Engineering Mathematics With Applications
Engineering Mathematics With Applications
with Applications to
Fire Engineering
http://taylorandfrancis.com
Engineering Mathematics
with Applications to
Fire Engineering
Khalid Khan
Tony Lee Graham
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
This book contains information obtained from authentic and highly regarded sources. Reasonable efforts have been made to
publish reliable data and information, but the author and publisher cannot assume responsibility for the validity of all materi-
als or the consequences of their use. The authors and publishers have attempted to trace the copyright holders of all material
reproduced in this publication and apologize to copyright holders if permission to publish in this form has not been obtained.
If any copyright material has not been acknowledged please write and let us know so we may rectify in any future reprint.
Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted, or utilized in
any form by any electronic, mechanical, or other means, now known or hereafter invented, including photocopying, micro-
filming, and recording, or in any information storage or retrieval system, without written permission from the publishers.
For permission to photocopy or use material electronically from this work, please access www.copyright.com (http://www
.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923, 978-750-
8400. CCC is a not-for-profit organization that provides licenses and registration for a variety of users. For organizations that
have been granted a photocopy license by the CCC, a separate system of payment has been arranged.
Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used only for identifi-
cation and explanation without intent to infringe.
Names: Khan, Khalid Mahmood, 1963- author. | Graham, Tony Lee, author.
Title: Engineering mathematics with applications to fire engineering / Khalid Khan
and Tony Lee Graham.
Description: Boca Raton : Taylor & Francis, 2018. | Includes bibliographical references
and index.
Identifiers: LCCN 2018001261| ISBN 9781138098848 (hardback : acid-free paper) | ISBN
9781315104270 (ebook)
Subjects: LCSH: Engineering mathematics. | Engineering mathematics--Problems,
exercises, etc. | Fire protection engineering--Mathematics.
Classification: LCC TA332.5 .K43 2018 | DDC 620.001/51--dc23
LC record available at https://lccn.loc.gov/2018001261
Preface xiii
Authors xv
v
vi Contents
Answers 357
Index 363
http://taylorandfrancis.com
Preface
xiii
xiv Preface
We consider the presentation in this text as unique in that there are real-world fire
engineering applications given alongside the mathematical content that underpins
those applications.
The overall structure of the book is that it begins with a review of the key
elementary mathematical concepts and focuses on important concepts such as
transposition of formulae, as these form an essential part of many engineering
solutions. An introduction to probability theory with discrete and continuous ran-
dom variables are important concepts used in fault tree analysis in risk assessment
and reliability theory, respectively. Determinants and matrices lead to solving a
system of linear equations using different techniques such as Gaussian elimina-
tion and the matrix inversion method. Vectors and normal vectors to surfaces are
considered and form the basis of concepts in surface integrals. Use of complex
numbers and their applications in electrical circuit theory and the role they play
in the solutions to differential equations are covered. There is an introduction to
one-variable calculus with the fundamentals of differentiation, integration tech-
niques, and importance of integration as a summing process for a multitude of
engineering applications. Methods of solving ordinary linear differential equa-
tions are introduced with emphasis on the Laplace transform method as a valuable
tool in finding solutions to problems. Higher-dimensional multivariable calculus
dealing with partial derivatives, double and triple integrals, and general change
of coordinate systems are covered. Finally, in the last chapter on vector calculus,
vector fields representing physical phenomenon are considered along with con-
cepts of divergence and curl, and applications of Green’s, Stokes’, and divergence
theorems are given.
We would like to thank the staff at CRC/Taylor & Francis who have contrib-
uted to the production of this book and to Dr. Alan Burns for his valuable com-
ments on the draft version of the book. Finally, we wish to express our gratitude
to the University of Central Lancashire for providing a conductive academic envi-
ronment that allowed us to complete this project.
Authors
Khalid Khan, BSc (Hons), MSc, PhD, received his BSc (Hons) in mathemat-
ical physics, and MSc and PhD in control systems all from the University of
Manchester Institute of Science and Technology in the United Kingdom. Dr Khan
then spent two years as a consultant engineer working on safety, reliability and
risk assessment problems in the energy industry. Subsequently, he moved abroad
and after spending some period of time as an assistant professor of mathemat-
ics at Etisalat University in the United Arab Emirates he returned to the United
Kingdom and took up a position at the University of Central Lancashire.
Dr Khan is currently a senior lecturer in engineering mathematics in the
School of Engineering at the University of Central Lancashire. He teaches on a
range of mathematics modules within the fire degree programs and contributes
to other mathematics teaching within the school and college. He is currently the
course leader for the Foundation Degree in Fire Safety Engineering. Dr Khan as
part of the fire team has been involved in the development of a range of courses
in fire safety engineering and management that are currently running at one of
University of Central Lancashire’s international partnerships in Qatar. Dr Khan is
also a senior fellow of the Higher Education Academy (HEA), which is a British
professional institution promoting excellence in higher education.
Dr Khan’s research interests are in the area of mathematical modeling of sys-
tem behavior in a range of applications. His current work focuses on fire suppres-
sion using sprinkler systems and on mathematical models of collective motion of
self-propelled particles in homogeneous and heterogeneous mediums. Dr Khan
currently has over thirty publications and is also a member of two journal review
panel boards.
Tony Graham, BSc (Hons), PhD, is a senior lecturer and course leader at the
University of Central Lancashire, United Kingdom. He is best known for teaching
fire safety engineering to thousands of students over twenty-four years in differ-
ent countries and for papers on compartment fire dynamics and the phenomenon
of flashover fire. His teaching also includes risk engineering and engineering
analysis. He has taught courses at International College of Engineering and
Management in Sultanate of Oman and also at City University of Hong Kong.
xv
xvi Authors
1.1 Degrees of Accuracy
1.1.1 Rounding Numbers (Common Method)
In the real world, when dealing with numbers a degree of accuracy is needed. For
example, if a piece of wood of a certain length was required, asking for a length of
123.732461 centimeters would not be sensible as such accurate measurements are
not possible. What is more usual is some form of rounding. The method of round-
ing is commonly used in mathematical applications in science and engineering.
It is the one generally taught in mathematics classes in high school. The method
is also known as round-half-up. It works as follows:
• Leave it the same if the next digit is 4 or less (this is called rounding
down).
Example 1.1
3.044 rounded to hundredths is 3.04 (because the next digit, 4, is less
than 5).
3.045 rounded to hundredths is 3.05 (because the next digit, 5, is 5 or
more).
3.0447 rounded to hundredths is 3.04 (because the next digit, 4, is less
than 5).
For negative numbers, one rounds the absolute value and reapplies the sign
afterward.
1
2 Review of Basic Concepts
Example 1.2
−2.1349 rounded to hundredths is −2.13.
−2.1350 rounded to hundredths is −2.14.
1.1.2 Round-to-Even Method
The round-to-even method method, also known as unbiased rounding or Gaussian
rounding, exactly replicates the common method of rounding except when the
digit(s) following the rounding digit starts with a 5 and has no nonzero digits
after it. The new algorithm becomes
• Otherwise, if all that follows the last digit is a 5 and possibly trailing
zeros, then increase the rounded digit if it is currently odd; else, if it is
already even, leave it alone.
All rounding schemes have two possible outcomes: increasing the rounding
digit by one or leaving it alone. With traditional rounding, if the number has a
value less than the halfway mark between the possible outcomes, it is rounded
down; if the number has a value exactly halfway or greater than halfway between
the possible outcomes, it is rounded up. The round-to-even method is the same
except that numbers exactly halfway between the possible outcomes are some-
times rounded up, sometimes down.
Despite the custom of rounding the number 4.5 up to 5, 4.5 is no nearer to 5
than it is to 4 (it is 0.5 away from both). When dealing with large sets of scientific
or statistical data, where trends are important, traditional rounding on average
biases the data upward slightly. Over a large set of data, or when many subsequent
rounding operations are performed as in digital signal processing, the round-to-
even rule tends to reduce the total rounding error, with (on average) an equal por-
tion of numbers rounding up as rounding down. This generally reduces upward
skewing of the result.
Examples 1.3
3.016 rounded to hundredths is 3.02 (because the next digit, 6, is 6 or
more).
3.013 rounded to hundredths is 3.01 (because the next digit, 3, is 4 or
less).
3.015 rounded to hundredths is 3.02 (because the next digit is 5, and the
hundredths digit, 1, is odd).
3.045 rounded to hundredths is 3.04 (because the next digit is 5, and the
hundredths digit, 4, is even).
3.04501 rounded to hundredths is 3.05 (because the next digit is 5, but it
is followed by nonzero digits).
1.1 Degrees of Accuracy 3
Example 1.4
Table 1.1 shows how the round-to-even system compares with the old sys-
tem of rounding. There are five original data points that are rounded to the
tenths and then their average is calculated. It can be seen that the round-to-
even method is much more accurate than the old method of rounding.
1.1.3 Decimal Places
When rounding a number, one is usually told how to round it. It is simplest when
one is told how many places to round to, but one should also know how to round
to a named place, such as to the nearest thousand or to the ten-thousandths place.
Also, it may be required to know how to round to a certain number of significant
digits; this is dealt with later.
Using the first few digits of the decimal expansion of pi = π = 3.14159265... in
the following examples.
Example 1.5
Round pi to five decimal places. First, count out the five decimal places,
then look at the sixth place:
3.14159 265
A little line separating the fifth place from the sixth place has been drawn.
This can be a good way of keeping your place, especially if dealing with
lots of digits.
The fifth place has a 9 in it. Looking at the sixth place, it has a 2 in it.
Since 2 is less than 5, the 9 will not be rounded up; that is, just leave the
9 as it is. In addition, delete the digits after the 9. Then pi, rounded to five
decimal places, is given as 3.14159.
Example 1.6
Round pi to four decimal places. First, go back to the original number:
3.14159265. Count off four places, and look at the number in the fifth place:
3.1415 9265
That is, the 5 becomes a 6, the 9265… part disappears, and pi, rounded to
four decimal places, is given as 3.1416.
This rounding works the same way when rounding to a certain named
place, such as the hundredths place. The only difference being a bit more
careful in counting off the places needed. Just remember that the decimal
places count off to the right in the same order as the counting numbers
count off to the left. That is, for regular numbers, the place values are
For decimal places, a “oneths” is not there, but all the other fractions are
Example 1.7
Round pi to the nearest thousandth. The “nearest thousandth” means that
one needs to count off three decimal places (tenths, hundredths, thou-
sandths), and then round:
3.141 59265
Example 1.8
Round 18.796 to the hundredths place. The hundredths place is two decimal
places, so count off two decimal places, and round according to the third
decimal place:
18.79 6
Since the third decimal place contains a 6, which is greater than 5, one has
to round up. But rounding up a 9 gives a 10. In this case, round the 79 up to
an 80 as 18.80.
One might be tempted to write this as 18.8, but, since rounded to the
hundredths place (to two decimal places), one should write both decimal
places. Otherwise, it looks like rounding to one decimal place, or to the
tenths place, and the answer could be counted off as being incorrect.
1.1.4 Significant Places
Rounding can also be carried out to an appropriate number of significant digits.
What are significant digits? Well, they are sort of the “interesting” or “important”
digits. For example:
3.14159 has six significant digits (all the numbers give you useful information).
1000 has one significant digit (only the 1 is interesting; you do not know
anything for sure about the hundreds, tens, or units places; the zeroes
1.1 Degrees of Accuracy 5
may just be placeholders; they may have rounded something off to get
this value).
1000.0 has five significant digits (the “.0” tells us something interesting
about the presumed accuracy of the measurement being made: that the
measurement is accurate to the tenths place, but that there happens to be
zero tenths).
0.00035 has two significant digits (only the 3 and 5 tell us something; the
other zeroes are placeholders, only providing information about relative
size).
0.000350 has three significant digits (that last zero tells us that the measure-
ment was made accurate to that last digit, which just happened to have a
value of zero).
1006 has four significant digits (the 1 and 6 are interesting, and the zeros
have to be counted, because they are between the two interesting
numbers).
560 has two significant digits (the last zero is just a placeholder).
560. (notice the point after the zero) has three significant digits (the decimal
point tells us that the measurement was made to the nearest unit, so the
zero is not just a placeholder).
560.0 has four significant digits (the zero in the tenths place means that
the measurement was made accurate to the tenths place, and that there
just happen to be zero tenths; the 5 and 6 give useful information, and
the other zero is between significant digits, and must therefore also be
counted).
3. All zeros that are both to the right of the decimal point and to the right of
all nonzero significant digits are themselves significant.
Following are some rounding examples; each number is rounded to four, three,
and two significant digits.
Example 1.9
Round 742,396 to four, three, and two significant digits:
Example 1.10
Round 0.07284 to four, three, and two significant digits:
Example 1.11
Round 231.45 to four, three, and two significant digits:
Scientists have developed a shorter method to express very large and very small
numbers. This method is called scientific notation. Scientific notation is based on
powers of the base number 10.
