the definition of the
definite integral:
Z b Xn
f (x)dx = lim f (x∗i )∆x
a n→∞
i=0
In a similar manner, the volume under a surface (given by a function of two variables z = f (x, y)) and above
the xy plane can be found by integrating the function z = f (x, y) twice, once with respect to x and once
with respect to y.
We start out by assuming that a function z = f (x, y) is defined on the region R which is a rectangle as
follows
R = [a, b] × [c, d]
27
28 CHAPTER 3. MULTIPLE INTEGRATIONS
This means that the ranges for x and y are a ≤ x ≤ b and c ≤ y ≤ d. We divide up a ≤ x ≤ b into n
subintervals and divide up c ≤ y ≤ d into m subintervals. This will divide up R into a series of smaller
rectangles and from each of these we choose a point (x∗i , yj∗ ). Here is a sketch of this set up as shown in
Fig.3.2. Now, over each of these smaller rectangles we construct a box whose height is given by f (x∗i , yj∗ ) as
shown in Fig.3.3. Each of the rectangles has a base area of ∆A and a height of f (x∗i , yj∗ ), so the volume of
each of these boxes is f (x∗i , yj∗ )∆A. The volume under the surface z = f (x, y) is then approximately,
n X
X m
V ≈ f (x∗i , yj∗ )∆A
i=0 j=0
Here, we have a double sum since we need to add up volumes in both the x and y directions. To get a better
estimation of the volume we will take n and m larger and larger and to get the exact volume we will need
to take the limit as both n and m go to infinity. In other words,
n X
X m
V = lim f (x∗i , yj∗ )∆A
m,n→∞
i=0 j=0
This looks a lot like the definition of the integral of a function of single variable. In fact this is also the
3.2. DOUBLE INTEGRALS 29
definition of a double integral, or more exactly an integral of a function of two variables over a rectangle,
ZZ X m
n X
f (x, y)dA = lim f (x∗i , yj∗ )∆A
R m,n→∞
i=0 j=0
Note the similarities and differences in the notation to single integrals. We have two integrals to denote the
fact that we are dealing with a two dimensional region and we have a differential here as well. Note that
the differential is dA instead of the dx and dy that we are used to seeing. Note that we have the R written
below the two integrals to denote the region that we are integrating over.
Note that one interpretation of the double integral of f (x, y) over the rectangle R is the volume under
the function (surface) f (x, y) (and above the xy-plane),
ZZ
Volume = f (x, y)dA
R
In the next section we start looking at how to actually compute double integrals.
where A(x) is the cross-section area at x. For each value of x, we may calculate A(x) as the integral
Z y=1
A(x) = (4 − x − y)dy (3.2.2)
y=0
which is the area under the curve z = 4 − x − y in the plane of the cross section at x. In calculating A(x), x
is held fixed and the integration takes place with respect to y. Combining equations (3.2.1) and (3.2.2), we
Figure 3.4: To obtain the cross section area A(x) we hold x fixed and integrate with respect to y.
30 CHAPTER 3. MULTIPLE INTEGRATIONS
If we had just wanted to write instructions for calculating the volume, without carrying out any of the
integrations, we could have written
Z 2Z 1
Volume = (4 − x − y)dydx
0 0
This expression says that the volume is obtained by integrating 4 − x − y with respect to y from y = 0 to
y = 1, holding x fixed, and then integrating the resulting expression in x with respect to x from x = 0 to
x = 2.
What would have happened if we had calculated the volume by slicing with planes perpendicular to the
y-axis as shown in Fig.3.5? As a function of y, the typical cross-section area is
Z x=2
A(y) = (4 − x − y)dx = 6 − 2y (3.2.4)
x=0
Figure 3.5: To obtain the cross section area A(y) we hold y fixed and integrate with respect to x.
ZZ
(4 − x − y)dA
R
3.2. DOUBLE INTEGRALS 31
over the rectangle R : 0 ≤ x ≤ 2, 0 ≤ y ≤ 1? The answer is that they both give the same value of the
double integral. The following theorem published in 1907 by Guido Fubini says that the double integral of
any continuous function over a rectangle can be calculated as an integral in either order of integration.
Solution:
This integral is evaluated over the area as shown in Fig.3.6. By Fubini theorem, we have
O 2 x
-1
Figure 3.6: The plot of the integration domain R = [0, 2] × [−1, 1].
ZZ Z 1 Z 2 Z 1 x=2
Z 1
2 2 3
(1 − 6x y)dxdy = (1 − 6x y)dxdy = x − 2x y x=0
dy = (2 − 16y)dy = 4
R −1 0 −1 −1
Example 3.2.2 Compute the following double integral over the rectangle R = [−2, −1] × [0, 1]
ZZ
(x2 y 2 + cos(πx) + sin(πy))dA
R
Solution:
In this example, we integrate with respect to x first,
ZZ Z 1Z −1
2 2
(x y + cos(πx) + sin(πy))dA = (x2 y 2 + cos(πx) + sin(πy))dxdy
R 0 −2
1 −1
1 3 2 1
Z
= x y + sin(πx) + x sin(πy) dy
0 3 π −2
Z 1 1
7 2 7 3 1 7 2
= y + sin(πy) dy = y − cos(πy) = +
0 3 9 π 0 9 π
Note that here we have used only the basic calculus skills.
In the previous examples, we looked at double integrals over rectangular regions. However, the problem
with this is that most of the regions are not rectangular so we need to now look at a double integral over a
bounded non-rectangular region R. There are two types of regions as shown in Fig.3.7.
The double integral for both of these cases are defined in terms of iterated integrals as follows:
32 CHAPTER 3. MULTIPLE INTEGRATIONS
Figure 3.7: Case 1: R = {a ≤ x ≤ b, g1 (x) ≤ y ≤ g2 (x)}, Case 2: R = {h1 (y) ≤ x ≤ h2 (y), c ≤ y ≤ d}.
Example 3.2.3 Evaluate the following integral over the given domain R
Z Z
(x − 2y)dA
R
where R = {0 ≤ x ≤ 1, 0 ≤ y ≤ 1 − x}.
Solution:
Z Z Z 1 Z y=1−x
(x − 2y)dA = (x − 2y)dydx
R 0 y=0
Z 1 y=1−x
Z 1
y 2 y=0 (−1 + 3x − 2x2 )dx
= xy − dx =
0 0
1
3 2 3 2 1
= −x + x2 − x3 = −1 + − =−
2 3 0 2 3 6
Example 3.2.4 Evaluate the following integral over the given region R
ZZ
(4xy − y 3 )dA
R
√
where R is the region bounded by y = x and y = x3 .
Solution:
In this case we need to determine the two inequalities for x and y that we need to do the integral. The best
way to do this is the graph the two curves, which is shown in Fig.3.8. So, from the sketch we can see that
two inequalities are √
0 ≤ x ≤ 1, x3 ≤ y ≤ x
we can now evaluate the integral as
√
Z Z Z 1 Z x
3
(4xy − y )dA = (4xy − y 3 )dydx
R 0 x3
1 √x 1
1 7 2 1
Z Z
= 2xy − y 42
dx = x − 2x7 + x12 dx
0 4 x3 0 4 4
1
7 3 1 8 1 55
= x − x + x13 =
12 4 52 0 156
3.2. DOUBLE INTEGRALS 33
√
Figure 3.8: Integration domain R is bounded by the curves y = x and y = x3 .
Example 3.2.5 Find the area of the region R bounded by y = x and y = x2 in the first quadrant.
Solution:
We sketch the region as shown in Fig.3.9 and calculate the area as
1.0
0.8
0.6
0.4
0.2
1 x 1 1 1
x2 x3
1
Z Z Z Z
x
Area = dydx = [y]x2 dx = (x − x2 )dx = − =
0 x2 0 0 2 3 0 6
Another method:
√ 1
1 y 1 √ 1
√ 2 3/2 1 2 1
Z Z Z Z
y
Area = dxdy = [x]y dy = ( y − y) dy = y − y =
0 y 0 0 3 2 0 6
Example 3.2.6 Calculate
sin x
ZZ
dA
R x
where R is the triangle in the xy-plane bounded by the x-axis, the line y = x, and the line x = 1.
Solution:
If we integrate first with respect to y and then with respect to x, we find
Z 1Z x Z 1 y=x Z 1
sin x sin x
dy dx = y dx = sin xdx = − cos(1) + 1 ≈ 0.46
0 0 x 0 x y=0 0
y=x
x=1
R
O
X
Figure 3.10: The integration domain R is bounded by the line y = x and the line x = 1.
Solution:
What is meant by this expression is
Z x=1 Z y=1 Z z=1
I= (x + y + z)dz dy dx
x=0 y=0 z=0
where, as before, the inner integral is performed first, integrating with respect to z, with x and y being
treated as constants. So,
Z x=1 Z y=1 1 !
z2
I= x + z + yz + dx
x=0 y=0 2 0
Z x=1 Z y=1
1
= (x + y + )dy dx
x=0 y=0 2
Z x=1 1
y2
1
= xy + + y dx
x=1 2 2 0
Z 1 2 1
1 1 x 3
= (x + + )dx = +x =
0 2 2 2 0 2
Solution:
Z 3 Z − 32 x+2 Z 6−2x−3y Z 3 Z − 23 x+2 Z 6−2x−3y
2xdzdydx = 2xdz dydx
0 0 0 0 0 0
Z 3 Z − 23 x+2
= 2x(6 − 2x − 3y)dydx
0 0
3 3
4 3 1 4 8 3
Z
2 2
= x − 8x + 12x dx = x − x + 6x =9
0 3 3 3 0
3.4. EXERCISES 35
3.4 Exercises
1. Evaluate the following integrals
Z 1Z 3 Z 4Z 2 Z 2Z 3
(a) y 3 dxdy (b) xdxdy (c) (x2 y + 1)dxdy
0 0 0 0 1 2
1 3 Z 2Z 1 Z 2Z 4
y
Z Z
xy
(d) dxdy (e) xe dydx (f ) 6xy 2 dxdy
−1 2 x −1 0 1 2
√
Z 2Z y3 Z 9Z y Z 2Z x3p
x 3
(g) e y dxdy (h) x3 ey dxdy (i) x4 + 1dydx
1 y 0 0 0 0
2. Evaluate each of the following double integrals over the indicated rectangles
Z Z Z Z
(a) 6xy 2 dA, R = [2, 4] × [1, 2] (b) (2x − 4y 3 )dA, R = [−5, 4] × [0, 3]
Z Z R
Z Z R
1
(c) xexy dA, R = [−1, 2] × [0, 1] (d) dA, R = [0, 1] × [1, 2]
R (2x + 3y)2
ZZ Z ZR
√
(e) (x2 + y)dA, R : −1 ≤ x ≤ 1, 0 ≤ y ≤ x2 (f ) 4dA, R : x2 + y 2 ≤ 9
ZZ R
ZZ R
4. Sketch the region of integration, reverse the order of integration if necessary, and evaluate the integrals
Z 1Z 1 Z πZ π Z 2Z 2x
sin y
(a) x2 exy dxdy (b) dydx (c) (4x + 2)dydx
0 y 0 x y 0 x2
Z 1/16Z 1/2 Z 1 Z 1 Z 1 Z √ 1−x2
(d) cos(16πx5 )dxdy (e) (x2 + y 2 )dydx (f ) (5x2 y)dydx
0 y 1/4 0 x −1 0
Z 1 Z 1 Z 3Z 9 Z 8Z 2
2 3
p
(g) ey dydx (h) x3 ey dydx (i) √
x4 + 1dxdy
0 x 0 x2 0 3 y
5. Evaluate
1
Z Z
p dA
R (x + 2y)
over the region R: x − 2y ≤ 1 and x ≥ y 2 + 1.