Example 1.12
Our galaxy to which the sun belongs is called the Milky Way. It con-
tains at least 100,000,000,000 stars. Now let’s look at this number:
100,000,000,000. It can be written as 1.0 × 100,000,000,000. It is the large
number 100,000,000,000 that causes the problem. But that is just a multiple
of 10. In fact, it is 10 times itself 11 times:
A × 10 N (1.1)
101 = 10
10 2 = 100
10 3 = 1, 000
and so on. Similarly, 100 = 1, since the zero exponent means that no zeros follow
the 1.
Negative exponents indicate negative powers of 10, which are expressed as
fractions with 1 in the numerator (on top) and the power of 10 in the denominator
(on the bottom). So
10 −1 = 1/10
10 −2 = 1/100
10 −3 = 1/1, 000
and so on. This allows one to express other small numbers this way. For example,
Example 1.13
(2.0 × 10 2 ) + (3.0 × 10 3 )
can be rewritten as
(0.2 × 10 3 ) + (3.0 × 10 3 )
Just add 0.2 + 3 and keep the 103 intact. Your answer is 3.2 × 103, or 3,200.
This can be checked by converting the numbers first to the more familiar
form. So,
Example 1.14
(2.0 × 10 7 ) − (6.3 × 10 5 )
The problem needs to be rewritten so that the exponents are the same.
So this can be written as
1.2.3 Multiplication
When multiplying numbers expressed in scientific notation, the exponents can
simply be added together. This is because the exponent represents the number of
zeros following the one. So,
Example 1.15
Multiply the following numbers: (4.0 × 105) × (3.0 × 10 –1). The 4 and the
3 are multiplied, giving 12, but the exponents 5 and –1 are added, so the
answer is 12 × 104, or 1.2 × 105.
Checking: (4 × 105) × (3 × 10 –1) = 400,000 × 0.3 = 120,000 = 1.2 × 105.
1.3 Basic Algebra 9
1.2.4 Division
Example 1.16
(6.0 × 108 )
(3.0 × 10 5 )
To solve this problem, first divide the 6 by the 3, to get 2. The exponent
in the denominator is then moved to the numerator, reversing its sign (this
will be explained further when dealing with indices). So, move the 105 to
the numerator with a negative exponent, which then looks like this: 2 ×
108 × 10 –5. All that is left now is to solve this as a multiplication problem,
remembering that all that needs to be done for the 108 × 10 –5 part is to add
the exponents. So, the answer is 2.0 × 103 or 2,000.
1.3 Basic Algebra
1.3.1 Algebraic Notation
Algebraic notation describes how algebra is written. It follows certain rules and
conventions, and has its own terminology.
For example, the expression in Figure 1.1 has the following components to it:
1: exponent (power); 2: coefficient; 3: terms; 4: operators; 5: constant, and with x,
y: variables.
A coefficient is a numerical value or letter representing a numerical constant
that multiplies a variable (the operator is omitted).
A group of coefficients, variables, constants and exponents that may be sepa-
rated from the other terms by the plus and minus operators. Letters represent
variables and constants.
By convention, letters at the beginning of the alphabet (e.g., a, b, and c) are
typically used to represent constants, and those toward the end of the alphabet
(e.g., w, x, and y) are used to represent variables.
Algebraic operations work in the same way as arithmetic operations, such
as addition, subtraction, multiplication, division, and exponentiation, and are
applied to algebraic variables and terms.
2 1 2
7 x11 + 3xy – 8
3 4 3 4 5
1.3.2 Evaluating Expressions
Algebraic expressions may be evaluated and simplified, based on the basic prop-
erties of arithmetic operations (addition, subtraction, multiplication, division, and
exponentiation).
Added terms are simplified using coefficients. For example, x + x + x + x can
be simplified as 4x (where 4 is a numerical coefficient).
Multiplied terms are simplified using exponents. For example, x × x × x × x is
represented as x4.
Like terms are added together, for example, 5x + 4p + 1 − 2x + 6p + 4 is w ritten
as 3x + 10p + 5.
Brackets can be multiplied out, using the distributive property. For example,
4(x + 3) can be written as 4 × x + 4 × 3, which can be written as 4x + 12.
This idea can be extended to multiply out two brackets as (x + 4) (x + 3). Here
in the first bracket the x term multiplies the (x + 3) and then +4 term multiplies
the (x + 3) as follows:
( x + 4)( x + 3) = x ( x + 3) + 4( x + 3) = x 2 + 3x + 4 x + 12 = x 2 + 7 x + 12.
1.4 Linear Equations
1.4.1 Solving Linear Equations
In engineering, the physical modeling of system behavior is done using the language
of mathematics. Different types of equations are derived that model the systems and
the simplest type of equations are called linear. An equation of the form ax + b = 0,
where a and b are constants, is said to be a linear equation with variable x.
Note: For an equation to be linear, the power the variable, in this case x, has to
be raised to is one.
The method of solving linear equations is to collect all the terms involving x
on one side of the equation and everything else on the other side. The idea is to
isolate the variable x to be on its own. The way this is achieved is through using
certain operations like addition, subtraction, multiplication, division, and others
(i.e., squaring and square rooting) to manipulate the equation so as to keep both
sides of the equation the same. This is illustrated in the following examples.
Example 1.17
Solve the following equation to find x: x + 3 = 7.
Here it is easy to see what the answer should be for x, x = 4. But how can
this be arrived at systematically since for more complicated equations
1.4 Linear Equations 11
the answer will not be obvious. The approach is to consider the equa-
tion as
So, to keep this balanced, any operation carried out on the LHS must
also be carried out on the RHS. Starting with the equation given earlier:
Example 1.18
Solve the following equation to find x: x – 5 = 4.
Starting with the equation:
Example 1.19
Solve the following equation to find x: 3x = 12.
Starting with the equation:
Example 1.20
x
Solve the following equation to find x: = 6
Starting with the equation, 5
x
=6 (multiply by 5 on both sides of the equationn)
5
x
×5=6×5 (tidying up both sides)
5
x = 30 (solved for x )
Example 1.21
Solve the following equation to find x: 5x − 3 = 2x + 15
Starting with the equation:
Example 1.22
Solve the following equation to find x: 5(x + 3) + 4 (2x − 3) = 2(2x + 15)
Starting with the equation:
Note: Sometimes the linear equations are disguised because they involve
fractions. A good strategy is to multiply every term by the lowest common
multiple (LCM) of the denominators, as seen in the next example.
Example 1.23
x−5 4−x
Solve the following equation to find x: − =5
4 3
1.4 Linear Equations 13
Note: These basic ideas of solving linear equations are very important in
transposing equations.
1.4.2 Transposing Formulae
In science and engineering, formulae are used to relate physical quantities to each
other. It is found in electrical circuit theory that the power P is related to the cur-
rent I and resistance R by the following equation:
P = I 2R
W IV
=
I I
or
W W
=V ⇒ V=
I I
V
R×I= ×I
I
or
RI = V ⇒ V = RI
V − E = IR
Divide both sides by R:
V−E V−E
=I ⇒ I=
R R
R = P + 5Pt
R − P = 5Pt
1.4 Linear Equations 15
R − P 5Pt R−P
= ⇒ t=
5P 5P 5P
J
I2 =
Rt
I
I=
Rt
Example 1.31
x2 = z − y
x2 + y = z ⇒ z = x2 + y
( x − 1) y = x + 2
xy − y = x + 2
xy − x = y + 2
x ( y − 1) = y + 2
16 Review of Basic Concepts
Divide by (y − 1):
y+2
x=
y −1
Example 1.33
Alpert’s equation for the ceiling jet velocity when the jet is far away from
the fire is given as
13 1 2
Q H
U = 0.195
5
r 6
where velocity, U, is in meters per second (m/s); total energy release rate,
is in kilowatts (kW); and the ceiling height and radial position (r and
Q,
H) are in meters (m).
Transpose this equation to make Q the subject. First, rewrite to make Q
on the LHS:
13 1 2
Q H
0.195 =U
5
r 6
13 1 2
Q H U
=
5 0 . 195
r 6
5
Multiply both sides by r 6 :
1 1 U 56
Q 3 H = r
0.195
2
1
Divide both sides by H 2 :
5
1 U r 6
Q =
0.195 1 2
3
H
1.5 Linear Simultaneous Equations 17
3
5
3 5
U r 6 U r 2
Q = =
0.195 H 21
0.195 H 2 3
Example 1.34
9
If F = C + 32, then find F if C = 20.
5
9
F= (20) + 32 ⇒ F = 9 × 4 + 32
5
F = 68
Now, if one needed to find the value of C given F, then it would be easier
to rearrange the equation first to make C the subject and then substitute in
the given value of F (see next example).
Example 1.35
9
Given that F = C + 32, find the value of C if F = 86.
5
Solution: Rearranging
9
C = F − 32
5
5
C = (F − 32)
9
So putting in F = 86 yields
5
C= (86 − 32) ⇒ C = 30
9
not as accurate as the others and is not considered here. The substitution method
can sometimes involve awkward fractions and so the elimination method is gener-
ally the most preferred method.
1.5.1 Elimination Method
The method of elimination makes the coefficients of one of the variables equal,
and then the two equations are either added or subtracted in order to eliminate
that variable.
Note: In the equation 3x − 5y = 7, the coefficients of x and y are 3 and –5, respectively.
Example 1.36
Solve the simultaneous equations
x+y=6 (1)
x−y=4 (2)
Solution: This is the simplest case, and since the coefficients of both the x
and y are the same it is easy. Just add Equations 1 and 2 to eliminate the
y variable since the y coefficients are of opposite signs. If, however, the x
variable was to be eliminated first, then Equations 1 and 2 would need to be
subtracted since the coefficients are of the same signs.
Adding Equation 1 to 2 gives
2 x = 10 ⇒ x = 5
x + y = 6 ⇒ 5+ y = 6 ⇒ y =1
Note: Remember to check the answers by seeing if they fit both the original
Equations 1 and 2.
Example 1.37
Solve the following simultaneous equations:
x + 2y = 5 (1)
3x − y = 1 (2)
1.5 Linear Simultaneous Equations 19
Now both the coefficients of x and y are different, so the task is first to
make one of the variable coefficients the same. Here there is no preferred
choice and so try to eliminate the y variable.
To make the coefficients of y the same, multiply Equation 2 by 2, which
gives
x + 2y = 5 (3)
6x − 2 y = 2 (4)
7x = 7 ⇒ x =1
x + 2y = 5 ⇒ 1 + 2y = 5 ⇒ 2y = 4 ⇒ y = 2
Example 1.38
Solve the simultaneous equations
2 x + 3 y = 13 (1)
7 x − 5 y = −1 (2)
10 x + 15 y = 65 (3)
21x − 15 y = −3 (4)
31x = 62 ⇒ x = 2
2 x + 3 y = 13 ⇒ 4 + 3 y = 13 ⇒ 3 y = 9 ⇒ y = 3
1.5.2 Substitution Method
The second method of solving simultaneous linear equations involves rearrang-
ing one of the equations and substituting into the other. This technique is called
the method of substitution. It requires making one of the variables the subject and
then using this value into the second equation.
Example 1.39
Solve the following simultaneous equations using the method of substitution:
7 x + 2 y = 11 (1)
4x + y = 7 (2)
y = 7 − 4x
7 x + 2(7 − 4 x ) = 11
7 x + 14 − 8 x = 11
− x = −3
x=3
1.6 Quadratic Equations
We have seen that equations like 2x + 1 = 0 and x + 2 = 0 are examples of linear
equations. Now, suppose one has an equation like this: (x + 2)(2x − 1) = 0. This is
not a linear equation; it is called a quadratic equation because if the two sets of
brackets are expanded, it gives x2 + 3x − 2 = 0.
Note: Here the highest power of the variable x is now 2, and the word quadratic
is derived from the Latin word quadratus meaning “square.”
A quadratic equation has not one solution like the linear equations already seen,
but two solutions. What are they? Start by looking at the shape of this equation:
The only way that this can happen is if one of the numbers in the brackets is
zero. In other words, either x + 2 = 0 or 2x − 1 = 0. Now, it has already been shown
how to solve simple linear equations like these.
1.6 Quadratic Equations 21
1
If x + 2 = 0, then x = −2. And if 2x − 1 = 0, then 2 x = 1 ⇒ x = .
2
So the answers to the equation (x + 2)(2x − 1) = 0 are x = −2 or x = 0.5.
Example 1.40
One can write the answers to the following quadratic equations fairly easily.
1
( x − 4)(3 x + 1) = 0 Solution: x = 4 or x = −
3
1 2
(10 x + 1)(5 x − 2) = 0 Solution: x = − or x =
10 5
1
x (2 x − 1) = 0 Solution: x = 0 or x =
2
3
( x − 7)(8 x − 3) = 0 Solution: x = 7 or x =
8
These equations are quadratic equations that have already been factor-
ized, that is, written as
that is, (x − a)(x − b) = 0, where a and b are called the “roots” of the equation.
Leaving the topic of factorizing quadratics aside, let’s turn to the situation
when the quadratic equation doesn’t factorize. For example, x2 + 8x + 7 = 0
factorizes and can quickly be solved as shown earlier. But changing the
final number to say 8, that is, x2 + 8x + 8 = 0, means that this cannot be
done by the factorizing method. Also in realistic engineering problems it is
highly unlikely that the quadratic equation generated will factorize. What
then? In all general cases, a quadratic formula can be used to solve the
problem. It is better to consider this method as it can be used in all cases
encountered.
ax 2 + bx + c = 0 (1.2)
where a, b, and c are numbers, then the solutions to this equation are given by the
formula known as the general solution to a quadratic equation as
− b ± b 2 − 4 ac
x= (1.3)
2a
This term underneath the square root sign is given a special name called the
discriminant and the symbol Δ, where, Δ = b2 − 4ac.
22 Review of Basic Concepts
Notes:
• If the number under the square root sign is positive, there are two real
and distinct solutions.
• If the number under the square root sign is zero, there is one real and
repeated solution.
• If the number under the square root sign is negative, there are no real
solutions. This topic is dealt with in Chapter 5, where it will be shown that
in this third case the answers can in fact be written as complex solutions.
Example 1.41
Solve the equation x2 + 8x + 8 = 0 correct to four decimal places.
Solution: First compare the given quadratic equation with the general qua-
dratic ax2 + bx + c = 0. This then determines the values of the coefficients
a, b, and c to be used in the formula as a = 1, b = 8, and c = 8. Substituting
these values into the general formula given by Equation 1.3 gives
− b ± b 2 − 4 ac
x=
2a
−8 ± 82 − 4(1)(8) −8 ± 32 −8 ± 4 2
x= = = = −4 ± 2 2
2(1) 2 2
Example 1.42
Solve the equation 4x2 − 3x − 11 = 0 correct to three decimal places.
Solution: Again, first compare the given quadratic equation with the general
quadratic ax2 + bx + c = 0. Determining the values of the coefficients a, b,
and c as follows: a = 4, b = –3, and c = –11. Substituting these values into
the general formula given by Equation 1.3 gives
− b ± b 2 − 4 ac
x=
2a
Note: All quadratic equations can are solved in the same way with just dif-
ferent coefficients for the values of a, b, and c.
1.7 Trigonometry 23
1.7 Trigonometry
Many problems in engineering especially mechanics involve forces and these can
be represented as either right angled triangles or general scalene triangles. The
need to solve these problems requires solutions to triangular problems. First, the
methods of solution to right angled triangles is considered.
1.7.1 Right-Angled Triangles
A right-angled triangle is one in which one of the angles is 90 degrees. Consider
the general right-angled triangle shown in Figure 1.2. First, the labeling of the
sides is very important. The longest side of the right-angled triangle is always
called h the hypotenuse, the side facing opposite the angle is o the opposite, and
the side next to the angle is called a the adjacent side, as shown in Figure 1.2.
Hypotenuse
h
Opposite
o
Adjacent
a
When dealing with just the sides of the triangle, Pythagorean theorem can be
used to relate the different sides as follows:
h 2 = a 2 + o2 (1.4)
Hence, given any two sides of a right-angled triangle, the third side can be calcu-
lated using Equation 1.4.
Example 1.43
Find the missing length AB as shown in the triangle given in Figure 1.3.
h
6 cm
A C
8 cm
h 2 = a 2 + o2 = 82 + 62 = 64 + 36 = 100
When dealing with angles and sides, for the right-angled triangle there are
three trigonometric ratios sin θ, cos θ, and tan θ and these are defined as follows:
o
sin θ = (1.5)
h
a
cos θ = (1.6)
h
o
tan θ = (1.7)
a
These three formulae can more easily be remembered by using the acronym:
o
Note: Given that sin θ = then to find the angle θ, the inverse sine (i.e., sin –1)
h o
function would have been used on both sides to give θ = sin− 1 . Similarly,
h
inverse function expressions exist for the cosine and tangent functions.
When solving any triangle problem first make sure all sides are labeled cor-
rectly, and then identify which one of the ratios is required to solve the problem.
Example 1.44
Find the missing sides o and a in the triangle shown in Figure 1.4.
h = 10 cm
o
30°
A C
a
Solution: To find the side o, use a trigonometry identity involving the sides
o
o and side h. Using, SOH CAH TOA, that is, sin θ = needs to be used and
substituting in values for θ and h gives h
o
sin 30° = ⇒ o = 10 sin 30° ⇒ o = 5 cm
10
Example 1.45
Find the angle θ in the triangle given in Figure 1.5.
o= 3
θ
A C
a=1
3 3
∴ tan θ = ⇒ θ = tan −1 ,
1 1
1.7.2.1 Sine Rule
Given a general triangle as shown in Figure 1.6, the labeling is again very important
as the formulae depend on it. Here, the angles are denoted by the capital letters and
the side opposite that angle is denoted by the corresponding lowercase letter.
The sine rule relates the sine of an angle with its opposite side as follows:
a b c
= = (1.9)
sin A sin B sin C
26 Review of Basic Concepts
c
a
A
b C
Note: In Equation 1.9 any two of the relationships are used as required,
that is,
a b a c b c
= or = or =
sin A sin B sin A sin C sin B sin C
Example 1.46
In the triangle given in Figure 1.7, find the length of side AC.
Solution: First find the missing angle B, angle B is given by 180° − (83° +
62°) = 35°
a b c
So now using the sine rule, = = , since the side AC (i.e.,
sin A sin B sin C
side b) is required and side a and angle A are also known, then using
a b 10 b 10 × sin 35
= ⇒ = ⇒ b= = 5.78 cm
sin A sin B sin 83 sin 35 sin 83
Note: It is not always possible to make use of the sine rule directly, so
sometimes there is a need for the second rule.
c 83° b
B 62°
a = 10 C
c
a
A
b C
1.7.2.2 Cosine Rule
The cosine rule is needed when there are two given sides and only the angle
between them is given. Consider the general triangle in Figure 1.8.
The Cosine rule is stated as
And by symmetry,
To find the angles when all three sides are known, Equation 1.10 can be rear-
ranged to give
b2 + c2 − a2
cos A = (1.13)
2bc
Example 1.47
Find the length AC given in Figure 1.9.
Solution: Using the cosine rule given by Equation 1.11, b2 = a2 + c2 − 2ac cos B
gives
c=8 b
62°
B
a = 10 C
1.7.3 Resultant Forces
1.7.3.1 Adding Two Forces
Consider two forces of magnitude, F1 and F2 , that act upon a particle. If these
forces are placed end to end, it can be seen that they have the same effect as
a single force of magnitude F, as in Figure 1.10. This force is known as the
resultant force. The resultant force will form the third side in a triangle of
forces. To calculate this resultant force and its direction, use of the cosine and
sine rule are made.
F2 Resultant
R
F2
F1
F1
Example 1.48
Two forces of magnitude, 6 N and 5 N, act on a particle. The angle
between the forces is 40°. Find the magnitude and direction of the resul-
tant force.
To find the direction of the resultant force with respect to the 6 N force
means finding the angle θ in Figure 1.11. This can be done using the sine
rule as
5 F 5 × sin 140
= ⇒ sin θ = ⇒ θ = 18.1°
sin θ sin 140 10.34
So, the resultant force has magnitude 10.34 N and is at an angle 18.1° to the
6 N force.
Resultant
5N R
40° 140°
θ 5N
6N 6N
Figure 1.11 Diagram showing resultant force forming a general triangle of forces.
1.7 Trigonometry 29
o a o
sin θ = , cos θ = , tan θ =
h h a
There are important trigonometric identities that can be derived and are very
useful in applications later on. Two basic identities are
sin θ
tan θ = (1.15)
cos θ
Note: These can be proved from the aforementioned trigonometric ratios and
can be used to prove other trigonometric identities as well as solve trigonometric
equations.
Example 1.49
3 4
Find the value of tan θ when sin θ = and cos θ = − .
5 5
Solution:
sin θ 3 4 3 5 3
tan θ = = ÷ − = × − =−
cos θ 5 5 5 4 4
Example 1.50
1 − sin θ 1
Show that ≡ − tan θ .
cos θ cos θ
Note: When asked to “show that” or “prove that” then consider starting
with one side of the identity and, step by step, reduce it to the same form as
the other side of the identity.
Solution:
1 − sin θ 1 sin θ 1
LHS: ≡ − ≡ − tan θ
cos θ cos θ cos θ cos θ
Example 1.51
2 − cos2 θ
Show that ≡ 1.
1 + sin 2 θ
30 Review of Basic Concepts
There are other useful compound formulae for the sine, cosine, and tangent as
follows:
cos( A ± B) = cos A cos B sin A sin B (1.16)
tan A ± tan B
tan( A ± B) = (1.18)
1 tan A tan B
1
cos( A) cos( B) = ( cos( A + B) + cos( A − B)) (1.19)
2
1
sin( A)sin( B) = ( cos( A − B) − cos( A + B)) (1.20)
2
1
sin( A) cos( B) = (sin( A + B) + sin( A − B)) (1.21)
2
The preceding formulae are very important whenever the need arises to trans-
form the product of sines and cosines into sums and are a very useful in the
techniques of integration and applications in areas such as Fourier series, seen
later in Chapter 9.
1.7.5 Radian Measure
When it comes to measuring angles generally the measurement that has been used
is degrees. A whole circle is 360°, a straight line is 180°, and a right angle is 90°,
for example. However, when it comes to finding gradients of curves, or rates of
change (i.e., to differentiate trigonometric functions), a different unit of measure-
ment is used called the radian.
Radian is short for the radius angle, and it means the angle given at the center
of a circle by an arc of one radius as shown in Figure 1.12.
Now the relationship between radians and degrees can be determined as fol-
lows: There are 360° in a complete revolution of the circle, the circumference of a
circle is 2πr, and so the number of radiuses r around the circumference is given by
2π r
= 2π radians. Therefore, this gives an important formula relating degrees
r
to radians:
360° ≡ 2π radians (1.22)
1.8 Statistics 31
1 radian r
The standard relationships that exist between degrees and radians is shown in
Table 1.2.
When talking about an angle in degrees, one should write the degrees sign
as 45.2°. When talking about an angle in radians, a sign is not used. Thus
π
cos(60°) = cos = 0.5 and sin (25°) = 0.4226, but sin (25) = −0.1324 because
3
this means sin (25 radians).
1.8 Statistics
1.8.1 Introduction
Statistics deals with all aspects of data: collecting the data, pictorial representa-
tion of the data, and numerical analysis and drawing final conclusions on findings.
There are vast areas where statistics are used ranging from opinion pollsters (test-
ing public opinions on issues), governmental national statistics, and in science and
engineering testing theories.
32 Review of Basic Concepts
Qualatative Quantative
(no numerical value,
(having numerical value)
colour of eyes etc.)
There are two types of variables, as shown on Figure 1.13, and any quantity
that varies is called a variable.
There are some different aspects to statistical data that need defining and fur-
ther explanation:
Population—This is all the possible data for the research question. If the
question was to determine the average height of male adults in the United
Kingdom, then the population would need to include measuring every
single adult male.
Sampling without bias—A bias is anything that makes the sample unrepresenta-
tive (e.g., asking only certain members of a community their views a topic).
1.8.2 Measures of Averages
There are three main measures of average of a set of numerical data:
Median—Arrange the data in order of magnitude (size), then find the cen-
tral value. If there is an even number of data values, then use halfway
between the two middle values.
Mean—Add all the data values and divide by the total number of data points.
Note: For most purposes the mean is considered the most useful measure of aver-
age since it uses all data.
The symbol μ (pronounced myü) is used for the mean of the whole population
and the symbol x (x bar) is used for the mean of the sample. The definition of the
mean can be more concisely written in formula form as
x=
∑x i
(1.23)
n
Note: Here the symbol ∑ x means to sum or add together all the x values.
i i
Equation 1.23 is used for a simple set of data (i.e., a small number of data
points) and where n is the number of data points.
Example 1.52
The number of fires reported in 20 consecutive weeks is given by the data
set as:
4 7 12 13 0 5 21 13 10 6
6 8 15 9 6 0 14 12 6 8
Find the modal, median, and mean number of fires per week.
0 0 4 5 6 6 6 6 7 8 8 9 10 12 12 13 13 14 15 21
The modal number is, 6 as this occurs more times than all others.
For the median, since there is an even number of data points (i.e., 20 data
points) look at the 10th and 11th data values. In this case they are both 8 and
8, which gives the middle of these as
8+8
Median = =8
2
34 Review of Basic Concepts
x=
∑x i
=
175
= 8.75.
n 20
x=
∑x f i i
(1.24)
∑f i
where fi is just the frequency associated with the ith data point.