6. Evaluate Z Z
(x2 + y 2 )dxdy
R
8. Find the volume of the region bounded by the paraboloid z = x2 + y 2 and below by the triangle
enclosed by the lines y = x, x = 0, and x + y = 2 in the xy-plane.
36 CHAPTER 3. MULTIPLE INTEGRATIONS
Solution:
3
(b) 8 (c) 10 1 . (d) 0., (e) e2 − e−1 − 3, (f) 84, (g) 1 e4 − 2e, (h) 1 (e729 − 1), (i) 1 ( 173 − 1). 2. (a) 84, (b) −756, (c) e2 − e−1 − 3, (d)
p
1. (a)
4 2 2 12 6
2
− 1 (ln 8 − ln 2 − ln 5). 3. (a) 16, (b) π + 2, (c) 8 ln 8 − 16 + e, (d) π (e) 1 − sin(1) (f) 1/2(e − 1). 4. (a) (e − 1)/2, (b)2, (c) 8, (d) 1/80π,
6 2
(e) 1/3, (f) 2/3 (g) 1/2(e − 1), (h) 1 (e729 − 1), (i) 1 ( 173 − 1). 5. 2/3. 6. 1/6. 7. 1. 8. 16/3. 9. (a) 1, (b) 0, (c) π 3 /2(1 − cos 1), (d) 1. ,10.
p
12 6
(a) 14 (e − 1), (b) 1
3
Chapter 4
Vector calculus
4.1 Introduction
In many practical problems it is necessary to measure the rate of change of a scalar point function. For
example, in heat transfer problems we need to know the rate of change of temperature from point to point,
because that determines the rate at which heat flows. Similarly, if we are investigating the electric field due
to static charges, we need to know the variation of electric potential from point to point. To determine such
information, the ideas of calculus were extended to vector quantities. The first development of this was the
concept of the gradient of a scalar point function.
A vector field is a function that assigns a vector to each point at given space. Vector fields arise in
differential equations and differential geometry. Vector field is an important topic in multivariable calculus.
Two vector fields are illustrated as below in Fig.4.1:
-1
-2
-3
(a) -3 -2 -1 0 1 2 3 , (b)
Figure 4.1: Vector fields (a) F = (sin2 (x) + y) i + cos(x + y 2 ) j and (b) F = −y i − z j + x k .
37
38 CHAPTER 4. VECTOR CALCULUS
Solution:
∂v
= 6xy i + 2yz j − 12x3 y 2 k
∂x
∂v
= 3x2 i + 2xz j − 6x4 y k
∂y
∂v
= 0 i + 2xy j − 0 k = 2xy j
∂z
∂2v
= 6y i + 0 j − 36x2 y 2 k = 6y i − 36x2 y 2 k
∂x2
∂2v
= 2y j
∂x∂z
In the following sections, we study the three derivatives, that is: (i) the gradient of a scalar field (ii) the
divergence of a vector field and (iii) the curl of a vector field.
Figure 4.2: Temperature contours and heat flow lines for a metal plate.
edges and surfaces are insulated. After a while a steady state situation exists in which the temperature φ at
any point remains the same. Some temperature contours are shown in Fig.4.2.
The direction in which φ changes fastest is along the line of greatest slope which is orthogonal (i.e.
perpendicular) to the contours. Hence, at each point of a scalar field we can define a vector field giving the
magnitude and direction of the greatest rate of change of φ locally.
A vector field, called the gradient, written grad φ, can be associated with a scalar field φ so that at
every point the direction of the vector field is orthogonal to the scalar field contour. This vector field is the
direction of the maximum rate of change of φ. As an example, the direction of the heat flow is along the flow
lines which are orthogonal to the contours (see the dashed lines in Fig.4.2(b)); this heat flow is proportional
to the vector field grad φ.
Definition 4.3.1 Given a scalar filed function of φ = φ(x, y, z) then the vector
∂φ ∂φ ∂φ
i + j + k = ∇φ = grad φ
∂x ∂y ∂z
Solution:
(a)
∂ ∂ ∂
grad φ = ∇ φ = (xy 2 z 3 ) i + (xy 2 z 3 ) j + (xy 2 z 3 ) k = y 2 z 3 i + 2xyz 3 j + 3xy 2 z 2 k
∂x ∂y ∂z
(b)
∂φ ∂φ ∂φ
grad φ = ∇ φ = i + j + = 2x i + 6y j + 4z k
∂x ∂y ∂z
Example 4.3.2 If φ = x3 y + xy 2 + 3y, find (1) ∇φ, (2) ∇φ(1, 2, 3), (3) |∇φ| at (1, 1, 1)
Solution:
1.
∂φ ∂φ ∂φ
= 3x2 y + y 2 , = x3 + 2xy + 3, =0
∂x ∂y ∂z
Hence,
∇φ = gradφ = (3x2 y + y 2 ) i + (x3 + 2xy + 3) j + 0 k
2.
∇φ(1, 2, 3) = (3(1)2 (2) + 22 ) i + (13 + 2(1)(2) + 3) j = 10 i + 8 j
3.
∇φ(1, 1, 1) = 4 i + 6 j + 0 k
p √
|∇φ(1, 1, 1)| = |4 i + 6 j + 0 k | = 42 + 62 + 02 = 52
The change in a function φ in a given direction (specified as a unit vector a) is determined from the scalar
product (dot product) (gradφ) · a. This scalar quantity is called the directional derivative. Note that
1. grad φ
1.
∂φ ∂φ ∂φ
grad φ = ∇φ = i + j + k = 2xy 2 z 2 i + 2x2 yz 2 j + 2x2 y 2 z k
∂x ∂y ∂z
grad φ −2 i + 2 j + 2 k 1 1
=p = √ (−2 i + 2 j + 2 k ) = √ (− i + j + k )
|grad φ| (−2)2 + 22 + 22 2 3 3
3 4
3. At (2, 1, −1), grad φ = 4 i + 8 j − 8 k . To find the derivative of φ in the direction of b = i + k
5 5
take the scalar product (vector dot product)
3 4 3 4
(4 i + 8 j − 8 k ) · i + k = 4 × + 0 + (−8) × = −4
5 5 5 5
Example 4.3.4 Find grad f for f (x, y, z) = 3x2 + 2y 2 + z 2 at the point (1, 2, 3). Hence calculate the
1
directional derivative of f at (1, 2, 3) in the direction of the unit vector (2, 2, 1).
3
Solution:
Since
∂f ∂f ∂f
= 6x, = 4y, = 2z
∂x ∂y ∂z
we have that
grad f = ∇f = 6x i + 4y j + 2z k
At the point (1, 2, 3)
grad f (1, 2, 3) = 6 i + 8 j + 6 k
2 2 1
Thus the directional derivative of f at (1, 2, 3) in the direction of the unit vector , , is
3 3 3
2 2 1 34
(6 i + 8 j + 6 k ) · i + j + k =
3 3 3 3
1 34
Therefore, the directional derivative in the direction of the unit vector (2, 2, 1) is .
3 3
This is commonly denoted by ∇ · v (vector dot product). If we use the vector operator notation introduced
in the previous section we have
∂ ∂ ∂ ∂vx ∂vy ∂vz
div v = ∇ · v = i + j + k · (vx i + vy j + vz k ) = + +
∂x ∂y ∂z ∂x ∂y ∂z
Example 4.4.1 Find the divergence of the vector v = (2x − y 2 ) i + (3z + x2 ) j + (4y − z 2 ) k at the point
(1, 2, 3).
Solution:
Here
vx = 2x − y 2 , vy = 3z + x2 , vz = 4y − z 2
∂vx ∂vy ∂vz
= 2, = 0, = −2z
∂x ∂y ∂z
Thus, the divergence of the vector v is
div v = ∇ · v = 2 − 2z
Example 4.4.2 Find the divergence of F = x2 z i − 2y 3 z 3 j + xyz 2 k at the point (1, 0, 3).
Solution:
Solution:
We have
vx = x sin y, vy = y sin x, vz = −z(sin x + sin y)
so that
∂vx ∂vy ∂vz
= sin y, = sin x, = −(sin x + sin y)
∂x ∂y ∂z
Therefore,
∂vx ∂vy ∂vz
∇·v = + + = sin y + sin x − (sin x + sin y) = 0
∂x ∂y ∂z
and hence v satisfies the equation of continuity.
Note that the curl of a vector field is always a vector field. We can write the curl more compactly as
curl v = ∇ × v
Solution:
i j k
∂ ∂ ∂
∇ × v = ∂x ∂y ∂z
x2 yz −2xy yz
∂(x2 yz) ∂(yz) ∂((−2xy) ∂(x2 yz)
∂(yz) ∂(−2xy)
= − i + − j + − k
∂y ∂z ∂z ∂x ∂x ∂y
= (z − 0) i + (x2 y − 0) j + (−2y − x2 z) k = z i + x2 y j − (2y + x2 z) k
Hence,
∇ × v = z i + x2 y j − (2y + x2 z) k
v = (2x − y 3 , 3z + x2 , 4y − z 2 )
Solution:
Since
vx = 2x − y 3 , vy = 3z + x2 , vz = 4y − z 2
so that
i j k
∂ ∂ ∂
curl v = ∂x ∂y ∂z
2x − y 3 3z + x2 4y − z 2
∂(4y − z 2 ) ∂(3z + x2 ) ∂(2x − y 2 ) ∂(4y − z 2 ) ∂(3z + x2 ) ∂(2x − y 3 )
=i − + j − + k −
∂y ∂z ∂z ∂x ∂x ∂y
= i (4 − 3) + j (0 − 0) + k (2x + 3y 2 ) = i + (2x + 3y 2 ) k
∇ × v = (1, 0, 14) = i + 14 k
Given the magnetic field B = B0 x k , find the associated current I. A mathematical statement of this problem
is that we need to evaluate the curl of B.
Solution:
i j k
∂ ∂ ∂
∇ × B = ∂x ∂y ∂z = 0 i − B0 j + 0 k = −B0 j
0 0 B0 x
and so
B0
I=− j
µ0
The current is perpendicular to the field and to the direction of variation of the field.
curl F = 0
44 CHAPTER 4. VECTOR CALCULUS
is a conservative field.
Solution:
We find
i j k
∂ ∂ ∂
∇ × F = ∂x ∂y ∂z
zex sin y zex cos y x
e sin y
= (e cos y − e cos y) i + (ex sin y − ex sin y) j
x x
v = ∇φ = (2y + z) i + 2x j + x k
and
i j k
∂ ∂ ∂
∇ × v = ∂x ∂y ∂z = 0 i + (1 − 1) j + (2 − 2) k = 0
2y + z 2x x
∇φ ⇒ a vector field
∂φ ∂φ ∂φ
gradφ = ∇φ = i + j + k
∂x ∂y ∂z
where
∂ ∂ ∂
∇= i + j + k
∂x ∂y ∂z
2. The divergence of a vector field:
4.6. EXAMPLES INVOLVING GRAD, DIV AND CURL 45
div v = ∇ · v ⇒ a scalar
Example 4.6.1 Verify that ∇×(∇×v) = ∇(∇·v)−∇2 v for the vector field v = 3xz 2 i −yz j +(x+2z) k .