Example 1.53
A survey was carried out to see the distribution of the ages (in years) of fire
engines in a particular region of the country. Twenty-seven vehicles were
surveyed and the results are shown in Table 1.3. Find the mode, median, and
mean age of the fire engines.
Solution: The first thing to do is redraw the table with an extra column for
the product term xi fi, as shown in Table 1.4. The columns can be filled in
and the products computed. At the bottom of the columns, the sums of the
columns can also be calculated.
Mean: x =
∑x f =
i i 45
= 1.67 years.
∑f i
27
∑ f = 27
i ∑ x f = 45
i i
1.8.2.2 Grouped Data
Sometimes the individual data points are not provided but only the grouped fre-
quency table is given. In this case it is not possible to give the exact values of the
three averages but only estimates.
For the mean, the method is the same as for Example 1.48, but first the mid-
point of the grouped data set must be determined. This is a good approximation
since some of the data values will be higher than this and some will be lower than
this middle value. The next example shows how this method works.
Example 1.54
Consider the marks for 61 students who took a test as given in Table 1.5.
Find the modal class, median, and the mean percentage mark for this class.
Solution:
6
51 + × (class width = 10) = 53.73%
22
∑ f = 61i ∑x f = 3085.5
m i
The mean is obtained as before, but now the midpoint of the class width is
needed since the exact data values are not known, as shown in Table 1.6.
So the estimated mean is
∑ x f = 50.58%
m i
∑f i
1.8.3 Measures of Spread
In the previous section it was seen that for a data set an average value could be
calculated to indicate something about the data points. However, this does not
say anything about how the data points are spread out. There are three common
measures of spread: The range (R), the interquartile range (IQR), and the standard
deviation (σ) are defined next.
1.8.3.1 Range
The range for a set of data is the difference between the highest value and the
lowest value (i.e., it considers the extreme values of the data set).
Example 1.55
See the following two data sets A and B:
A: 4, 5, 5, 6, 7, 9
B: 1, 3, 3, 5, 6, 8, 10, 12
1.8.3.2 Interquartile Range
When a set of n data is written in order of magnitude (size), the median is given
(n + 1)
by th item of data.
2
1.8 Statistics 37
• The lower quartile is the median of the lower half of the data.
• The upper quartile is the median of the upper half of the data.
Example 1.56
Find the interquartile range (IQR) for the following set of data:
24, 24, 25, 26, 26, 26, 27, 27, 30, 33, 33, 35, 35, 36, 43
Solution: There are 15 data values, so the median is the eighth data value.
Thus, median = 27.
Lower quartile has 7 data values ⇒ median of this is the fourth value = 26.
Upper quartile has 7 data values ⇒ median of this is 35.
∴ IQR = UQ − LQ = 35 − 26 = 9
σ=
∑ (x − µ)i
2
(1.25)
n
σ=
∑ (x − µ) f
i
2
i
(1.26)
∑f i
s=
∑ (x − x ) i
2
(1.27)
n −1
s=
∑ (x − x ) f
i
2
i
(1.28)
∑ f −1 i
Note: Generally, it is much easier to calculate the mean and standard deviation
using a calculator (in statistics mode).
Example 1.57
Look at the following set of data and find the mean and standard deviation:
xi : 1, 2, 3, 4, 5
Solution:
Mean: µ =
∑x i
=
15
=3
n 5
Now use Table 1.7 to find the standard deviation σ as
∴σ=
∑ (x − µ)
i
2
=
10
= 2 = 1.414
n 5
∑ (x − µ)i
2
= 10
1.8 Statistics 39
Example 1.58
The number of fires over 52 weeks was recorded at a particular fire station.
The results are shown as follows:
No. of fires, xi 4 5 6 7 8 9 10
Frequency, fi 3 10 18 6 5 6 4
Calculate the mean and standard deviation of the number of fires per
week.
µ=
∑x f i i
=
346
= 6.65
∑f i
52
σ=
∑ (x − µ) f
i
2
i
=
143.77
= 2.765 = 1.66
∑f i
52
Finally, the square of the standard deviation (σ2) is given a special name
as the variance.
Table 1.8 Calculating Mean and Standard Deviation for Data
in a Frequency Table
xi fi xi fi (xi − μ) (xi − μ)2 fi
4 3 12 –2.65 21.0675
5 10 50 –1.65 27.225
6 18 108 –0.65 7.605
7 6 42 0.35 0.735
8 5 40 1.35 9.1125
9 6 54 2.35 33.135
10 4 40 3.35 44.89
∑ f = 52 ∑ x f = 346
i i i ∑ (x − µ)
i
2
fi = 143.77
1.8.4 Change of Scale
The weekly wages of employees in a small company have a mean of £290 and stan-
dard deviation of £42. If there is a pay rise of £15 for each employee, what happens
to the mean and standard deviation? Since each wage is increased by £15, the mean
wage will be increased by £15 to £305. However, the standard deviation measures
variability and this is unchanged since all wages have increased by the same amount.
If instead of a flat rate increase an increase of 10% is given to each employee, the
variability would increase because the higher paid employees would get a larger rise.
In this case both the mean and standard deviation would increase by 10%. In sum-
mary, if a variable is increased by a constant amount, its average will be increased
by this amount, but the spread will be unchanged. If the variable is multiplied by a
constant amount, both its average and spread will be multiplied by this amount.
40 Review of Basic Concepts
1.9 Applications
Example 1.59: Relationship between Sizes of Pool Fires to Flame Heights
A schematic diagram of a pool fire with its flame height is shown in
Figure 1.14. From pool fire experiments, a power law fit was determined
relating the heat release rate Q (kW) and the diameter of the pool fire D (m)
to the flame height L (m) using Heskestad’s equation as follows:
2
L = 0.235 Q 5 − 1.02 D (1.29)
Knowing the heat release rate for a fire and its diameter size, then using
Equation 1.29 can predict the height of the flame.
Consider a 500 kW diesel pool fire with a diameter of 1.5 m, the flame
height is given by substituting in the data values into Equation 1.29 as
2
L = 0.235(500) 5 − 1.02(1.5) = 1.29 m
D = 1.5 m
Figure 1.14 Diesel pool fire with its associated flame height.
Fuel array
smoke layer measured radially far away from the fire is given by Alpert’s
equation as
2
(Q /r ) 3
Tjet − T0 = 5.38 (1.30)
H
where Tjet is the smoke layer temperature, T0 is the ambient room tempera-
ture, Q is the heat release rate, r is the radial distance of the smoke, and H
is the room height.
If a sprinkler was located 5m radially from the fire with a ceiling height
of 3m and the fire size is 1200 kW with an ambient temperature of 20°C
then what is the expected gas temperature at the sprinkler?
Equation 1.30 can be transposed to make Tjet the subject of the formula
to give
2
(Q /r ) 3
Tjet = T0 + 5.38 (1.31)
H
2
(240) 3
Tjet = 20 + 5.38 = 89°C
3
Vw
Vs
the resultant velocity VR of the smoke flow and the direction it flows with
reference to the horizontal floor.
First constructing a triangle of velocities gives Figure 1.17.
Vw = 0.6
Vs = 1.2
VR
α
But for the standard deviation, only carry out the multiplication since the
addition of 32 will have no effect on the variability giving
Problems
1.1 Transpose the following formulae for the variable given:
3.6V
a. L = for A
A
b. F = G + 7H for H
k
c. α = for k
ρc
d. v = h − gt for t
Problems 43
d2
e. t = for d
4α
f. I = εσ AT 4 for T
g H2
g. τ = t for t
H S
1 2
h. E = mv for v
2
h+3
i. g = for h
h−5
1
Q 3
j. U = 0.96 for Q
H
1.2 The equation for the smoke jet Tj is given by Equation 1.31 as
Tj = T0 + 5.38
(Q /r ) 3
a. a + b = 2 and 5a + b = 14
c. x2 + 5x − 3 = 0
d. 3x2 − 6x − 11 = 0
30 m
(Burning building)
A B
V ms–1
20 ms–1 200 m
70° θ°
O
Fire boat station
a. Find the time it takes the fireboat to cross the river and the actual
direction of the boat θ.
2.1 Introduction
Most people deal with uncertain events each day in their lives but rarely pause to
calculate chances or probabilities. Yet whole professions, such as the insurance
industry, pensions, investment advisors, and bookmakers are founded upon prob-
ability. In engineering, the ideas of probability are very important when consider-
ing the overall reliability of systems and in the areas of risk assessment.
How is the probability of something happening defined? The usual definition
supposes that one repeats an “experiment” many times and records the outcomes.
For example, roll a fair dice many times and record how often the result is a six.
The probability of a six occurring is given by
1
If the die was fair and rolled lots of times, the expected probability would be 6 .
Note: If A denotes the event A happening, then A′ or (Ac) is called the complement
of A and denotes the event A not happening.
45
46 Introduction to Probability Theory
P( A′) = 1 − P( A) (2.2)
For example, for a fair dice P(1) = P(2) = P(3) = P(4) = P(5) = P(6). And of
course 1 + 1 + 1 + 1 + 1 + 1 = 1. But it can also be concluded, for example, since
6 6 6 6 6 6
P(6) = 1
, then P(not a 6) = 1 − 1 = 5 .
6 6 6
It is not always possible to assign a probability using equally likely outcomes.
If someone wanted to know the probability of going to the bus stop and having
to wait for a bus for more than 5 minutes, then some trials would have to be done
to find the waiting times. The relative frequency of an event is the proportion of
times it has been observed to happen. If somebody went for a bus on 40 weekday
mornings and on 16 of these they had to wait more than 5 minutes, one could
16
assign a probability of = 0.4 to the event of having to wait more than 5 minutes.
40
Example 2.1
When rolling a dice, the
1 1 1
P(either getting a 3 or 4) = + =
6 6 3
Example 2.2
When choosing one card from a pack, the probability of getting a ten card
or a jack card is
4 4 8 2
P(a ten or a jack ) = P( ten) + P( jack ) = + = =
52 52 52 13
P( A or B) = P( A) + P( B) (2.3)
( )
P A1 or A2 or A3 ……or An = P( A1 ) + P( A2 ) + P( A3 ) +……. + P( An )
2.1 Introduction 47
n n
P Ai =
i =1
∑ P( A )
i =1
i
If the events are not mutually exclusive these formulae are modified. For
example Equation 2.3 now becomes
P ( A or B) = P( A) + P( B) − P( A and B) (2.4)
This is because these events can happen together and the probability of
B is included in the probability of A and vice versa, so this means that the
probability has been included twice and so one of these must be subtracted
from the answer.
2.1.2 Independent Events
Independent events are those that have no influence on each other. When tossing
a coin twice, the result of the first throw has no effect on the second.
In such a case, the probability of both events happening is found by multiply-
ing the separate probabilities.
Example 2.3
1 1 1
P (head on both tosses of a coin) = P (head) × P (head) = × =
2 2 4
1 1 1
P (rolling two sixes with a dice) = P (6) × P (6) = × =
6 6 36
P( A and B) = P( A) × P( B) (2.5)
( )
P A1 and A2 and A3 ……and An = P( A1 ) P( A2 ) P( A3 )…….P( An )
n n
P Ai =
i =1
∏ P( A )
i =1
i
If the events are not independent, then one has conditional probabilities
and these formulae are modified (see later).
48 Introduction to Probability Theory
Example 2.4
The probability that telephone calls to a railway timetable inquiry service
are answered is 0.7. If three calls are made find the probability that all three
are answered and exactly two are answered.
Solution: If A is the event of a call being answered, P(A) = 0.7, then A′ is the
probability of a call not being answered, and P(A′) = 1 – 0.7 = 0.3.
Using the multiplication rule the probability of AAA = 0.7 × 0.7 ×
0.7 = 0.343.
If one call is unanswered it could be the first, second, or the third call:
These three outcomes are mutually exclusive and so one can apply the
addition law and find the probability of exactly two calls being answered to
be 0.147 + 0.147 + 0.147 = 0.441.
2.1.3 Conditional Probability
Sometimes the events A and B are not independent, and so the probability of the
event A happening depends on the event B happening. This is represented as follows:
P(A|B) denotes the probability that event A happens given that event B
happens.
P( A and B) = P( A) × P ( B A ) (2.6)
Example 2.5
James buys ten apparently identical oranges. Unknown to him the flesh of
two of these oranges is rotten. He selects two of the ten oranges at random
and gives them to his friend. Find the probability that
2
a. The probability of both oranges being rotten is = 0.022.
90
16 16
b. The probability of exactly one orange being rotten is + = 0.35.
90 90
2.1 Introduction 49
2 1 2
1 × =
Rotten Rotten, rotten 10 9 90
9
2 2 8 16
Rotten Rotten, o.k. × =
10 10 9 90
8
O.k.
9
2
Rotten 8 2 16
9 O.k., rotten × =
8 10 9 90
O.k.