Solution:
i j k
∂ ∂ ∂
∇ × v = ∂x ∂y ∂z = (y, 6xz − 1, 0)
3xz 2 −yz x + 2z
i j k
∂ ∂ ∂
∇ × (∇ × v) = ∂x ∂y ∂z = (−6x, 0, 6z − 1)
y 6xz − 1 0
∂(3xz 2 ) ∂(−yz) ∂(x + 2z)
∇·v = + + = 3z 2 − z + 2
∂x ∂y ∂z
∇(∇ · v) = (0, 0, 6z − 1)
and
∇2 v = (∇2 (3xz 2 ), ∇2 (−yz), ∇2 (x + 2z)) = (6x, 0, 0)
Thus,
∇(∇ · v) − ∇2 v = (−6x, 0, 6z − 1) = ∇ × (∇ × v)
Example 4.6.2 Show for any vector field F = F1 i + F2 j + F3 k that ∇ · (∇ × F) = 0.
Solution:
i j k
∂ ∂ ∂
∇ · (∇ × F) = div ∂x ∂y ∂z
F1 F2 F3
∂F3 ∂F2 ∂F1 ∂F3 ∂F2 ∂F1
= div − i + − j + − k
∂y ∂z ∂z ∂x ∂x ∂y
∂ ∂F3 ∂F2 ∂ ∂F1 ∂F3 ∂ ∂F2 ∂F1
= − + − + −
∂x ∂y ∂z ∂y ∂z ∂x ∂z ∂x ∂y
∂ 2 F3 ∂ 2 F2 ∂ 2 F1 ∂ 2 F3 ∂ 2 F2 2
∂ F1
= − + − + − =0
∂x∂y ∂z∂x ∂y∂z ∂y∂x ∂z∂x ∂z∂y
Here we have assumed mixed partial derivatives are the same, for example,
∂ 2 F3 ∂ 2 F3
=
∂x∂y ∂y∂x
46 CHAPTER 4. VECTOR CALCULUS
4.7 Exercises
1. Given v = 2xi + 3yzj + 5xz 2 k find
∂v ∂v ∂v
(a) , (b) , (c) .
∂x ∂y ∂z
2. Given v = 2i − xyzj + 3x2 zk find
∂v ∂v ∂v ∂2v ∂2v ∂2v
(a) , (b) , (c) , (d) 2
, (e) 2
, (f) .
∂x ∂y ∂z ∂x ∂y ∂z 2
3. If φ = x2 − y 2 − 3xyz, find
(a) ∇φ,
(b) evaluate ∇φ at the point (0, 0, 0),
(c) Is ∇φ the same as ∇(−φ)? Explain your answer.
4. The force in an electrostatic field given by f (x, y, z) has the direction of the gradient. Find ∇f and its
value at point P
x
(a) f = xy, P (3, −4) (b) f= , P (1, 1)
x2 + y2
π
(c) f = ln(x2 + y 2 ), P (4, 2) (d) f = ex cos y, P (1, )
2
1
(e) f=p , P (12, 0, 16) (f ) f = 2x2 + 4y 2 + 9z 2 , P (−1, 2, −4)
(x2 + y2 + z2 )
9. Determine div v and curl v for v = (x − cos yz)i + (y − cos xz)j + (z − cos xy)k.
10. Given p
r = xi + yj + zk, r = |r| = x2 + y 2 + z 2
12. Express each of the following in operator notation using ∇, ∇· and ∇×.
(a) grad(div v).
(b) curl (gradφ).
(c) curl (curl v).
(d) div (curl v).
(e) div (gradφ).
13. The surface water velocity on a straight uniform river 20 metres wide is modelled by the vector
1
v= x(20 − x) j where x is the distance from the west bank (see Fig.4.3).
50
(a) find the velocity v at each bank and at the midstream
(b) find ∇ × v at each bank and at the midstream
14. Let
u = zi + xj + yk, v = (y + z)i + (z + x)j + (x + y)k, f = x + y − z, g = xyz
find the following expressions
16. Find Curl (F) and Div(F) for the given vector function
p p
F = (x2 + y 2 + z 2 ) i + (x2 + y 2 + z 2 ) j + x k
Solution:
2. (a) −yzj + 6xzk (b) −xzj (c) −xyj + 3x2 k, (d) 6zk, (e) 0, (f) 0.
12. (a) grad(div v) = ∇(∇ · v), (b) curl (gradφ) = ∇ × (∇φ), (c) curl (curl v) = ∇ × (∇ × v),
z z x y x y
16. curl(F)= − q i + (−1 + q j + (q − q ) k , div(F)= q + q
x2 +y 2 +z 2 x2 +y 2 +z 2 x2 +y 2 +z 2 x2 +y 2 +z 2 x2 +y 2 +z 2 x2 +y 2 +z 2
48 CHAPTER 4. VECTOR CALCULUS
Chapter 5
5.1 Introduction
This chapter treats integration in vector fields. The mathematics in this chapter is the line integral. The
importance of line integrals lies in their applications. These are the integrals in which we calculate the
work done by variable forces along paths in space and the rates at which fluids flow along curves and across
boundaries. vector integral calculus is very important to the engineer and has many applications in solid
mechanics, in fluid flow, in heat problems and others.
Recall that, here we integrate the function f (x) from x = a along the x-axis to x = b. Now, in a line integral
we shall integrate a given function along a curve in space or in the plane. Hence, curve integral would be a
better name but line integral is standard.
Similar to the standard integral, a line integral for a function f (x, y) along a curve C can be defined as
Z Z Z Z
f (x, y)dx, f (x, y)dy, f (x, y)ds (f (x, y)dx + f (x, y)dy)
C C C C
The subscript C indicates that the integration is along the given curve C. This curve is not restricted to two
dimensions, and may be in as many dimensions as we please. Note that there are three main cases involving
the variable of
R the integration: dx, dy and ds.
The integral RC f (x, y)ds represents
R the area beneath the surface z = f (x, y) but above the curve C. However,
the integrals C f (x, y)dx and C f (x, y)dy represent the projections of this area onto the xz and yz planes,
respectively. These have been illustratedR in Fig.5.1. The last integral
R is a combination of the cases dx and
dy. A particular case of the integral C f (x, y)ds is the integral C ds. This is a means of calculating the
length along a curve, i.e., an arc length.
A line integral is normal evaluated by expressing all variables in terms of one variable. In general,
Z Z Z
f (x, y)ds 6= f (x, y)dx 6= f (x, y)dy
C C C
The technique for evaluating a line integral is to express all quantities in the integral in terms of a single
variable. If the integral is with respect to x or y, then the curve C and the function f (x, y) may be expressed
in terms of the relevant variable x or y. If the integral is with respect to ds, normally all quantities are
expressed in terms of x. For example, consider the integral
Z b
f (x, y)dx, where y = g(x)
a
This can be evaluated in the usual way by first substituting for y in terms of x in the integrand and then
performing the integration
Z b
f (x, g(x))dx
a
49
50 CHAPTER 5. INTEGRATION IN VECTOR FIELDS
Figure 5.1: Representation of a line integral and its projections onto the xz and yz planes.
Clearly the value of the integral will, in general, depend on the function y = g(x). If x and y are given in
terms of a parameter t, then t is used as the integration variable.
from A(1, 0) to B(0, 1) along the curve C that is the portion of x2 + y 2 = 1 in the first quadrant as shown in
Fig.5.2.
B 1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.2 0.4 0.6 0.8 1
A
Solution:
The curve C is the first quadrant of the unit circle as shown in Fig.5.2. On the curve,
p
y = 1 − x2
so that
0 0
1 1
Z Z p
xydx = x 1 − x2 dx = − (1 − x2 )3/2 = −
C 1 3 1 3
Solution:
This is the same as the previous example other than dx being replaced by dy. As this integral concerns only
points along C and the integration is carried out with respect to y, everything may be expressed in terms of
y. On the curve C, we have
√
y = x2 =⇒ x = y
5.2. LINE INTEGRATION 51
b
dx dy dz
Z Z
f dx + g dy + h dz = f (t) + g(t) + h(t) dt
C a dt dt dt
into f (x, y, z), g(x, y, z) and h(x, y, z), obtaining functions of t. Further, substitute
dx dy dz
dx = dt, dy = dt, dz = dt
dt dt dt
into the integrations with respect to only the variable t. In this way, the line integral results in the integral
of a function of t over the range of values of this variable.
Example 5.2.3 Evaluate the line integral
Z
xdx − yzdy + ez dz
C
if C is given by
x = t3 , y = −t, z = t2 , 1≤t≤2
8 6 4
2
0
-2.0
-1.5
-1.0
Figure 5.3: The integral path C: x = t3 , y = −t, z = t2 , 1 ≤ t ≤ 2.
Solution:
Here
f (x, y, z) = x, g(x, y, z) = −yz, h(x, y, z) = ez
52 CHAPTER 5. INTEGRATION IN VECTOR FIELDS
and, on C,
dx = 3t2 dt, dy = −dt, dz = 2tdt
Then
Z
xdx − yzdy + ez dz
C
Z 2
2
= t3 (3t2 ) − (−t)(t2 )(−1) + et (2t) dt
1
2
111
Z
2
= 3t5 − t3 + 2tet dt = + e4 − e
1 4
Solution:
20
15
z=3t
10
0
1
0.5 1
0 0.5
0
−0.5 −0.5
y=sin(t) −1 −1
x=cos(t)
Figure 5.4: The integral path C : x(t) = cos(t), y = sin(t), z = 3t, 0 ≤ t ≤ 2π.
We have
dx dy dz
= − sin(t), = cos(t), =3
dt dt dt
thus Z Z 2π
zdx + xdy + ydz = (−3t sin t + cos2 t + 3 sin t)dt = 6π + π + 0 = 7π
C 0
We use a ds here to acknowledge the fact that we are moving along the curve, C, instead of the x-axis
(denoted by dx) or the y-axis (denoted by dy). Because of the ds this is sometimes called the line integral
of f (x, y) with respect to arc length. Actually, ds can be written as
p
ds = (dx)2 + (dy)2
or s s
2 2
dy dx
ds = 1+ dx, ds = + 1 dy
dx dy
5.2. LINE INTEGRATION 53
So, to compute a line integral with ds we will convert everything over to the x or y or other parametric
equations. The line integral is then
s s
2 2
dy dx
Z Z Z
f (x, y)ds = f (x, y) 1+ dx = f (x, y) + 1 dy
C C dx C dy
where C is the curve y = x2 , starting from x = 0, y = 0 and ending at x = 1, y = 1. This is the same
integral and curve as Example 5.2.2, but the integration is now carried out with respect to ds, the arc length
parameter.
Solution:
As this integral is with respect to x, all parts of the integral can be expressed in terms of x. Along the curve,
y = x2 , we obtain
s 2
2 dy p p
y = x , =⇒ ds = 1 + dx = 1 + (2x)2 dx = 1 + 4x2 dx
dx
This can be done by substituting method. Let u = 1 + 4x2 , then du = 8xdx. When x = 0, u = 1 and when
x = 1, u = 5, hence,
Z 1 Z 5
3 1 1 5
3
x(1 + 4x2 ) 2 dx = u 2 du = 5 2 − 1 ≈ 2.745
0 1 8 20
Note that the result here is different from the result in Example 5.2.2.
where C is the helix given by x = cos t, y = sin t and z = 3t, 0 ≤ t ≤ 4π as show in Fig.5.4,
Solution:
The line integral is
Z Z 4π p
xyzds = 3t cos(t) sin(t) sin2 t + cos2 t + 9 dt
C 0
Z 4π √ Z
√ 3 10 4π
1
= 3t sin(2t) 1 + 9 dt = t sin(2t)dt
0 2 2 0
√ 4π
3 10 1 t √
= sin(2t) − cos(2t) = −3 10 π
2 4 2 0
So, as we can see there really is not too much difference between ds and dx or dy in the line integrals. All
are the basic integration skills.
Although we defined line integrals over a single smooth curve, if C is a piecewise smooth curve, that is
[ [ [
C = C1 C2 · · · Cn
S
Figure 5.5: The integral path C = C1 C2 for Example 5.2.7.
The following steps tell you how to evaluate a work done integral:
1. evaluate F on the curve as a function of t
2. find dr/dt
3. dot product F with dr/dt
4. integrate from t = a to t = b
Example 5.3.1 Find the work done by the force
F = (y − x2 ) i + (z − y 2 ) j + (x − z 2 ) k
over the curve C as shown in Fig.5.6
r(t) = t i + t2 j + t3 k , 0≤t≤1
5.3. WORK DONE BY A FORCE OVER A CURVE IN SPACE 55
1
0.75
0.5
0.25 1
0
0.75
0.5
0.25
0
0.25 0
0.5
0.75
1
Solution:
Step 1: Evaluate F on the curve
F = (y − x2 ) i + (z − y 2 ) j + (x − z 2 ) k
= (t2 − t2 ) i + (t3 − t4 ) j + (t − t6 ) k
= (t3 − t4 ) j + (t − t6 ) k
dr
Step 2: Find
dt
dr d
= (t i + t2 j + t3 k ) = i + 2t j + 3t2 k
dt dt
dr
Step 3: Dot product F with
dt
dr
F· = ((t3 − t4 ) j + (t − t6 ) k ) · ( i + 2t j + 3t2 k )
dt
= (t3 − t4 )(2t) + (t − t6 )(3t2 )
= 2t4 − 2t5 + 3t3 − 3t8
dr
Step 4: Integrate F · from t = 0 to t = 1
dt
Z 1 1
4 5 3 8 2 5 1 6 3 4 1 9 29
Work = (2t − 2t + 3t − 3t )dt = t − t + t − t =
0 5 3 4 3 0 60
3 1
Example 5.3.2 Find the work done by force F = i + j along the curve C traced by a half circle
4 2
r(t) = cos t i + sin t j from t = 0 to t = π as show in Fig.5.7.
1.0
0.8
0.6
0.4
0.2
Solution:
The work done by F is
Z Z π
3 1 3 1
Z
W = F · dr = i + j · dr = i + j · (− sin t i + cos t j )dt
C C 4 2 0 4 2
Z π π
3 1 3 1 3
= − sin t + cos t dt = cos t + sin t = −
0 4 2 4 2 0 2
56 CHAPTER 5. INTEGRATION IN VECTOR FIELDS
The units of work depend on the units of |F| and on the units of distance.
does not intersect itself. Note that any closed curve can be regarded as a union of simple closed curves. We
use the special notation
I I
f (x, y)ds, and F · dr
C C
to denote line integrals of scalar and vector fields, respectively, along closed curves.
Firstly, let us consider the following example.
where
(a) C : x = t, y = 2t, 0 ≤ t ≤ 1; (b) C : x = t, y = 2t2 , 0 ≤ t ≤ 1
as shown in Fig.5.9.
Solution:
(a) Since x′ (t) = 1 and y ′ (t) = 2, then
Z Z 1
(x2 + y 2 )dx + 2xydy = (x(t)2 + y(t)2 )x′ (t) + 2x(t)y(t)y ′ (t) dt
C 0
1 1 1
13t3
13
Z Z
2 2 2
= (t + 4t )(1) + 2t(2t)(2) dt = 13t dt = =
0 0 3 0 3
5.4. THE LINE INTEGRAL AROUND A CLOSED LOOP 57
So in both cases, if the vector field F(x, y) = (x2 + y 2 ) i + 2xy j represents the force moving an object from
13
(0, 0) to (1, 2) along the given curve C, then the work done is . This may lead you to think that work
3
(and more generally, the line integral of a vector field) is independent of the path taken. However, this is not
always the case. The following theorem gives a necessary and sufficient condition for the path independence:
R
Theorem 5.4.1 InH a region R the line integral C F · dr is independent of the path between any two points
in R if and only if C F · dr = 0 for every closed curve C which is contained in R.
Clearly, the above theorem does not give a practical way to determine path independence, since it is impos-
sible to check the line integrals around all possible closed curves in a region. What it mostly does is give
an idea of the way in which line integrals behave, and how seemingly unrelated line integrals can be related
(in this case a specific line integral between two points and all line integrals around closed curves). For a
more practical method for determining path independence, the following result provides a convenient way
to evaluate a line integral in a conservative field.
Theorem 5.4.2 If F is a conservative field, all the following statements are equivalent:
I Z B
∇×F=0 ⇐⇒ F · ds = 0 ⇐⇒ F · ds = f (B) − f (A)
C A
1
0.8
0.6
0.4
0.2
0
Solution:
1. Evaluate F on C1
F = y 2 i + 2xy j = x4 i + 2x3 j
and
ds
= i + 2x j
dx
58 CHAPTER 5. INTEGRATION IN VECTOR FIELDS
Hence,
Z Z Z 1 Z 1
4 3 4 4
F · ds = (x i + 2x j ) · ( i + 2x j ) = (x + 4x )dx = 5x4 dx = [x5 ]10 = 1
C1 0 0
2. Evaluate F on C2
F = y 2 i + 2xy j = x2 i + 2x2 j
and
ds
= i + j
dx
Hence,
Z Z Z 0 Z 0
2 2 2
F · ds = (x i + 2x j ) · ( i + j ) = (3x )dx = 3x2 dx = [x3 ]01 = −1
C2 1 1
3. I Z Z
F · ds = + F · ds = 1 + (−1) = 0
C1 +C2 C1 C2
R
So, we found that the integral F · ds is path independence.
where C is the boundary (traversed counterclockwise) of the region R = {(x, y) : 0 ≤ x ≤ 1, 2x≤ y ≤ 2x} as
show in Fig.5.11.
Solution:
R is the shaded region in Fig.5.11. By Green’s Theorem, for M (x, y) = x2 + y 2 and N (x, y) = 2xy, we have
Figure 5.11: Closed curve C and the region R = {(x, y) : 0 ≤ x ≤ 1, 2x≤ y ≤ 2x}.
Z Z
∂N ∂M
I
(x2 + y 2 )dx + 2xydy = − dA
C R ∂x ∂y
Z Z Z Z
= (2y − 2y)dA = 0dA = 0
R R
5.5. GREEN’S THEOREM 59
We actually already knew that the answer was zero. Recall from Example 5.4.1 that the line integral
I
(x2 + y 2 )dx + 2xydy = 0
C
F(x, y) = (x − y) i + x j
Solution:
We have
M = (x − y) = cos t − sin t
N = x = cos t
dx = d(cos t) = − sin tdt
dy = d(sin t) = cos tdt
∂M ∂M ∂N ∂N
= 1, = −1, = 1, =0
∂x ∂y ∂x ∂y
From the Green’s theorem, we have from the left hand side
I Z t=2π Z t=2π
M dx + N dy = (cos t − sin t)(− sin tdt) + (cos t)(cos tdt)
C t=0 t=0
Z t=2π
= (− sin t cos t + 1)dt = 2π
t=0
where C is the square cut from the first quadrant by the lines x = 1 and y = 1.
Solution:
We can use either form of Green’s theorem to change the line integral into a double integral over the square.
1. Taking M = xy, N = y 2 , and C and R as the square boundary and interior gives
I Z Z Z 1 Z 1
xydy − y 2 dx = (y + 2y)dxdy = 3ydxdy
C 0 0
1 1 1
3 2 3
Z Z
= [3xy]x=1
x=0 dy = 3ydy = y =
0 0 2 0 2
3
I Z Z
−y 2 dx + xydy = (y − (2y))dxdy =
C R 2
60 CHAPTER 5. INTEGRATION IN VECTOR FIELDS
5.7 Exercises
R
1. Evaluate C
3ydx + 2xdy along the straight line C between (1, 1) and (3, 3).
(7x + 3y)dx + 2ydy along the curve C, y = x2 , between (0, 0) and (2, 4).
R
2. Evaluate C
3. Find the line integral C xydx + x2 ydy where C is the rectangular curve from (2, 5) to (4, 5), to (4, 6),
R
(0 ≤ t ≤ 1).
12. For f (x, y) = 2x + y 2 , find the following line integrals
Z Z Z
(a) f (x, y)dx, (b) f (x, y)dy, (c) f (x, y)ds
C C C
x = t, y = t2 , z = t3 , 0≤t≤1
for the vector field F = x2 i + yz j + xy k , along the curve x = cos t, y = sin t, z = t, with 0 ≤ t ≤ 6π.
17. Find the total work done by a force F = x2 i +y 2 j +z 2 k along the curve C : r = cos t i +sin t j +et k
form (1, 0, 1) to (1, 0, e2π ).
18. Find the total work done by a force F = z i + x j + y k along the curve C, which is defined by the
parametric equations
x = t2 , y = t3 , z = t2 , 0 ≤ t ≤ 1.
21. Evaluate Z
I= x2 ydx + (x − 2y)dy
C
over the part of the parabola y = x2 from (0, 0) to (1, 1). Then you do the same integral by using a
different curve C which is defined as x = sin t, y = sin2 t, 0 ≤ t ≤ π2 . You should obtain the same value
I = − 25 , this illustrate the very important point that the line integral is independent of how the curve
is defined.
is path independent in any domain in space and find its value in the integral from A(0, 0, 0) to B(2, 2, 2).
where C is the boundary of the triangle formed by the points (0, 0), (2, 0) (0, 5). Express the line
integral in terms of an appropriate double integral and evaluate this. Verify Green’s theorem.
24. Evaluate I
ey dx + ex dy
C
where C is the boundary of the triangle formed by the lines y = x, y = 5 and x = 0. By converting
this line integral into a double integral verify Green’s theorem in the plane.
Solution:
1. 20.
2
2. 38., 3. 60, 4. 1 − π + π
2 4 4
5. −1149.
7 (a) F = 4yi + 4xj, (b) 64, (c) 64.