10
7 8 7 56
O.k. O.k., o.k. × =
9 10 9 90
2.1.4 Bayes’ Theorem
Founded by the Rev. Thomas Bayes (1701–1761), who apart from being a minister
was a statistician and philosopher, Bayes’ theorem is a fact about probabilities
and has a lot of real-world applications. It gives a way of working out what the
conditional probabilities should be.
A notation that is used for the following case is that the probability of a hypoth-
esis, H, given that a new piece of evidence E is written as P(H\E).
Bayes’ theorem states
P( E \H ) P( H )
P( H \E ) = (2.7)
P(E )
P( H and E ) = P( H ) P( E \H )
P( E and H ) = P( E ) P( H \E )
P( E ) P( H \E ) = P( E \H ) P( H )
P( E \H ) P( H )
P( H \E ) =
P(E )
Example 2.6
One day a person who does not feel well decides to go onto the Internet to
find out what might be wrong. Let’s say the person found an illness called
hypothesitis, H. So the probability of the symptoms given hypothesitis is
P(E\H) = 0.95.
But Bayes’ theorem can be used to find the probability that the person
will have hypothesitis given the symptoms as
P( E \H ) P( H )
P( H \E ) =
P(E )
P( E \H ) P( H ) 0.95 × 0.0001
P( H \E ) = = = 0.00095
P(E ) 0.01
P ( B) = ∑ P(B \A ) P( A )
j
j j
P( B \Ai ) P( Ai )
P( Ai \B) = (2.8)
∑ P(B \A ) P( A )
j
j j
Example 2.7
The output from a factory is produced on three machines: A1, A 2 , and
A 3. The three machines account for 20%, 30%, and 50% of the out-
put, respectively. The fraction of defective items produced is 5% for
A1, 3% for A 2 , and 1% for A 3. If an item is chosen at random and the
output is found to be defective, what is the probability it was produced
by machine A 3?
Solution: Let Ai denote the event that a randomly chosen item was made by
the ith machine (i = 1, 2, 3). Let B denote the event that a randomly chosen
item is defective. So we know P(A1) = 0.2, P(A2) = 0.3, and P(A3) = 0.5.
Also, P(B\A1) = 0.05, P(B\A2) = 0.03, and P(B\A3) = 0.01.
2.1 Introduction 51
P( B \A3 ) P( A3 )
P( A3 \B) =
∑
3
P( B \A j ) P( A j )
j =1
0.01 × 0.5 5
P( A3 \B) = = = 0.208
0.05 × 0.2 + 0.03 × 0.3 + 0.01 × 0.5 24
2.1.5 Tree Diagrams
An alternative approach to solving probability problems involving a series of
events is with a tree diagram. Consider the case of telephone calls made to an
engineering supply company. The probability that a call is answered is given as
0.7. A tree diagram showing the possible outcomes is given in Figure 2.2. Each
branch shows the possible outcomes of each call and their probabilities. Here, A =
call answered and A′ = call not answered. The outcome of the three calls is found
by reading along the branches leading to it, and the probability of this outcome is
found by multiplying the individual probabilities along these branches.
Note: This is an important concept in assessing the risk of certain events occur-
ring. See examples in applications section.
From the tree diagram the probabilities of different outcomes can be calcu-
lated. The probability of all three calls being answered AAA can be seen to be
0.343. The probability of exactly two calls being answered is the sum of the
probabilities of the three outcomes: AAA′, AA′A, and A′AA = 0.147 + 0.147 +
0.147 = 0.441.
Note: The capital letter stands for the variable itself and the small letter
stands for the value the variable takes.
This notation can be used in Table 2.1 below to give the possible values
for the number of squares moved and the probability of each value. This
table is called the probability distribution of X.
Example 2.9
A bag contains two red and three blue marbles. Two marbles are selected at
random without replacement and the number, X, of blue marbles is counted.
Find the probability distribution of X.
The tree diagram illustrating this situation is shown in Figure 2.3, where
R1 denotes the event that the first marble is red and R2 the event that the
second marble is red. Similarly, B1 and B2 stand for the events that the first
and second marbles, respectively, are blue. X can take the values 0, 1, and 2.
P( X = 0) = P( R1 and R2 ) = P( R1 ) × P R2 R1 = ( ) 2 1
× =
2 1
=
5 4 10 10
P( X = 1) = P( B1 and R2 ) + P( R1 and B2 )
( )
= P( B1 ) × P R2 B1 + P( R1 ) × P B2 B1 ( )
3 2 2 3 12 3
= × + × = =
5 4 5 4 20 5
P( X = 2) = P( B1 and B2 ) = P( B1 ) × P B2 B1 = ( ) 3 2
× =
6
=
5 4 20 10
3
Note: For any random variable, X, the sum of the probabilities is 1 and
this is given as
∑ P( X = x ) = 1 (2.9)
1
P (R2|R1)= R1 and R2
4
2
P (R1)=−
5 R1 and B2
3
P (B2|R1) =
4
2
P (R2|B1)= B1 and R2
4
3
P(B1) =
5
2
P (B2|B1)= B1 and B2
4
2.2.2 Expectation Values
The expected value or the mean of a probability distribution is denoted by μ. The
new symbol is used in order to distinguish the mean of a probability distribution
from x, the mean of a data set. μ is often called the expectation or expected value
of X and is denoted by E(X).
The expectation of a random variable X is defined by
E(X ) = µ = ∑x p i i (2.10)
Example 2.10
Find the expected value of the variable X, which has the probability distri-
bution shown in Table 2.3.
Solution:
E(X ) = ∑ x p = 1 × 16 + 2 × 16 + 3 × 16
i i
1 1 1
+ 4 × + 5 × + 6 × = 3.5
6 6 6
Example 2.11
Find the expected value of the variable Y, which has the probability distri-
bution shown in Table 2.4.
E (Y ) = ∑ y p = 4 121
i i
σ 2 = Var ( X ) = ∑ ( x − µ ) p =∑ x p − µ
i
2
i
2
i i
2
(2.11)
Example 2.12
Calculate the standard deviation of the random variable X given in
Example 2.10.
Solution: First calculate ∑x p :2
i i
1
∑ x p = 1 × 16 + 2 × 16 + 3 × 16 + 4
2
i i
2 2 2 2
×
6
1 1 91 1
+ 52 × + 6 2 × = = 15
6 6 6 6
Var ( X ) = ∑x p − µ
2
i i
2 1
= 15 − 3.52 =
6
35
12
= 2.92
Then calculate the standard deviation, which is just the square root of
the variance:
35
σ= = 1.71 (correct to three significant figurees)
12
Note: In this section, finding the integral of functions is required and so the
topic of integration is important. If one is not familiar with integrating func-
tions, then it is advised to study Chapter 6 first, then return to this section
afterward.
It has already been shown that for a discrete random variable X, it is possible to
allocate probabilities to each discrete value, x, that X can take. When considering
a continuous random variable this is not the case.
For a continuous random variable X, probabilities are allocated to each of the
range of values that the variable can take. This is done by defining a function f(x)
called the probability density function (pdf).
56 Introduction to Probability Theory
∫ f (x) dx = 1
all x
(2.12)
Then, the probability that the random variable X takes a value in the range
a ≤ x ≤ b is given by the integral
x =b
P (a ≤ X ≤ b) =
∫ f (x) dx
x =a
(2.13)
Example 2.13
A continuous random variable X has the pdf defined by
3
(1 − x 2 ) −1 ≤ x ≤ 1
f (x) = 4
0 otherwise
Solution:
0.5
3
P(0.2 ≤ X ≤ 0.5) =
∫ 4 (1 − x ) dx
0.2
2
0.5
3 x 3
3 0.53 0.23
x − = 0.5 − − 0.2 −
4 3 0.2 4 3 3
= 0.196 (correct to three significant digits)
F ( x ) = P( X ≤ x ) =
∫ f (x) dx
−∞
(2.14)
dF ( x )
where = f ( x ).
dx
Note: If f(x) is defined only on the range of values a ≤ x ≤ b, then this becomes
2.3 Continuous Random Variables 57
F ( x ) = P( X ≤ x ) =
∫ f (x) dx
a
(2.15)
Example 2.14
The continuous random variable X has the following cumulative distribu-
tion function:
0 x≤0
3
x
F (x) = 0<x≤4
64
1 x>4
Solution:
33 27
a. P( X ≤ 3) = F (3) = = .
64 64
dF ( x )
b. Using, f ( x ) =
dx
3/4
x
0 4
3x 2
0≤x≤4
f ( x ) = 64
0 otherwise
58 Introduction to Probability Theory
E(X ) =
∫ x f (x) dx
allx
(2.16)
Example 2.15
A random variable T has the pdf given by
0 02 0 < x < 10
f (t ) = . t
0 otherwise
Find E(T).
10
E (T ) =
∫ t f (t) dt
0
10
=
∫ 0.02t dt
0
2
10
0.02t 3 2
= =6
3 0 3
E (aX ) = aE ( X ) (2.17)
E (aX + b) = aE ( X ) + b (2.18)
2
Var ( X ) = E ( X 2 ) − E ( X ) (2.19)
σ = Var ( X ) (2.20)
2.3 Continuous Random Variables 59
Example 2.16
A random variable T has the pdf given by
0 02 0 < x < 10
f (t ) = . t
0 otherwise
2
Var (T ) = E (T 2 ) − E (T )
So
10
E (T ) =
2
∫ t f (t) dt
0
2
10
=
∫ 0.02t dt
0
3
10
0.02t 4
= = 50
4 0
Therefore,
2
2 5
Var (T ) = 50 − 6 = 5
3 9
5
σ= 5 = 2.36
9
Note: In general, the following results can be shown to be true for the vari-
ance of a continuous random variable X:
2.4 Applications
Example 2.17: Event Tree Analysis (ETA) Showing
Probabilities of Outcomes
Failure of a complicated engineering system can lead to different damage
scenarios. The consequence of a particular failure event may depend on a
sequence of events following the failure. The means for systematic identi-
fication of the possible event sequences is the so-called event tree. This is
a visual representation, indicating all events that can lead to different sce-
narios. In the following example, first identify the events.
Consider the initiation event A, fire ignition reported to a fire squad.
After the squad has been alerted and done its duty at the place of accident,
a form is completed where a lot of information about the fire can be found:
type of alarm, type of building, number of staff involved, and much more.
Here the focus is on the condition of the fire at the arrival of the fire brigade.
This is described as
The place where the fire was extinguished is described by the event
65 30 30
E1c E2c = × = = 0.3
100 65 100
2.4 Applications 61
( ) ( ) (
P E1c and E2c = P E1c P E2c E1c = ) 65 30 30
× =
100 65 100
= 0.3
(32)
E2
(35)
E1
E2C (3)
(100)
(65) E2 (35)
E1C
E2C (30)
Figure 2.5 The event tree with the numbers within the parentheses indicating the
number of cases observed after 100 fire ignitions.
Water fails
(0.01206)
AND
OR OR
Figure 2.6 Fault tree diagram showing the probabilities of failure in parentheses.
Here, the diesel pump D is the most unreliable component and so mak-
ing this more reliable (i.e., reducing its failure rate to 0.1) would mean
the top event failure rate now becomes 0.00606 or 1/165 failures, which
is a good improvement. Even further analysis could be carried out by
the company to reduce the risk of failure by introducing a second (i.e.,
a spare) diesel pump and seeing what the effect of this would have on
the overall failure rate. It turns out that introducing “redundancy” is a
more effective method for increasing the overall reliability of a system
(see Problem 2.8).
P(Y ) P( D \Y )
P(Y \D) =
P( X ) P( D \X ) + P(Y ) P( D \Y ) + P( Z ) P( D \Z )
that is, there is a 40% chance that the defective sprinkler was supplied by
company Y.
Problems
2.1 A card is picked at random from a shuffled pack. What is the probability
of getting
a. A diamond
b. An ace or a king
2.3 Two used fire engines were bought by a fire station. The probability
that fire engine A is working in a year’s time is 0.9 and the probability
that the second fire engine B is working in a year’s time is 0.7. Find the
probability that
2.4 A fair coin is tossed three times. Find the probability that the number of
tails is 0, 1, 2, or 3.
2.6 Calculate the expected value and the standard deviation of the variable
X, which has the probability distribution given in Table 2.6.
64 Introduction to Probability Theory
2.7 Given that a system has a failure probability density function f(t), where
time is measured in years as follows,
0 t<0
1
f (t ) t 0≤t ≤4
8
0 t>4
Calculate the mean time to failure (MTTF) and the standard deviation
for the system.
2.8 Consider the system given in Example 2.18. It was found that reducing
the diesel pump failure rate by half to 0.1 reduced the overall failure
rate from 1/83 to 1/165. To further improve this system, the company
decides to install a second pump (i.e., a spare) with the same failure rate
of 0.2 rather than improving the reliability of the first pump. Calculate
the new system failure rate for the top event “water fails” and comment
on your findings.