8. −70 or 70
9. (a) 1 − e−2π , (b) 1 − e−2π
10. 303. √
√
11. 11 − 1 . 12. (a) 7 , (b) 14 , (c) 7 5. 13. 1 (143/2 − 1). 14. 56 77 . 15. − 161 16. I = − 3π . 18. 3 . 19. 3 i + 4 j + 9 k . 20. 0. 22. 16.
8 e 3 3 3 6 3 10 2 2 2
23. 20. 24. −452.239.
Chapter 6
Fourier Series
63
64 CHAPTER 6. FOURIER SERIES
where an and bn depend upon f (x). This type of series is called a Fourier series. It might not be unexpected
that Fourier series would not converge to a function f (x) if f (x) is continuous over some interval, but the
amazing thing about Fourier series is that f (x) does not even have to be continuous. At the time, it was
incredible that series of continuous functions could converge to a discontinuous function and Fourier’s work
was severely criticized. Nevertheless, Fourier’s work not only survived almost two centuries of mathematical
scrutiny, but has fostered several areas of modern mathematical research.
f (t + P ) = f (x)
for any value of t. The most obvious examples of periodic functions are the trigonometric functions sin t and
cos t, both of which have period 2π (using radian measure as we shall do throughout this chapter) as shown
in Fig.6.2. This follows since
f (t + 2π) = f (t)
6.4. FOURIER COEFFICIENTS FOR F (X) WITH PERIOD 2π 65
1 2π
Here the period is 2, the frequency is and the angular frequency is = π.
2 2
• Triangular wave Here we can conveniently define the function as
−t −π < t < 0
f (t) =
t 0<t<π
f (t + 2π) = f (t)
The first term on the right equals 2πa0 . All the other integrals on the right are zero, as can be readily seen
by integration. Hence, the first term a0 is
π
1
Z
a0 = f (x)dx (6.4.2)
2π −π
Integration shows that the four terms on the right are zero, except for the last term in the first line, which
equals π when n = m. Since in (6.4.3) this term is multiply by am , the right side in (6.4.3) equals am π. Our
second result is
1 π
Z
am = f (x) cos mxdx, m = 1, 2, 3, · · · (6.4.4)
π −π
The first integral is zero. The next integral is of the kind considered before, and is zero for all n = 1, 2, 3, · · · .
For the last integral we obtain
Z π
1 π 1 π
Z Z
sin nx sin mxdx = cos(n − m)xdx − cos(n + m)xdx
−π 2 −π 2 −π
The last term is zero. The first term on the right is zero when n 6= m and is π when n = m. Since in (6.4.5)
this term is multiplied by bm , the right side in (6.4.5) is equal to bm π, and the last result is
π
1
Z
bm = f (x) sin mxdx, m = 1, 2, 3, · · · (6.4.6)
π −π
6.4. FOURIER COEFFICIENTS FOR F (X) WITH PERIOD 2π 67
∞
X
f (x) = a0 + (an cos nx + bn sin nx) (6.4.7)
n=1
where
π
1
Z
a0 = f (x)dx
2π −π
π
1
Z
an = f (x) cos nxdx n = 1, 2, 3 · · ·
π −π
π
1
Z
bn = f (x) sin nxdx n = 1, 2, 3 · · ·
π −π
Example 6.4.1 Obtain the Fourier series of the half-rectified square wave as shown in Fig.6.6.
Solution:
We have the function
1 0<t<π
f (t) = f (t + 2π) = f (t)
0 π < t < 2π
The calculation of the Fourier coefficients is merely straightforward integration using the results already
obtained Z π Z π
1 1 1 1
a0 = f (t)dt = dt = (π) =
2π −π 2π 0 2π 2
Z π Z π π
1 1 1 sin nt
an = f (t) cos ntdt = cos ntdt = =0
π −π π 0 π n 0
π
1 π 1 π
1 cos nt 1
Z Z
bn = f (t) sin ntdt = sin ntdt = − = (1 − cos nπ)
π −π π 0 π n 0 nπ
Some care is needed now here,
1 0 n = 2, 4, 6, · · ·
bn = (1 − cos nπ) = 2
nπ nπ n = 1, 3, 5, · · ·
Fig.6.7 shows the Fourier series solutions with n = 10 and n = 100, respectively.
68 CHAPTER 6. FOURIER SERIES
1.0 1.0
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
-5 5 -5 5
Figure 6.7: The Fourier series solutions with n = 10 (left) and n = 100 (right), respectively.
∞
X nπ nπ
f (x) = a0 + an cos x + bn sin x (6.5.8)
n=1
L L
L
1
Z
a0 = f (x)dx
2L −L
L
1 nπx
Z
an = f (x) cos dx n = 1, 2, 3 · · ·
L −L L
L
1 nπx
Z
bn = f (x) sin dx n = 1, 2, 3 · · ·
L −L L
Example 6.5.1 Determine the Fourier series of f (x) = x for [−l, l) and f (x + 2l) = f (x) outside this
interval (Fig.6.8 (left) shows a periodic function with l = 2).
Solution:
The Fourier coefficients are
1 l nπx 1 l nπx
Z Z
an = f (x) cos dx = x cos dx = 0
l −l l l −l l
1 l nπx 1 l nπx
Z Z
bn = f (x) sin dx = x sin dx
l −l l l −l l
l
(−1)n+1 2l
2 sin(nπx/l) x cos(nπx/l)
= 2 2 2
− = n = 1, 2, 3, · · ·
l n π /l nπ/l 0 nπ
The an = 0 because the integrand is an odd function of x. The Fourier series of f (x) is
∞
2l X (−1)n+1 nπx
f (x) = sin (6.5.9)
π n=1 n l
Fig.6.8 and Fig.6.9 show f (x) and a few partial sums of its Fourier series representation plotted over its
periodic extension interval.
Example 6.5.2 Determine the Fourier series of f (x) = l2 − x2 for (−l, l) and f (x) = f (x + 2l) outside this
interval as show in Figure 6.10 (right) with l = 2.
Solution:
6.5. FOURIER COEFFICIENTS FOR F (X) WITH ANY PERIOD P = 2L 69
2 2
1 1
-10 -5 5 10 -10 -5 5 10
-1 -1
-2 -2
Figure 6.8: The periodic function f (x) = x on the interval (−2, 2) (left) and the Fourier series solution with
n = 3 (right), respectively.
2
2
1 1
-10 -5 5 10 -10 -5 5 10
-1 -1
-2
-2
Figure 6.9: The Fourier series solutions with n = 10 (left) and n = 100 (right), respectively.
l
1 nπx
Z
an = (l2 − x2 ) cos dx
l −l l
l
2 2x cos(nπx/l) (n2 π 2 x2 /l2 − 2) sin(nπx/l)
= +
l (n2 π 2 /l2 ) n3 π 3 /l3 0
4l2 (−1)n+1
= n 6= 0
π2 n2
4l2
a0 =
3
The bn are equal to zero because (l2 − x2 ) sin(nπx/l) is an odd function of x. Thus,
∞
2l2 4l2 X (−1)n+1 nπx
f (x) = + 2 cos
3 π n=1 n2 l
Fig.6.10 (left) shows f (x) = l2 − x2 and the first few partial sums of the Fourier series representation of f (x)
over the periodic interval −2 to 2, where n = 1, 2, 3. Note that the Fourier series represents not only f (x) in
the interval l = −2 to l = 2. but also its periodic extension as well as shown in Fig.6.10 (right).
4
4
3
3
2
2
1 1
-2 2 4 6 8 10 -2 2 4 6 8 10
Figure 6.10: The Fourier series solutions to f (x) = l2 − x2 with n = 1, 2, 3 (left) and n = 10 (right),
respectively, where l = 2.
Frequently in physical problems, x represents time and the function to be expanded as a Fourier series
is a periodic signal. Recall that sin t goes through one cycle as t goes from 0 to 2π ( or from any point t0
to t0 + 2π). Therefore, sin 2πt goes through one cycle as t goes from 0 to 1, and sin 2πνt goes through v
cycles as t goes from 0 to 1. The function sin 2πνt represents a sinusoidal signal with frequency of ν cycles
per second, or ν hertz (Hz). Because there are 2π radians in one cycle, ω = 2πν is the frequency in unit
of radians per second. If a signal has a frequency of ν cycles per second, then the time between successive
maxima or minima is 1/ν seconds per cycle, which means that the period, τ , of the signal is τ = 1/ν, or
τ = 2π/ω.
70 CHAPTER 6. FOURIER SERIES
Example 6.5.3 Let us consider the square wave show in Fig.6.11. We can express f (t) mathematically by
−1 −t0 ≤ t < 0
f (t) =
1 0 ≤ t < t0
1.0
0.5
-6 -4 -2 2 4 6
-0.5
-1.0
Solution:
The period of f (t) is τ = 2t0 and its frequency is ω = 2π/τ = π/t0 . If we write f (t) as
∞
a0 X nπ nπ
f (t) = + an cos t + bn sin t
2 n=1
t0 t0
then
t0
1 nπt
Z
an = f (t) cos dt = 0
t0 −t0 t0
and
t0
1 nπt 2
Z
bn = f (t) sin dt = [1 − (−1)n ] n = 1, 2, 3, · · ·
t0 −t0 t0 nπ
Thus,
∞ ∞
2 X 1 − (−1)n
nπt 4X 1
f (t) = sin = sin(2n − 1)ωt (6.5.10)
π n=1 n t0 π n=1 2n − 1
Fig.6.12 shows a few partial sum of this series. Note the slow convergence due to the denominator being
of order n. Figure 6.12 also shows the partial sum consisting of 1000 terms, showing that it is possible to
represent a square wave by a Fourier series, provided enough terms are taken.
Figure 6.12: The first five partial sums (left) and 1000 terms (right) of the Fourier series of the square wave
in Figure 6.11 plotted against ωt.
Notice that the summation runs from −∞ to ∞. We can determine the cn in equation (6.6.11) by realizing
that {einπx/L } is an orthogonal set over the interval −L to L. Multiply equation (6.6.11) by e−ikπx/L and
integrate from −L to L to obtain
Z L X∞ Z L X∞
f (x)e−ikπx/L dx = cn ei(n−k)πx/L dx = 2Lcn δnk = ck 2L
−L n=−∞ −L n=−∞
thus
L
1
Z
ck = f (x)e−ikπx/L dx (6.6.12)
2L −L
Example 6.6.1 Let us determine the complex Fourier series representation of f (x) = x in [−l, l] with
f (x + 2l) = f (x).
Solution:
l
1
Z
ck = f (x)e−ikπx/l dx
2l −l
l
1 e−ikπx/l
kπx
= −i − 1n
2l −k 2 π 2 /l2 l −l
il (−1)k il
= cos kπ = (k 6= 0)
kπ kπ
1 l
Z
c0 = f (x)dx = 0
2l −l
Therefore,
∞
il X (−1)n inπx/l
f (x) = e
π n=−∞ n
n6=0
The property of even fe (x) and odd fo (x) functions that we shall make use of is
Z L Z L
fe (x)dx = 2 fe (x)dx
−L 0
and Z L
fo (x)dx = 0
−L
72 CHAPTER 6. FOURIER SERIES
In particular,
Z L nπx Z L nπx
fe (x) cos dx = 2 fe (x) cos dx
−L L 0 L
Z L nπx Z L nπx
fo (x) sin dx = 2 fo (x) sin dx
−L L 0 L
Z L nπx Z L nπx
fo (x) cos dx = fe (x) sin dx = 0
−L L −L L
The equation for the Fourier coefficients
L
1 nπx
Z
an = f (x) cos dx =⇒ an = 0 if f (x) is an odd function
L −L L
L
1 nπx
Z
bn = f (x) sin dx =⇒ bn = 0 if f (x) is an even function
L −L L
f (x) = π − x 0≤x<π
Solution:
To express f (x) as a Fourier sine series, we use the odd extension of f (x) (see Fig.6.14)
−π − x −π ≤ x < 0
fo (x) =
π−x 0≤x<π
Figure 6.14: The odd extension of the function defined in Example 6.7.1 plotted against x (the left figure).