3 Vectors and
Geometrical
Applications
3.1 Introduction
In engineering, there are different mathematical quantities that are used to
describe physical phenomena. These quantities can be divided into two catego-
ries: scalar or vector. These two quantities are defined as follows:
65
66 Vectors and Geometrical Applications
y
2-D
4 a = 3i + 4j
x
3
3-D
z
7 b = 2i + 3j + 7k
2
3
y
r = ai + bj
θ
x
0
r = xi + yj + zk
r
Example 3.1
In 2-D space:
If r = 3 i + 4 j , then the magnitude of r is given as r = 32 + 4 2 , r = 25 ,
so r = 5.
Example 3.2
In 3-D space:
Given the vector r = i + j − 4 k , find the magnitude r .
3.1.1.1 Unit Vectors
r
Given a vector r , a unit vector in the direction of r is denoted by r̂, where r̂ =.
r
This means that to find a unit vector, take the vector and divide its components by
the magnitude of that vector.
68 Vectors and Geometrical Applications
Example 3.3
For the vector given by r = 2 i − 4 j + 3k to find the unit vector in the direc-
r
tion of r , using r̂ = gives r = (2)2 + (−4)2 + (3)2 , r = 4 + 16 + 9 and
r
so r = 29. This then gives
1
rˆ = ⋅ (2 i − 4 j + 3 k )
29
Example 3.4
What does the vector field defined in 2-D space given by F ( x , y) = x i look
like? To see this, take different points in space and see what the vector field
becomes.
So for the point (1,0), F (1, 0) = 1 i and for any point (1, y), F (1, y) = 1 i .
Similarly, for the point (2, y), F (2, y) = 2 i . This vector field is shown in
Figure 3.5.
x
–4 –3 –2 –1 0 1 2 3 4
b c
b
a a
a
b
–b
d
a
a ⋅ b = a b cos θ = b ⋅ a (3.1)
Note: The scalar product is always written with a dot between the two vectors
being multiplied.
The formula given by Equation 3.1 is useful for the following situations: (1) to
find the angle between two vectors, and (2) to prove if two vectors are perpen-
dicular to each other.
Consider the arbitrary vectors a = a1 i + a2 j + a3 k and b = b1 i + b2 j + b3 k .
To calculate the a ⋅ b term, this is given as
Example 3.5
Given a = 3 i + 2 j + k and b = i + 5 j − 3k , find the angle between the vec-
tors. Using Equation 3.2 first gives
Now, to find the angle θ between the vectors using Equation 3.1 gives
a ⋅ b = a b cos θ
10 = 32 + 22 + 12 ⋅ 12 + 52 + (−3)2 ⋅ cos θ
a⋅b 10 10
cos θ = , cos θ = = = 0.452
a⋅b 14 ⋅ 35 7
10
θ = cos−1 = 63.14°
7
Note: If a.b = 0, then the two vectors are perpendicular, since a and
b ≠ 0, so cos θ = 0, which implies that the angle θ = 90°.
a × b = a b sin θ nˆ (3.3)
where n̂ is a unit normal vector to the plane containing both a and b. There are
two directions perpendicular to any plane, here the direction is such that a, b, and
n̂ form a right-handed set of axes.
Note: The vector product can have a “×” or a “∧” notation between the vectors,
Figure 3.9 shows the arrangement of the vectors.
3.1 Introduction 71
n
b
Figure 3.9 The vector product of a and b lies along the z axis.
Equation 3.3 for the vector product is useful for producing a vector that is per-
pendicular to both the vectors a and b.
3.1.3.3 How to Calculate a × b
If two vectors are given as a = 3 i + 2 j + k and b = i + 2 j − 3k , then the vec-
tor product is given as a × b = (3 i + 2 j + k ) × ( i + 2 j − 3k ). Now this can
be slightly complicated to work out by expanding of the brackets and using
i × i = j × j = k × k = 0, while the cross products follow the right-handed axes
rules. This process can be seen using the cyclic diagram as shown in Figure 3.10.
However, the vector product is more easily calculated using a 3 × 3 determi-
nant. Since what is required is the vector c = a × b, in the determinant the first
row is just the i , j , and k components. Then on the second row will be the com-
ponents of the vector a and finally on the third row the components of the vector
b as follows:
i j k
c =a×b= 3 2 1 = +i 2 1 − j 3 1 +k 3 2
2 −3 1 −3 1 2
1 2 −3
c = −8 i + 10 j + 4 k
Now, if this is a vector perpendicular to both a and b, then this can be checked
using the scalar product of c with a and b. The result should be zero.
Check: a ⋅ c = 0 −33 + 20 + 13 = 0
Also, b⋅c = 0 − 8 + 20 − 12 = 0
This proves that both vectors are perpendicular to this vector c = a × b.
i
i × j = k
j × k= i
k× i = j
etc.
k j
Figure 3.10 Cyclic rotation showing vector products for unit vectors i , j , and, k.
72 Vectors and Geometrical Applications
3.1.4 Projection of Vectors
When talking about the projection of a vector onto another vector, this generally
means the projection of vector a on to the line created by the vector b. The fol-
lowing cases could arise as shown in Figure 3.11.
a or
a
b b
Projb a Projb a
Projb a
First find the length of the projection of a on to b. This is given by using trigonom-
etry. The length of projection onto b is = a cos θ . This is in terms of the angle θ.
But in terms of the vectors a and b, make use of the dot product as follows:
a ⋅ b = a b cos θ
gives
a⋅b
cos θ =
a b
a⋅b a⋅b
a =
a b b
a⋅b
So the length of the projection vector is , but its direction is that of the
b
vector b. Therefore, multiplying the length by a unit vector b̂ will give the vector
projection of a onto b as follows:
a⋅b ˆ
Projb a = b (3.4)
b
3.1 Introduction 73
b
Now, = bˆ can be written out again in a neater form as
b
( )
Projb a = a . bˆ bˆ (3.5)
Note: This can be useful when imagining a fire hose spraying foam from the
ground toward a fire on an oil tank. This gives a way to combine the horizontal
wind speed with the form jet in that direction.
Example 3.6
Find the distance from the point (5, 5) to the line y = 2x shown in Figure 3.13.
In this problem, the use of projections can help to reach a solution.
Solution: To find the perpendicular distance d of the point (5, 5) to the line,
first project the vector (5, 5) onto the line. A vector that describes the line
y = 2x is given as the vector 〈1, 2〉 . You need to find
〈1, 2〉 〈1, 2〉
Proj〈1,2 〉 〈5, 5〉 = 〈5, 5〉.
5 5
15
= 〈1, 2〉 = 3〈1, 2〉 = 〈3, 6〉
5
So the projection of 〈5, 5〉 onto the line is at the point P (3, 6). This is the
closest point to (5, 5) that exists on the line y = 2x.
So the distance required is the distance between two points (5, 5) and
(3, 6), which is just given by the distance between two points formula,
d = ( x1 − x 2 )2 + ( y1 − y2 )2 , that is, d = (5 − 3)2 + (5 − 6)2 = 5 .
y y = 2x
P
d
(5, 5)
3.2 Vector Geometry
3.2.1 Vector Equation of a Line
To find the vector equation of the straight line shown in Figure 3.14, suppose P is
an arbitrary point on a straight line and A and B are given points on the same line
and O is the origin. P has a position vector r, the line is parallel (in the direction)
of AB, and A and B are both points on the line.
B Straight-line
a
b P
OP = OA + AP
r = a + t AB (3.6)
Note: The vector equation of a straight line through two fixed points with position
vectors a and b is given by
r = a + t (b − a) (3.7)
Example 3.7
Find the vector equation of the straight line parallel to the vector i + 2 j − 5k
going through the point with position vector 2 i − 3 j + k .
r = 2 i − 3 j + k + t ( i + 2 j − 5k )
3.2 Vector Geometry 75
3.2.1.1 Intersection of Lines
In 2-D space two lines intersect or are parallel. In 3-D space two lines can inter-
sect or they may be either parallel or skew (i.e., missing each other).
Given two lines in space r1 = a1 + t b1 and r2 = a2 + s b2, then for them to inter-
sect means r 1 = r2 (i.e., a1 + t b1 = a2 + s b2 ), where the unique values of t and s
can be found.
Note: If unique values of t and s cannot be found, then the lines do not intersect
and are said to be skew.
Example 3.8
Find the point of intersection of the lines:
r1 = 2 i + 3 j + t (− i + 2 j ) and r2 = − i + j + s(3 i − 2 j )
2 i + 3 j + t (− i + 2 j ) = − i + j + s(3 i − 2 j )
2 − t = −1 + 3s (i.e. 3s + t = 3)
and
3 + 2t = 1 − 2s (i.e. t + s = −1)
Note: If the lines are in 3-D space, follow the same method except now
there will be three simultaneous equations involving s and t. Solve for s
and t using any two of the three equations and test to see if these fit into the
third equation. If they fit into the third equation, then the lines do intersect.
If they do not fit the third equation, then the lines are said to be skew.
76 Vectors and Geometrical Applications
Example 3.9
Two lines r1 and r2 are given by the equations:
r1 = 3 i + j − 4 k + s(−2 i − 3 j + 6 k )
r2 = 2 i − 3 j + k + t ( i − j − k )
3 − 2s = 2 + s , 1 − 3s = −3 − t , − 4 + 6s = 1 − t
Solving the first two equations gives s = 1 and t = –1, and these fit the
third equation so therefore the lines do intersect.
The point of intersection is given by using s = 1 into equation for r1 and
the point becomes P( i − 2 j + 2 k ).
Example 3.10
Find the equation of the plane with normal vector n = 〈−1, 2, 3〉 and goes
through the point Q (1, 3, 1) as shown in Figure 3.15.
R –1, 2, 3
P (x, y, z)
Q (1, 3, 1)
Figure 3.15 Equation of a plane, normal to a vector and passing through a point.
3.2 Vector Geometry 77
To find this plane, for a given point (x, y, z) the vector PQ and QR must
be perpendicular to each other. So, using the properties of the scalar prod-
uct, then the dot product of QR with PQ must be zero.
Therefore, 〈−1, 2, 3〉 ⋅ 〈 x − 1, y − 3, z − 1〉 = 0
−( x − 1) + 2( y − 3) + 3( z − 1) = 0
a, b, c
(x, y, z)
〈a, b, c〉 ⋅ 〈 x − x0 , y − y0 , z − z0 〉 = 0
a( x − x 0 ) + b( y − y0 ) + c( z − z0 ) = 0
or collecting all the constants onto one side gives the general equation of a plane
as
ax + by + cz = d (3.8)
Example 3.11
Find the equation of a plane that acts as a “mirror” for the points (1,3,–1)
and (1,–1,1) as shown in Figure 3.17.
The plane is perpendicular to the line joining the two points.
(1, –1, 1)
M
y
(1, 3, –1)
x +x y +y z +z
To find the midpoint M, use the midpoint formula 1 2 , 1 2 , 1 2 ,
2 2 2
which gives
1 + 1 3 − 1 −1 + 1
, , = (1,1, 0)
2 2 2
Having a point and the normal vector, the equation of the plane is given
by using Equation 3.8 as
0( x − 1) − 4( y − 1) + 2( z − 0) = 0
4 y − 2z = 4
3.3 Applications
Example 3.12: Resultant Smoke Flow
A small fire in a corridor has smoke flowing vertically at a speed of Vs =
1.3 ms–1. An open window blows wind due east at a speed of Vw = 2.5 ms–1
as shown in Figure 3.18.
The resultant speed VR and direction θ the smoke moves can be obtained
using vectors as follows: VR2 = 1.32 + 2.52 = 7.94 and VR = 2.8 ms–1.
The direction of smoke flow relative to the horizontal is
1.3
θ = tan −1 = 27.5°.
2.5
Vw = 2.5
Vs = 1.3
VR
θ
W = F . d = F d cos θ (3.9)
Here, the cos θ factor in the scalar product allows for the fact that
it is only the component of F along the direction of motion that does
the work.
2. Given two unit vectors û1 and û 2, then the angle between these two
vectors is given by
cos θ = uˆ 1. uˆ 2 (3.10)
θ B
Problems
3.1 Find a unit vector û in the direction of the vector AB, where the points
A and B are given by A( i + 3 j + 5k ) and B(3 i + 2 j + k ).
3.2 Find the angle between the vectors a = 〈1, 2, −4 〉 and b = 〈4, 2, 7〉.
3.3 Find the value of the constant α, which makes the vectors a = 〈1, 2, α 〉
and b = 〈2, 3, 4 〉 perpendicular to each other.
3.4 Figure 3.20 shows a rhombus OACB, with sides OA and OB being given
by the vectors a and b, respectively. Given the diagonals OC = a + b
and BA = a − b, show that the diagonals are perpendicular to each other.
80 Vectors and Geometrical Applications
B
C
a–b
b
a+b
0 A
a
r1 = i + j + 3k + s(3 i + 8 j + 2 k )
r2 = 2i + 5 j − k + t ( i + 2 j + 3k )
3.8 The 2-D rotational flow of a fluid can be represented by the vector field
〈− y, x 〉
F ( x , y) =
x 2 + y2
3.9 Smoke flows vertically upward at a speed of 0.8 ms–1 when it is blown
by wind traveling at 0.6 ms–1 in a southeasterly direction. Find the resul-
tant speed of the smoke flow and its direction relative to the vertical.
4 Determinants
and Matrices
4.1 Background
Matrices with different dimensions are a standard method for solving a wide
range of problems in many disciplines. In science and engineering, when consid-
ering particular problems, the derivation of the solutions to these problems can
come down to solving a system of linear equations. In this chapter, different meth-
ods for solving a linear system of equations will be discussed and these methods
involve the concepts of determinants and matrices.
4.2 Introduction to Determinants
Determinants arise naturally in the solution of a set of linear equations. Also
they will help solve linear equations using the matrix inversion method (see
Section 4.3.6). Starting with a general set of two linear simultaneous equations
as follows:
ax + by = e (4.1)
cx + dy = f (4.2)
where a, b, c, d, e, and f are constants. How do you find the values of x and y?
Since the values of a, b, c, and d are not known, all that can be done is try to
eliminate either x or y.
If one decides to get rid of y, then to make the coefficients of y the same, first
multiply Equation 4.1 by d and Equation 4.2 by b:
81
82 Determinants and Matrices
de − bf
adx − bcx = de − bf ∴ x (ad − bc) = de − bf ∴ x=
ad − bc
af − ce
y=
ad − bc
a b ad − bc
(4.5)
c d
This definition given by Equation 4.5 is just the difference in product of the
diagonals.
Notice also how these answers for x and y contain lots of similar expressions.
Using the definition of a determinant given earlier, they can be written in the fol-
lowing form:
e b a e
f d c f
x= , y=
a b a b
c d c d
or alternatively as
x y 1
= = (4.6)
e b a e a b
f d c f c d
This is called Cramer’s rule. Equation 4.6 is a nice and neat way to represent
the solutions to Equations 4.1 and 4.2.
4.2 Introduction to Determinants 83
4.2.1 2 × 2 Determinants
Now using the definition of a determinant given by Equation 4.5
a b ad − bc
c d
Example 4.1
6 2 = 6 × 4 − 2 × 3 = 24 − 6 = 18
3 4
Example 4.2
−2 −5 = −2 × 4 − (−5) × 3 = −8 + 15 = 7
3 4
4.2.2 Properties of Determinants
Looking at these answers in the above examples, one can notice some facts about
determinants.
a b = a c
c d b d
84 Determinants and Matrices
• If one row (or column) is equal to the other, the value of the determinant
is zero.
a b = 0 = a a
a b c c
• If one row (or column) is a multiple of the other, the value of the deter-
minant is zero.
a b = 0 = a ka
ka kb c kc
100 500 = 0
1 5
3 p 3r p r
= 9 ps − 9qr = 9( ps − pr ) = 9
3q 3s q s
m n 5m n m n
5 = = = 5(mp − on)
o p 5o p 5o 5 p
It can be seen why the last property mentioned earlier comes about:
In fact, there is a much “stronger” property that can be easily proved. First, a
numerical example to show this:
102 504 = 2 4 = − 4 = 6
10
1 5 1 5
4.2 Introduction to Determinants 85
Here the determinant is made much easier to work out by subtracting 100 and 500
from the top row.
This can be made a general rule as follows:
Or the value of a determinant does not change if one adds to one row a multiple
of another row.
This might not seem important for 2 × 2 determinants, but all the preceding
rules apply to determinants of any size, and it is with larger ones that they come
in useful.
4.2.3 3 × 3 Determinants
The basic building block of a determinant is the 2 × 2 one already discussed. All
larger ones are broken down into combinations of 2 × 2 determinants, according
to the rule of signs given as
+ − +…
− + −
+ − +
Example 4.3
Take the determinant
3 4 2
0 1 5
1 6 8
Suppose this is expanded along the top row. The element 3 is multiplied
3 4 2
1 5 −4 0 5 +2 0 1
0 1 5 = +3
6 8 1 8 1 6
1 6 8
= 3(−22) − 4(−5) + 2(−1) = −48
Suppose trying to expand this using a different row; let’s say row 2. Then
the working is similar, but remember that now the expansion starts with a
minus sign this time giving
3 4 2
4 2 +1 3 2 −5 3 4
0 1 5 = −0
6 8 1 8 1 6
1 6 8
= 0(not necessary) + 1(22) − 5(14) = −48
3 4 2
0 1 −5 3 4 +8 3 4
0 1 5 = +2
1 6 1 6 0 1
1 6 8
= 2(−1) − 5(14) + 8(3) = −48
Why all the fuss about being able to use any row or column to evaluate a deter-
minant? Because, as seen, it is very handy if some of the elements are zero. Thus,
using the first column or the second row would be quicker ways of evaluating the
above determinant.
Example 4.4
Evaluate the following determinants:
3 6 2
1. 5 2 1 Expanding along row 3 would speed the working.
0 4 0
4.3 Introduction to Matrices 87
−4 2 7
2. 3 5 0 Expanding along column 3 would help.
1 8 0
Example 4.5
Find the values of k for which
k 3 −5
0 k −1 4 = 0
0 0 k +5
k 3 −5
k −1 4
0 k −1 4 = k = k ( k − 1)( k + 5) = 0
0 k +5
0 0 k +5
4.3 Introduction to Matrices
In English, one thinks of the word matrix as meaning a “grid” or an array.” In
mathematics, it has a special meaning: An “object” made up of a rectangular
arrangement of “things,” or elements, which behaves according to certain rules.
These elements might be numbers, letters, complex numbers, or other things, but
the rules that govern matrix arithmetic are always the same.
When a matrix is written, it is always enclosed in brackets (either curly or
square). This is very important, because there are other rectangular arrays, as
seen in the previous section, called determinants that are written differently, and
the different types must not be confused!
Example 4.6
All the following are matrices
a
1 2+ j −5 − 3 j
0 , b ,
3 2 c j 1 − j
d
Note: Matrices that have a single column (as the second one above) are
called column vectors and those that have a single row are called row
vectors.
88 Determinants and Matrices
4.3.1 Order of a Matrix
Clearly, the preceding matrices have different sizes or “order.” The order of a
matrix is the number of rows by the number of columns it has and is written gen-
erally as m × n, where m is the number of rows and n is the number of columns.
Example 4.7
In Example 4.6, the order of the matrices are as follows: 2 × 2, 4 × 1, and
2 × 2.
If two matrices are equal, clearly they must be the same size and shape;
also corresponding elements must be identical. So if, for example, the two
matrices a b and 3 −5 are equal, then there are four equations
c d 2 1
that must be true: a = 3, b = −5, c = 2, and d = 1.
Example 4.8
2 1 8 6 1 8 8 2 16
+ =
1 5 3 0 2 1 1 7 4
0 6 4 5 3 0 5 9 4
or
2 1 8 6 1 8 −4 0 0
− =
1 5 3 0 2 1 1 3 2
0 6 4 5 3 0 −5 3 4
4.3.3 Matrix Multiplication
4.3.3.1 Multiplying a Matrix by a Scalar
Adding a matrix A to itself, it makes sense to call the result 2A. The effect will be
to multiply each element by 2 as follows:
3 6 3 6 6 12 3 6
+ = = 2
4 5 4 5 8 10 4 5
4.3 Introduction to Matrices 89
k a b = ka kb
c d kc kd
1 2 −3 7
3
down the left-hand side: 3 5 1 −2 This is a 3 × 4 matrix
0 2 6 8
4
then along the bottom
2 6
1 −2 7
If A = and B = 1 8 , then how to determine the product AB,
2 5 4
3 5
2×3 3× 2
Now since inner numbers are the same 3 = 3, that is, n = p, for these two matri-
ces, then this product exists. The answer will be a matrix of order given by the
outer numbers, which is 2 × 2. This can be represented at the moment generally as
1 −2 7 2 6 R1C1 R1C2
1 8 =
2 5 4 RC RC
3 5 2 1 2 2
2×3 3× 2 2×2
90 Determinants and Matrices
This product matrix has four elements in it. These elements are calculated by
seeing the location of each element in the product matrix and carrying out the
corresponding multiplication of the row and column from the original two matri-
ces and adding all the multiplications together.
In the above example, the first element in the product matrix, R1C1 is given by
row 1 of the first matrix multiplying column 1 of the second matrix as follows:
2
R1C1 = (1 − 2 7) 1 = (1) × (2) + (−2) × (1) + (7) × (3) = 2 − 2 + 21 = 21
3
Note: Here the multiplication is carried out with corresponding elements, that
is, the first element of the row with the first element of the column, and so on.
6
R1C2 = (1 − 2 7) 8 = (1) × (6) + (−2) × (8) + (7) × (5) = 6 − 16 + 35 = 25
5
2
R2C1 = (2 5 4) 1 = (2) × (2) + (5) × (1) + (4) × (3) = 4 + 5 + 12 = 21
3
6
R2C2 = (2 5 4) 8 = (2) × (6) + (5) × (8) + (4) × (5) = 12 + 40 + 20 = 72
5
Now these four elements can be put into the result of the product as
1 −2 7 2 6 21 25
1 8 =
2 5 4 21 72
3 5
Finally,
AB = 21 25
21 72
4.3 Introduction to Matrices 91
Example 4.9
Given the following matrices,
1.5
50 100 150 50
2.0 , calculate the product AB.
A= and B =
60 80 25 50 3.5
4.0
Solution: Matrix A has order 2 × 4 and matrix B has order 4 × 1. Since the
inner numbers match, then the product exists and the result will be a matrix
of order 2 × 1, that is, the outer numbers as follows:
1.5
50 100 150 50 2.0 R1C1 1000
= =
60 80 25 50 3.5 R2C1 537.5
4.0
This is the essence of matrix multiplication. Note that in this last exam-
ple the orders of the matrices were
(2 × 4) (4 × 1) giving a (2 × 1) matrix
and that the multiplication could only be done because the “4’s matched”:
the number of columns in the first one matched the number of rows in the
second one.
Could these particular matrices have been multiplied the other way
around, that is, does the product BA exist? By considering the structure
this gives
1 .5
2 .0 50 100 150 50 Because these numbers
=?
3 .5 60 80 25 50 don’t match up, implies can’t
do that matrix multiplication
4 .0
this way round: the product
doesn’t exist.
(4 × 1) (2 × 4)
92 Determinants and Matrices
Example 4.10
3 5 1 0 23 10
=
2 1 4 2 6 2
Example 4.11
2 3 0 4 2 6 23 7
=
1 4 2 5 1 8 24 6
Example 4.12
3 18 6 24
(6 2 8) = 36 12 48
6
1 6 2 8
4.3.4 Special Matrices
As with normal algebra, the number 1 has the property that 1 × a = a × 1 = a.
With matrices, there is a similar situation using the identity matrix I as follows:
1 0 0
1 0 etc.
I2 = and I 3 = 0 1 0
0 1
0 0 1
These are called the identity matrix, I, which have the special property that for
any other matrix A gives A I = I A = A. The identity matrix is always square, but
can be any size, such as 2 × 2, 3 × 3, 4 × 4, and so forth.
4.3 Introduction to Matrices 93
Finally, this section will conclude with other special types of matrices:
• A square matrix could have all its elements equal to zero: this is called
the null matrix.
• If all the elements except those on the main diagonal are zero, then it is
called a diagonal matrix.
• As noted above, the identity matrix, I, is a diagonal matrix with all its
nonzero elements equal to 1.
A= 1 0 5
4 2 1
then
1 4
A = 0 2
T
5 1
3 1 8 3 1 8
B= 1 2 4 BT = 1 2 4
8 4 9 8 4 9
0 −1 −2 0 1 2
C= 1 0 5 CT = −1 0 −5
2 −5 0 −2 5 0
4.3.5 Powers of Matrices
As with normal algebraic expressions, powers of matrices are calculated as
follows:
A2 = A × A
A3 = A × A × A
A4 = A × A × A × A
Example 4.13
Given a square matrix A = 3 5 , calculate A2.
2 1
Solution:
A2 = A × A = 3 5 × 3 5 = 19 20
2 1 2 1 8 11
1 T
A−1 = Ac (4.7)
A
where |A| is the determinant of the matrix A, and AcT is the matrix of cofactors of
A transposed. The cofactors of the matrix A are given by first taking into account
an appropriate sign for each element in the matrix using the following structure:
+ − + …
− + −
+ − +
Then each element is made up of the sign with the determinant of elements that
are left after removing the row and column containing that element.
Generally, if A is a square matrix, then the “minor” of entry aij is denoted by
Mij and is defined to be the determinant of the submatrix that remains after the ith
row and jth column are deleted from A.