The right figure shows the partial sums of its Fourier series representation consisting of 4, 8 and 64 terms,
respectively.
6.8. DIFFERENTIATION AND INTEGRATION OF FOURIER SERIES 73
Example 6.8.1 Integrate term by term the Fourier series obtained in equation (6.5.10) for square wave
−1 −t0 ≤ t < 0
f (t) =
1 0 ≤ t < t0
Solution:
From equation (6.5.10), the Fourier series expansion for
∞
2 X 1 − (−1)n
nπt
f (t) = sin (6.8.13)
π n=1 n t0
We now need to integrate between the limits −t0 and t and, owing to the discontinuity in f (t) at t = 0, we
must consider separately values of t in the intervals −t0 < t < 0 and 0 < t < t0 .
Case i: In the interval −t0 < t < 0, we integrate (6.8.13) term by term to obtain
Z t ∞ Z
4 X t sin(2n − 1)t
(−1)dt = dt
−t0 π n=1 −t0 (2n − 1)
that is
∞ t
4 X cos(2n − 1)t
−(t + t0 ) = −
π n=1 (2n − 1)2 −t0
"∞ ∞
#
4 X cos(2n − 1)t X 1
=− +
π n=1 (2n − 1)2 n=1
(2n − 1)2
It can be shown that
∞
X 2 1
2
= π
n=1
(2n − 1) 8
so that the above simplifies to
∞
1 4 X cos(2n − 1)t
−t = t0 + π − (−t0 < t < 0)
8 π n=1 (2n − 1)2
Case ii: In the interval 0 < t < t0 , we integrate term by term to get
Z 0 Z t ∞ Z
4 X t sin(2n − 1)t
(−1)dt + dt = dt
−t0 0 π n=1 −π (2n − 1)
giving
∞
1 4 X cos(2n − 1)t
t = t0 + π − (0 < t < t0 )
8 π n=1 (2n − 1)2
74 CHAPTER 6. FOURIER SERIES
then, provided that f ′ (t) satisfies the required conditions, its Fourier series expansion is
∞
X
f ′ (t) = (nbn cos nt − nan sin nt)
n=1
In this case the Fourier coefficients of the derived expansion are nbn and nan , so, in contrast to the integrated
series, the derived series will converge more slowly than the original series expansion for f (t).
Example 6.8.2 Consider the process of differentiating term by term the Fourier series of function
Solution:
The Fourier series for f (t) is
∞
π2 X (−1)n cos nt
t2 = +4 , (−π ≤ t ≤ π)
3 n=1
n2
Since f (t) is continuous within and at the end points of the interval −π ≤ t ≤ π, we obtain
∞
X (−1)n+1 sin nt
t=2 (−π ≤ t ≤ π)
n=1
n
which can be confirmed from the Fourier series expansion obtained for the function
Figure 6.15: The response of a linear system to a sinusoidal input is also sinusoidal.
We know that sinusoidal signals can be represented by complex numbers and that an a.c. electrical circuit
can be analyzed using complex number. This is true for linear systems in general. It is possible to define
a complex frequency function, g(iω), where ω is the frequency of the input, G relates the output and the
input of a linear system.
6.9. SOME APPLICATIONS FOR THE FOURIER SERIES 75
If a sine wave of amplitude Ai , is applied to the system then the amplitude, Ao pf the output is given by
Ao = |G(iω)|Ai
vi = iR + vo
vo 1
= (6.9.14)
vi 1 + jωRC
Equation (6.9.14) relates the output of the system to the input of the system. Therefore,
1
G(jω) =
1 + jωRC
It is convenient to convert G(jω) into polar form
1∠0
G(jω) = p
1 + (ωRC)2 ∠ tan−1 ωRC
1
=p ∠ − tan−1 ωRC
1 + (ωRC)2
Therefore,
1
|G(jω)| = p
1 + (ωRC)2
∠G(jω) = − tan−1 ωRC
The amplitude and phase characteristics for the circuit of Fig.6.16 are shown in Fig.6.17. These show the
variation of |G(jω)| and ∠G(jω) with angular frequency ω. Note that the circuit is a low pass filter, it
allows low frequency to pass easily and rejects high frequencies. The cut-off point of the filter, that is the
point at which significant frequency attenuation being to occur, can be varied by changing the values of R
and C. The quantity of RC is usually known as the time constant for the system. Consider the case when
RC = 0.3. Then the above equations reduce to
1
|G(jω)| = √ (6.9.15)
1 + 0.9ω 2
76 CHAPTER 6. FOURIER SERIES
Figure 6.17: Amplitude and phase characteristics for the circuit of Figure 6.16
n=1
ω1 = 1
1
|G(jω1 )| = √ = 0.96
1 + 0.09 × 1
∠G(jω1 ) = − tan−1 0.3 = −16.7o
n=3
ω3 = 3
1
|G(jω3 )| = √ = 0.74
1 + 0.09 × 32
∠G(jω3 ) = − tan−1 0.9 = −42.0o
6.9. SOME APPLICATIONS FOR THE FOURIER SERIES 77
n=5
ω5 = 5
1
|G(jω5 )| = √ = 0.55
1 + 0.09 × 52
∠G(jω5 ) = − tan−1 1.5 = −56.3o
It is clear that high-frequency Fourier components are attenuated and phase shifted more than low-frequency
Fourier components. The effect is to produce a rounding of the rising and falling edges of the square wave
input signal. This is shown in Fig.6.18b. The output signal has been obtained by adding together the
attenuated and phase-shifted output Fourier components. This is possible because the system is linear.
Figure 6.18: (a) The input square wave signal. (b) The output signal.
Solution:
To solve equation (6.9.18) by means of Fourier series, first express f (x) as a Fourier series
∞
a0 X nπx nπx
f (x) = + an cos + bn sin (6.9.19)
2 n=1
l l
where An and Bn are to be determined. We substitute equation (6.9.19) and (6.9.20) into equation (6.9.18)
to obtain
∞
n2 π 2 n2 π 2
A0 X nπx nπx
+ An 1 − 2 cos + Bn 1 − 2 sin
2 n=1
l l l l
∞
(6.9.21)
a0 X nπx nπx
= + an cos + bn sin
2 n=1
l l
Suppose the boundary conditions were y ′ (0) = y ′ (l) = 0. In this case, we would want to express f (x) as a
cosine series because the derivative of cos nπ/l is equal to zero at x = 0 and x = l. Therefore, we would use
the even extension of f (x) in the above example.
6.10. EXERCISES 79
6.10 Exercises
1. Show
T
2nπt
Z
sin dt = 0, for all integers n
0 T
T
2mπt 2nπt
Z
sin cos dt = 0, for all integers n
0 T T
9. Find the Fourier series of the sawtooth wave function as shown in Fig.6.19
and
f (x + 2π) = f (x)
Solutions:
2. f (t) = 1 + 2 sin πt + 2 sin 3πt + 2 sin 5πt + 2 sin 7πt · · · .
2 π 5 3π 5 5π 5 7π 5
3. f (t) = π − 2 cos t − sin t + 1 sin 2t − 2 cos 3t − 1 sin 3t · · · .
4 π 2 9π 3
2
4. f (t) = π + 2π sin t − 4 cos t − π sin 2t + cos 2t + 2π sin 3t − 4 cos 3t · · · .
3 3 9
n(1−cos nπ)
5. f (t) = 2
P∞
n=1 sin nt.
π (n2 −4)
2L2 P 2+(−1)n (n2 π 2 −2)
6. f (x) = − ∞
n=1 sin nπx
π3 n3 L
cos 2nx
7. f (x) = 2 + 4
P∞
π π n=1 2
1 − 4n
8 P
8. f (x) = ∞ n sin 2nx 1 1
n=1 4n2 −1 9. f (x) = π + 2 sin x − 2 sin 2x + 3 sin 3x − · · ·
π
80 CHAPTER 6. FOURIER SERIES
Chapter 7
81
82 CHAPTER 7. PARTIAL DIFFERENTIAL EQUATIONS
Given suitable boundary or initial conditions, solutions are 3D graphs of the dependent variable u against
the independent variables (x, t) for the heat and wave equations, (x, y) for the Laplace equation.
Other Engineering examples of PDE:
• Beam equation, which derived from force balance, describes the vertical vibration u(x, t) of a one-
dimensional beam
∂2u ∂u ∂4u
m 2 +k + EI 4 = q(x, t)
∂t ∂t ∂x
• Cable equation governs the voltage (or current) u(x, t) in a transmission line
∂2u ∂2u ∂u
= LC + (LC + LG) + RGu
∂x2 ∂t2 ∂t
• Reaction diffusion equation: governs the chemical concentration u(x, t) of a reactant that both
diffuses and reacts with other chemicals
∂u ∂2u
= α2 2 + f (u)
∂t ∂x
• Korteweg-de Vries (KdV) equation governs amplitude u(x, t) of dispersive waves on the surface
of water.
∂u ∂u 3 ∂u ∂ 3 u
= + u +
∂t ∂x 2 ∂x ∂x3
• Navier Stokes equation the fundamental force balance equation for the (vector) velocity field u(x, t)
of viscous fluid flow
∂u 1
+ (u · grad) u + ∇p = η∇2 u
∂t ρ
This vector equation has been shown to be remarkably accurate at describing Newtonian fluid flow
both laminar and turbulent.
Because of the rich variety of PDE and the nature of their solutions, in these lectures, we shall give only
the briefest overview of the theory of PDE, what they are used for in Engineering, and how they can be
solved. We shall focus exclusively on single PDE for a scalar dependent variable u that depends on two (or
occasionally three) independent variables.
2 ∂2u 1 ∂u
Example 7.1.1 Show that u = e−2π t sin πx is a solution of the partial differential equation: =
∂x2 2 ∂t
7.1. WHAT ARE PARTIAL DIFFERENTIAL EQUATIONS? 83
Solution:
First we find
∂u 2 ∂u 2
= −2π 2 e−2π t sin πx, = πe−2π t cos πx
∂t ∂x
then
∂2u 2
= −π 2 e−2π t sin πx
∂x2
We see that
∂2u 1 2 1 ∂u
= (−2π 2 e−2π t sin πx) =
∂x2 2 2 ∂t
2
2 ∂ u 1 ∂u
Therefore, u = e−2π t sin πx is a solution of = .
∂x2 2 ∂t
Example 7.1.2 Verify that
πx πct
u(x, t) = u0 sin cos
L L
satisfies the one dimensional wave equation
∂2u 1 ∂2u
2
= 2 2
∂x c ∂t
where u0 , L and c are constants.