4.3 Introduction to Matrices 95
The number (−1)i+j Mij is denoted by Cij and is called the cofactor of entry aij.
The next examples show how this works for a 2 × 2 and 3 × 3 matrix.
Example 4.14
The determinant of a general 2 × 2 matrix
A= a b
c d
Ac = d −c and AcT = d − b
−b a −c a
1 d −b
A−1 = (4.8)
ad − bc −c a
Example 4.15
Calculate the inverse of the matrix A = 2 3 .
1 5
Solution:
A = 10 − 3 = 7
1 5 −3
A−1 = .
7 −1 2
Check
1 7 0 1 0
AA−1 = = =I
7 0 7 0 1
Example 4.16
Following is an example of the inverse of a 3 × 3 matrix using the formula
1 T
A−1 = Ac .
A
96 Determinants and Matrices
0 0 1
A = 2 −1 3
1 1 4
A = 1 2 −1 = 3.
1 1
Matrix of cofactors is
+ −1 3 − 2 3 + 2 −1
14 1 4 1 1
Ac = − 0 1 + 0 1 − 0 0
1 4 1 4 1 1
0 1 − 0 1 + 0 0
+
−1 3 2 3 2 −1
−7 −5 3
Ac = 1 −1 0
1 2 0
−7 1 1
A = −5 −1 2
T
c
3 0 0
−7 1 1
−1 1
So therefore, A = −5 −1 2 .
3
3 0 0
Check to see that AA = I.
−1
2 1 6 15
−2 5 3 = 3
One could say the vector 6 has been transformed into the vector 15
3 3
by the square matrix 2 1 .
−2 5
Both the direction and the length, or magnitude (see Chapter 3, Section 3.1.1),
of the vector have changed. But will this be true for every vector that gets trans-
formed with this matrix?
Clearly, all the matrix multiplications cannot be done by hand. An Excel
workbook could easily be created that shows both vectors and changes the direc-
tion of the “input” vector and the modulus. It is noticed that there are two direc-
tions for which the vectors seem to be “in line with one another,” in other words
the modulus may have changed as a result of the transformation but not the
direction.
If the input vector is any vector along the direction 1 , let’s call it
1
k
, where 0 ≠ k ∈ , then the “output” vector is also along the same direction.
k
Moreover, for any such vector, the output vector is always 3 times as long as the
input, as follows:
2 1 k 3k k
−2 5 = 3 = 3
k k k
Also, if the input vector is any vector along the direction 1 , let’s call it
2
k
, where 0 ≠ k ∈ , then the output vector is also along the same direction.
2k
Moreover, for such vectors, the output vector is always 4 times as long as the
input, for
2 1 k 4k k
−2 5 2 = 8 = 4 2
k k k
These two “multiplying factors,” 3 and 4, are called the eigenvalues of the
matrix 2 1 .
−2 5
Each eigenvalue is associated with a direction; this direction is called the eigen-
vector associated with that eigenvalue. (Eigen is German for “characteristic.”)
98 Determinants and Matrices
Hence the matrix 2 1 has an eigenvalue 3, with associated eigenvec-
−2 5
k
tor , 0 ≠ k ∈ ; and an eigenvalue 4 with associated eigenvector k ,
k 2k
0 ≠ k ∈ .
2 1 x x
−2 5 y = λ y
AX = λ X
∴ λ X − AX = 0
∴ (λ I − A)X = 0
det (λ I − A) = 0,
det λ 0 − 2 1 = λ 2 1 = (λ 2 − 7λ + 12) = 0
−
0 λ −2 5 2 λ−5
∴ (λ − 3)(λ − 4) = 0
so λ = 3 or 4.
For any square matrix A, the expression det (λI – A) is called the characteris-
tic polynomial of A. The characteristic polynomial is sometimes written simply
4.3 Introduction to Matrices 99
Example 4.17
Find the eigenvalues and associated eigenvectors of the matrix
1 0 −1
A= 1 2 1
2 2 3
λ −1 0 1
−1 λ − 2 −1 = 0
−2 −2 λ − 3
1 0 −1 x x
using λ = 1 in AX = λ X : = 1
1 2 1 y y
2 2 3 z z
∴ x − z = x, so z=0
Also, x + 2y + z = y so x = −y
k
So, an eigenvector could be, − k , for any real (non-zero) number k.
0
100 Determinants and Matrices
1 0 −1 x x
using λ = 2 in AX = λ X : = 2
1 2 1 y y
2 2 3 z z
∴ x − z = 2x, so x = −2
also, x + 2 y + z = 2 y, so 2y = 2y
and 2 x + 2 y + 3z = 2 z , so 2 y = − z − 2 x = − z + 2z = z
−2 k
So an eigenvector could be, k , for any real (nonzero) number k.
And finally, 2k
1 0 −1 x x
using λ = 3 in AX = λ X : = 3
1 2 1 y y
2 2 3 z z
k
So an eigenvector could be, − k , for any real (nonzero) number k.
Therefore, the matrix −2 k
1 0 −1
A= 1 2
1
2 2 3
1 −2
has eigenvalues 1, 2, and 3 with corresponding eigenvectors −1 , 1 ,
1 0 2
and −1 .
−2
P = X1 X 2 X n
then
AP = A X1 X 2 Xn = 1 X1 2 X2 n Xn
1 0
AP = X1 X 2 Xn
0 λn
Diagonal eigenvalue
matrix D
D = P −1 AP (4.9)
A = PDP −1 (4.10)
A = PDP −1
A2 = PDP −1PDP −1 = PD 2 P −1
A3 = PDP −1PDP −1PDP −1 = PD 3 P −1
Ak = PDP −1PDP −1PDP −1 … PDP −1 = PD k P −1
Ak = PD k P −1 (4.11)
• All the λ’s are different or distinct, that is, no repeated eigenvalues.
• Having repeated eigenvalues means there may or may not have been
independent eigenvectors.
Example 4.18
Determine the diagonal factorization of a 2 × 2 matrix given by
A= 4 1
−8 −5
AX = λ X
( A − λI )X = 0
det ( A − λ I ) = 0
1 =0
det 4 − λ
−8 −5 − λ
λ 2 + λ − 12 = 0
(λ + 4)(λ − 3) = 0
gives λ1 = −4 and λ2 = 3.
The corresponding eigenvectors are for
λ1 = 3, the eigenvector is X1 = 1 .
−1
λ2 = −4, the eigenvector is X 2 = 1 .
−8
4.4 Solving Systems of Linear Equations 103
So,
P= 1 1
−1 −8
which gives
1
P −1 = − −8 −1
7 1 1
Now
D= 3 0
0 −4
Ak = PD k P −1
A4 = PD 4 P −1
1
A4 = − 1 1 81 0 −8 −1
7 −1 −8 0 256 1 1
−25
A4 = 56
200 281
ax + by = e
cx + dy = f
x y 1
= =
e b a e a b
f d c f c d
The advantage of this method is that only one of the answers needs to be cal-
culated, if that is all that is needed.
Another option is to make use of the inverse matrix method (see Section 4.4.3).
Writing the set of equations in matrix form as AX = B, the solution, provided it
exists, is given by X = A–1B. The advantage of this method is that results can be
obtained for lots of inputs by changing the right-hand set of values and using the
same inverse matrix A–1 each time.
Each method has its advantages of course, but they also have the disadvantage
that they do not give us much information about situations when there is not a
unique solution. As engineers, it is very important to be able to tell what happens
in extreme cases! Also, there may be a need to know when there might be more
than one solution for a set of equations.
Example 4.19
Solve the set of equations:
− x + y + 2z = 2
3x − y + z = 6
− x + 3y + 4z = 4
−1 1 2 2
3 −1 1 6
−1 3 4 4
a b c d
0 e f g
0 0 h i
−1 1 2 2
Let’s get two zeros in the first column of the matrix 3 − 1 1 6 by
−1 3 4 4
Row 2 = Row 2 + 3 Row 1,
(R2: R2 + 3R1) −1 1 2 2
0 2 7 12
and Row 3 = Row 3 – Row 1, 0 2 2 2
(R3: R3 – R1)
Note: When putting a zero in the second column, one must always use row 2 with
a row 3 otherwise using row 1 with row 3 will change the zero value already cre-
ated in the first element of row 3 and so undoing what has already been achieved.
The objective has been achieved, since the nonzero elements start one
place to the right as we go down the rows. The augmented matrix is in
echelon form.
Step 3: The solution can now be found by back-substitution. Imagine the
equations returned to their original form. They would now read
− x + y + 2z = 2
2 y + 7 z = 12
−5z = −10
From the last equation it can be seen that dividing each side by –5 gives
z = 2. Then, looking at the middle equation and using the value for z gives
2y + 14 = 12, so 2y = −2 and y = –1. Finally, moving to the first equation and
using the two values that are known already gives −x + (−1) + 2(2) = 2, from
which x = 1. Putting all the results together gives the solutions as
x = 1, y = −1, z=2.
3x + 2 y + z = 3
2x + y + z = 0
6 x + 2 y + 4 z = −4
3 2 1 3
2 1 1 0
6 2 4 6
It can be seen that the last row contains a contradiction. By writing it out
as an equation it becomes 0 = 12? This tells one that the system of equations
has no solution. It is said that the system is inconsistent.
3x + 2 y + z = 3
2x + y + z = 0
−x − z = 3
3 2 1 3
2 1 1 0
−1 0 −1 3
4.4 Solving Systems of Linear Equations 107
The last row gives nothing useful, since 0 = 0. The second row gives
1 1
− y + z = −2
3 3
or, multiplying each side by 3,
− y + z = −6
Any set of numbers that satisfied these equations would also satisfy the
original set of simultaneous equations. The parameter t can be replaced with
any real number. Hence, there are infinitely many solutions to this system.
The three situations be generalized as (1) a unique solution, (2) no real
solutions, and (3) infinitely many solutions. It’s definitely got something to
do with the last row (or rows) of the echelon form matrix being zero.
The rank of a matrix is the number of nonzero rows when it has been
reduced to echelon form. The three preceding situations finished up looking
like the following:
x + 2y = 4
3x − 5 y = 1
1 2 x 4
3 −5 =
y 1
Denoting
x
A = 1 2 , X = , B = 4
3 −5 y 1
This gives the following matrix equation AX = B. This is the matrix of the simul-
taneous equations. Here the unknown is the matrix X, since A and B are already
known. A is called the matrix of coefficients.
AX = B (4.12)
X = A−1 B (4.14)
This result given by Equation 4.14 gives a method for solving simultaneous
equations. First write the equations in matrix form, calculate the inverse of the
matrix of coefficients A−1, and finally perform a matrix multiplication.
Note: When det(A) = 0 there are no solutions and so the matrix is not invertible
because of division-by-zero problems.
4.4 Solving Systems of Linear Equations 109
Example 4.22
Solve the simultaneous equations
x + 2y = 4
3x − 5 y = 1
1 2 x 4
3 −5 =
y 1
AX = B
Now calculate the inverse of A = 1 2 :
3 −5
1 −5 −2
A−1 = −
11 −3 1
1 −5 −2 4
X = A−1 B = −
11 −3 1 1
1 −22
=−
11 −11
= 2
1
Example 4.23
Solve the simultaneous equations
− x − 2 y + 2z = 1
2x + y + z = 7
3 x + 4 y + 5z = 26
110 Determinants and Matrices
−1 −2 2 x 1
2 1 1 y = 7
3 4 5 z 26
AX=B
1 18 −4
−11
A = −7 −11 5
23
5 − 2 3
1 18 −4 1 23
1 1
X = A−1 B = −7 −11 5 7 = 46
23 23
5 − 2 3 26 69
1
X = 2
3
4.5 Applications
Example 4.24: Fire Modeling Using Markov Chains
A Markov chain is a process that satisfies the Markov property, meaning
one can make predictions for the future of the process based solely on its
present state just as well as one could knowing the process’s full history.
The process involves the probabilities of being in states and can be mod-
eled using matrices. The use of matrices and matrix operations can provide
solutions to problems and is useful to consider them here.
Consider the four different states of a fire as follows: O, fire is out;
S, smoke development; F, flashover; and B, full burning fire. The time steps
are in minutes and Figure 4.1 shows the transition diagram modeling the
fire with transition probabilities between the different states. It is required
to find how long it will take the fire in the different states to go out, that is,
to reach the absorbing state O.
4.5 Applications 111
B F
0.3 0.4
0.1
S
1.0
O
0.6
For this setup, a transition matrix P can be derived from the transition
diagram for the different states as follows,
O S F B
O 1 0 0 0
S 0.6 0 0.4 0
P=
F 0 0.3 0.1 0.6
B 0 0.1 0 0.9
0 0.4 0
Q = 0.3 0.1 0.6
0.1 0 0.9
FM = ( I − Q)−1
where identity matrix I = I3 in this case and the power –1 indicates the
inverse of the (I − Q) matrix.
112 Determinants and Matrices
Constructing
1 0 0 