Solution:
By straightforward partial differentiation of the given function u(x, t)
∂2u
π 2
∂u π πx πct πx πct
= u0 cos cos = −u 0 sin cos
∂x L L L ∂x2 L L L
2
∂u πc πx πct ∂ u πc 2 πx πct
= −u0 sin sin = −u 0 sin cos
∂t L L L ∂t2 L L L
We see that
∂2u 1 ∂2u
=
∂x2 c2 ∂t2
which completes the verification.
One possible physical interpretation of this problem is that u(x, t) represents the displacement of a string
stretched between two points at x = 0 and x = L. Clearly the position of any point P on the vibrating string
will depend upon its distance x from one end and on the time t. The boundary conditions represent the fact
πx
that the string is fixed at these end-points as shown in Fig.7.1. The initial condition u(x, 0) = u0 sin
L
represents the displacement of the string at t = 0. From this example, we have found that PDEs are much
Figure 7.1: The displacement of a string stretched between two points at x = 0 and x = L.
more complicated than ODEs. They come in a variety of forms: linear homogeneous, linear inhomogeneous,
semi-linear and non-linear. Their solution depends significantly on the domain of the independent variables,
and the boundary conditions on the dependent variables. We have noted that Only the very simplest PDE
can be solved with analytical methods.
84 CHAPTER 7. PARTIAL DIFFERENTIAL EQUATIONS
∂2u
= 2xet
∂x2
where u is a function of x and t.
Solution:
Integrating with respect to x gives us
∂u
= x2 et + f (t)
∂x
where the arbitrary function f (t) replaces the normal arbitrary constant C of ordinary integration. This
function of t only is needed because we are integrating partially with respect to x, i.e. we are reversing a
partial differentiation with respect to x at constant t. We can integrate it again with respect to x gives the
general solution
x3 t
u(x, t) = e + xf (t) + g(t)
3
where g(t) is a second arbitrary function. We have now obtained the general solution of the given PDE
but to find the arbitrary function we must know two initial conditions. For example, suppose that these
conditions are
∂u
u(0, t) = t, (0, t) = et
∂x
Inserting the first of these conditions into the general solution gives g(t) = t. Inserting the second condition
into the general solution gives f (t) = et . So the final solution is
x3 t
u(x, t) = e + xet + t
3
Example 7.2.2 Solve the PDE
∂2u
= sin x cos y
∂x∂y
subject to the conditions
∂u π
= 2x at y= , u = 2 sin y, at x=π
∂x 2
Solution:
First integrate the PDE with respect to y: (it is equally valid to integrate first with respect to x).
∂u
= sin x sin y + f (x)
∂x
∂u
Remember here to add an appropriate arbitrary function f (x). Since one of the given conditions is on ,
∂x
π
impose this condition to determine the arbitrary function f (x). At y = , we have
2
∂u π
= 2x =⇒ sin x sin + f (x) = 2x =⇒ f (x) = 2x − sin x
∂x 2
So
∂u
= sin x sin y + 2x − sin x
∂x
7.2. SOLUTION OF PARTIAL DIFFERENTIAL EQUATIONS 85
Here don’t forget to add an appropriate arbitrary function g(y). Similarly, we can obtain the arbitrary
function g(y) by using
where X(x) is a function of x only, and T (t) is a function of t only. That is why the method is called as the
separation method. However, we should note that not all solutions to PDEs are of this type; for example,
it is easy to verify that u(x, y) = x2 − y 2 (which is not of the form u(x, y) = X(x)Y (y)), is a solution of
the Laplace equation. In this section we shall only consider this kind of product form to the PDE solutions.
Now, we shall see how to derive the general solution of the heat equation using the separation of variables
method.
∂2u 1 ∂u
2
= , 0 < x < 3, t>0
∂x 2 ∂t
with the boundary conditions
u(0, t) = u(3, t) = 0
and the initial condition
u(x, 0) = 5 sin 4πx
Solution:
The basic idea is to try to find a solution that is a function of x times a function of t. That is, we write
dT d2 X
= 2KT, = KX
dt dx2
The T equation has general solution
T (t) = Ae2Kt
which will increase exponentially with increasing t if K is positive and decrease with t if K is negative. In
any physical problem the latter is the meaningful situation. To emphasise that K is being taken as negative
we assume that
K = −λ2
so 2
t
T = Ae−2λ
Similarly, the X equation becomes
d2 X
= −λ2 X
dx2
86 CHAPTER 7. PARTIAL DIFFERENTIAL EQUATIONS
We cannot deduce that the constant E has to be zero because then the solution (7.2.1) would be the trivial
solution u(x, t) = 0. The only sensible deduction is that
sin(3λ) = 0 =⇒ 3λ = nπ
where n is some integer. Hence solutions of the form (7.2.1) satisfying the 2 boundary conditions have the
form nπx 2n2 π2 t
u(x, t) = En sin e− 9
3
where we have written En for E to allow for the possibility of a different value for the constant for each
different value of n.
We obtain the value of n by using the initial condition u(x, 0) = 5 sin(4πx) and forcing this solution to agree
with it. That is, nπx
u(x, 0) = En sin = 5 sin(4πx)
3
so we must choose n = 12 with E12 = 5. Finally, we have the solution
12πx − 2 (12)2 π2 t 2
u(x, t) = 5 sin e 9 = 5 sin(4πx)e−32π t
3
∂2u 1 ∂2u
= for 0 < x < 2, t>0
∂x2 16 ∂t2
with the boundary conditions
u(0, t) = u(2, t) = 0
and the initial conditions
∂u
u(x, 0) = 6 sin πx − 3 sin 4πx, (x, 0) = 0
∂t
Solution:
We first assume that the solution has a form
X ′′ T ′′
= K, =K
X 16T
Now decide on the appropriate sign for K and then write down the solution to these equations. Choosing
K as negative (say K = −λ2 ) will produce Sinusoidal solutions for X and T which are appropriate in the
context of the wave equation where oscillatory solutions can be expected. Then we have
Similarly,
T ′′ = −16λ2 T =⇒ T = C cos 4λt + D sin 4λt
Now obtain the general solution u(x, t) by multiplying X(x) by T (t) and insert the two boundary conditions
to obtain information about two of the constants.
u(x, t) = X(x)T (t) = (A cos λx + B sin λx)(C cos 4λt + D sin 4λt)
and
u(2, t) = 0 =⇒ B sin(2λ)(C cos 4λt + D sin 4λt) = 0
so, for a non-trivial solution
nπ
sin(2λ) = 0 =⇒ λ= for some integer n
2
At this stage we write the solution as
nπx
u(x, t) = sin (E cos(2nπt) + F sin(2nπt)
2
where we have multiplied constants and put E = BC and F + BD. Now we use the initial condition
∂u
(x, 0) = 0
∂t
to deduce the value of F . First, we differentiate the solution with respect to t
∂u nπx
= sin (−2nπE sin(2nπt) + 2nπF cos(2nπt))
∂t 2
so at t = 0 we have
∂u nπx
(x, 0) = sin 2nπF = 0 =⇒ F = 0
∂t 2
Finally using the other the initial condition u(x, 0) = 6 sin(πx) − 3 sin(4πx) deduce the form of u(x, t). At
this stage the solution reads nπx
u(x, t) = E sin cos(2nπt) (7.2.2)
2
We now insert the last initial condition
nπx
u(x, 0) = 6 sin πx − 3 sin 4πx =⇒ u(x, 0) = E sin (7.2.3)
2
At this point we seem to have incompatability because no single value of n will enable us to satisfy (7.2.3).
However, in the solution (7.2.2), any positive integer value of n is acceptable and we can in fact, superpose
solutions of the form (7.2.2) and still have a valid solution to the PDE Hence we first write, instead of (7.2.2)
∞
X nπx
u(x, t) = En sin cos(2nπt) (7.2.4)
n=1
2
for which
∞
X nπx
u(x, 0) = En sin (7.2.5)
n=1
2
Actually, this is a Fourier series.
To make the solution (7.2.5) fit the initial condition (7.2.3) we do not require all the terms in the infinite
Fourier series. We need only the terms with n = 2 with coefficient E2 = 6 and the term for which n = 8
with E8 = −3. All the other coefficients En have to be chosen as zero. Using these results in (7.2.4) we
obtain the solution
u(x, t) = 6 sin(πx) cos(4πt) − 3 sin(4πx) cos(16πt)
The above solution perhaps seems rather involved but there is a definite sequence of logical steps which can
be readily applied to other similar problems.
88 CHAPTER 7. PARTIAL DIFFERENTIAL EQUATIONS
Hence
dT 2
= Bae−λ t cos ax
dx
dT
Using another boundary condition (L, t) = 0, we can have one of the following conclusions:
dx
B = 0, λ = 0, cos(aL) = 0
The first two possibilities (B = 0 and λ = 0) can be discounted as they leave T = 0 for all x and t and it is
not possible to satisfy the initial condition (7.2.7). Hence, we have to take
1
cos aL = 0, =⇒ aL = n + π
2
7.2. SOLUTION OF PARTIAL DIFFERENTIAL EQUATIONS 89
This is √
λ 1 α 1
aL = √ L = n+ π =⇒ λ = n+ π
α 2 L 2
So the temperature T satisfies
α 1 2 1 πx
T = Be− L2 (n+ 2 ) t sin n+
2 L
πx πx
T (x, 7200) = 800e−0.438 sin = 516 sin
2L 2L
This means the inner wall of the furnace has cooled from 800o C to 516o C.
X ′′ Y ′′
=K = −K
X Y
To determine the sign of K and hence the appropriate solutions for X(x) and Y (y) we must impose appropri-
ate boundary conditions. We will investigate solving Laplace equation in the square {0 ≤ x ≤ l, 0 ≤ y ≤ l}
for the boundary conditions as shown in Fig.7.2
where U0 is a constant.
1. We must first deduce the sign of the separation constant K.
If K is chosen to be positive say K = −λ2 , then the X equation gives a general solution
Figure 7.2: The domain and boundary conditions for the Laplace equation.
If the sign of K is negative K = −λ2 the solutions will change to trigonometric in x and exponential
in y. These are the only two possibilities when we solve Laplace equation using separation of variables
and we must look at the boundary conditions of the problem to decide which is appropriate. Here the
boundary conditions are periodic in x (since u(0, y) = u(l, y)) and non-periodic in y which suggests we
need a solution that is periodic in x and non-periodic in y.
Hence we choose K = −λ2 to give
The appropriate general solution of Laplace equation for the given problem is
u(0, y) = 0 =⇒ A=0
nπ
u(l, y) = 0 =⇒ sin(λl) = 0 =⇒ λ=
l
where n is a positive integer n = 1, 2, 3, · · · . While n = 0 also satisfies the equation it leads to the
trivial u = 0 only.
u(x, 0) = 0 =⇒ C + D = 0 =⇒ D = −C
At this point the solution can be written as
nπx nπy nπy
u(x, y) = BC sin e l − e− l
l
This can be conveniently written as
nπx nπy
u(x, y) = E sin sinh E = 2BC (7.2.9)
l l
At this stage we have just one final boundary condition to insert to obtain information about the
constant E and the integer n. Our solution (7.2.9) gives
nπx
u(x, l) = E sin sinh(nπ)
l
and clearly this is not compatible, as it stands, with the given boundary condition
u(x, l) = U0 = constant
The way to proceed is again to superpose solutions of the form (7.2.9) for all positive integer values of
n to give
X∞ nπx nπy
u(x, y) = En sin sinh (7.2.10)
n=1
l l
from which the final boundary condition gives
∞
X nπx ∞
X nπx
U0 = En sin sinh(nπ) = bn sin 0<x<l (7.2.11)
n=1
l n=1
l
f (x) = U0 , 0<x<l
7.2. SOLUTION OF PARTIAL DIFFERENTIAL EQUATIONS 91
Recalling the work on half-range Fourier series we must extend this definition to produce an odd
function with period 2l. Hence we define
U0 0<x<l
f (x) =
−U0 −l < x < 0
f (x) = f (x + l)
which has the period 2l as shown in Fig.7.3.
3. We can now apply standard Fourier series theory to evaluate the Fourier coefficients bn in (7.2.11). We
obtain
4U0 l nπx
Z
bn = En sinh(nπ) = sin dx
2l 0 l
Carrying out the integration, we obtain
4U0
2U0 n = 1, 3, 5, · · ·
En sinh(nπ) = (1 − cos nπ) =⇒ En = nπ sinh(nπ)
nπ 0 n = 2, 4, 6, · · ·
Since f (x) is a square wave with half-period symmetry we are not surprised that only odd harmonics
arise in the Fourier series. Finally substituting these results for En into (7.2.10) we obtain the solution
to the given problem as the infinite series
∞
4U0 X 1 nπx nπy
u(x, y) = sin sinh
π n sinh nπ l l
n=1
(n odd)
Example 7.2.6 Determine the steady-state temperature distribution governing by the equation
∂2T ∂2T
2
+ =0
∂x ∂y 2
in a rectangular plate that is insulated along the edges x = 0 and x = a, so that
Tx (0, y) = Tx (a, y) = 0
and whose edges y = 0 and y = b are
πx
T (x, 0) = 0, T (x, b) = T0 cos
a
Solution:
Letting T (x, y) = X(x)Y (y) yields
X ′′ (x) = kX(x), X ′ (0) = X ′ (a) = 0
and
Y ′′ (y) = −kY (y), Y (0) = 0
If k = 0 yields a non-trivial solution X0 (x) = constant. The boundary conditions cannot be satisfied if k > 0,
n2 π 2
but, if k < 0, we find that kn = −λ2n = − 2 and that
a
nπx
Xn (x) = cos
a
92 CHAPTER 7. PARTIAL DIFFERENTIAL EQUATIONS
By matching coefficients on the two sides of this equation, we see that only the n = 1 term survives, yielding
T0
β1 =
sinh πb
a
so that
T0 πy πx
T (x, y) = sinh cos
sinh πb
a
a a
We can verify that T (x, y) satisfies Laplace equation and also fits the boundary conditions. Fig.7.4 shows
T (x, y) plotted over the rectangular region.
1.0
0.5
1.0
0.0
-0.5
-1.0
0.0 0.5
0.5
1.0
1.5
0.0
2.0
Figure 7.4: The steady-state temperature distribution for the system by the Laplace equation, where a =
2, b = 1, T0 = 1.
7.3. EXERCISES 93
7.3 Exercises
1. Show that
u(x, t) = sin x sin(vt)
is a solution to the one-dimensional wave equation
∂2u 1 ∂2u
=
∂x2 v 2 ∂t2
2. Show that
1
φ(x, y, z) = p , (x, y, z) 6= (0, 0, 0)
x2 + y2 + z2
satisfies Laplace equation
∂2φ ∂2φ ∂2φ
∇2 φ = + 2 + 2 =0
∂x2 ∂y ∂z
3. Show that 2
T (x, y, z, t) = Ae−3α t sin x sin y sin z
satisfies the three-dimensional heat equation
∂2T ∂2T ∂2T 1 ∂T
2
+ 2
+ = 2
∂x ∂y ∂z 2 α ∂t
4. Use the direct integration method to solve the equation
∂2T
= et sinx
∂x2
where T (x, t) is a function of x and t.
5. The temperature distribution in the slab governed by the heat equation is
∂2T 1 ∂T
= 2 , 0 ≤ x ≤ 1, 0<t
∂x2 α ∂t
subject to the boundary conditions
T (0, t) = T (1, t) = 0
and the initial condition
T (x, 0) = T0 = constant, 0<x<1
Find the temperature distribution T (x, t). (Hint: see Example 7.2.3).
6. Show that
2πx 2πvt
u(x, t) = U0 sin cos
L L
satisfies the wave equation
∂2u 1 ∂2u
= , 0 ≤ x ≤ L, t>0
∂x2 v 2 ∂t2
and the boundary conditions
u(0, t) = u(L, t) = 0
and the initial conditions
2πx ∂u
u(x, 0) = U0 sin , (x, 0) = 0
L ∂t
7. Verify (by substitution) that each u satisfies
∂2u ∂2u
+ 2 = f (x, y)
∂x2 ∂y
with f (x, y) as indicated
y 2y
(a) u= , f=
x x3
(b) u = sin xy, f = −(x2 + y 2 ) sin xy
2
+y 2 2 2
(c) u = ex f = 4(x2 + y 2 )ex +y
1 1
(d) u= p f=p
x2 + y2 (x + y 2 )3
2
94 CHAPTER 7. PARTIAL DIFFERENTIAL EQUATIONS
8. Verify (by substitution) that the given function is a solution of the heat equation
∂u ∂2u
= c2 2
∂t ∂x
with suitable c
2 2
c t
(a) u = e−t sin x (b) u = e−ω cos ωx
2
(c) u = e−9t sin ωx (d) u = e−π t cos 25x
9. Verify (by substitution) that the given function is a solution of the wave equation
∂2u ∂2u
2
= c2 2
∂t ∂x
with suitable c
Solution:
4. u(x, t) = sin x cos t + xf (t) + g(t)
P sin(2n−1)πx −α2 (2n−1)2 π 2 t
5. T (x, t) = 4Tπ0 ∞ n=1 2n−1
e
3
/3
10.(b) u = a(y) cos 4πx + b(y) sin 4πx, (c) u = c(x)e−y (d) u = e−3y (a(x) cos 2y + b(x) sin 2y) + 0.1e3y .
Chapter 8
Appendixes
A Formulas
1. Trigonometry formulas
sin A
sin2 A + cos2 A = 1 tan A =
cos A
sin 2A = 2 sin A cos A cos 2A = cos2 A − sin2 A
1 + cos 2A 1 − cos 2A
cos2 A = sin2 A =
2 2
sin(A + B) = sin A cos B + cos A sin B sin(A − B) = sin A cos B − cos A sin B
cos(A + B) = cos A cos B − sin A sin B cos(A − B) = cos A cos B + sin A sin B
sin(−A) = − sin A cos(−A) = cos A
1 1
sin A sin B = cos(A − B) − cos(A + B)
2 2
1 1
cos A cos B = cos(A − B) + cos(A + B)
2 2
1 1
sin A cos B = sin(A − B) + sin(A + B)
2 2
2. Laws of logarithm
If A = bc , then logb A = c.
These laws apply in any bases:
A
log A + log B = log AB log A − log B = log
B
log X ln X
n log A = log An loga X = =
log a ln a
3. Laws of powers
Am
Am An = Am+n Am ÷ An = n = Am−n
√ A √
(Am )n = Amn 1/n
Am/n = Am
n n
A = A
95
96 CHAPTER 8. APPENDIXES
The determinant is
a b
|A| = = ad − bc
c d
6. Vectors
(a) Normal
If r = x i + y j + z k then p
|r| = x2 + y 2 + z 2
(b) Scalar (dot) product
If a = a1 i + a2 j + a3 k and b = b1 i + b2 j + b3 k then
a · b = a 1 b1 + a 2 b2 + a 3 b3
a · b = |a||b| cos θ
(c) Vector (cross) product
If a = a1 i + a2 j + a3 k and b = b1 i + b2 j + b3 k then
i j k
a × b = a1 a2 a3 = (a2 b3 − a3 b2 ) i + (a3 b1 − a1 b3 ) j + (a1 b2 − a2 b1 ) k
b1 b2 b3
7. Complex numbers
(a) Cartesian form √
z = a + bj, j= −1
(b) Polar form
z = a + bj, z = r(cos θ + j sin θ)
p b
a = r cos θ, b = r sin θ, r = a2 + b2 tan θ = , (−π ≤ θ ≤ π)
a
(c) Exponential form z = rejθ
(d) Euler’s relation
ejθ = cos θ + j sin θ, e−jθ = cos θ − j sin θ
(e) Multiplication and division in polar form
z1 r1
z1 z2 = r1 r2 ∠(θ1 + θ2 ) = ∠(θ1 − θ2 )
z2 r2
z n = rn ∠(nθ)
(f) De Moivre’s theorem
(cos θ + j sin θ)n = cos nθ + j sin nθ
8. The binomial theorem
n(n − 1) n−2 2 n(n − 1)(n − 2) n−3 3
(a + b)n = an + nan−1 b + a b + a b + · · · + bn
2! 3!
9. Elementary derivatives
df
function f (x) derivatives = f ′ (x)
dx
k 0 (k is a constant )
n
x nxn−1
ekx kekx
1
ln(kx) (k is any constant)
x
sin kx k cos kx
cos kx − k sin kx
tan kx k sec2 kx
1
sin−1 x √
1 − x2
−1
cos−1 x √
1 − x2
A. FORMULAS 97
b u(b)
du du
Z Z Z Z
f (u) dx = f (u)du, f (u) dx = f (u)du
dx a dx u(a)
1 1
Z Z
ln xdx = x ln x − x + C, dx = ln(ax + b)
ax + b a
1 1 −1 x 1 1 x − a
Z Z
dx = tan + C, dx = ln +C
x 2 + a2 a a x 2 − a2 2a x + a
1 1 x
Z p Z
√ dx = ln |x + x2 ± a2 | + C, √ dx = sin−1 + C
2
x ±a 2 2
a −x 2 a
1
Z p p p
x2 ± a2 dx = x x2 ± a2 ± a2 ln |x + x2 ± a2 | + C
2
kx
ne n
Z Z
n kx
x e dx = x − xn−1 ekx dx
k k
Z Z
tan xdx = − ln cos x + C tan2 xdx = −x + tan x + C
x 1 x 1
Z Z
2
sin xdx = − sin 2x + C cos2 xdx = + sin 2x + C
2 4 2 4
Z Z
x cos xdx = cos x + x sin x + C x sin xdx = −x cos x + sin x + C
1 1
Z Z
ex sin xdx = ex (sin x − cos x) + C ex cos xdx = ex (sin x + cos x) + C
2 2
1 x
Z
x
xe sin xdx = e (cos x − x cos x + x sin x) + C
2
1
Z
xex cos xdx = ex (x cos x − sin x + x sin x + C
2
∂φ ∂φ ∂φ
gradφ = ∇φ = i + j + k
∂x ∂y ∂z
where
∂ ∂ ∂
∇= i + j + k
∂x ∂y ∂z
(b) The divergence of a vector field:
div v = ∇ · v ⇒ a scalar
(d) Higher order derivatives can be also formed, giving the following: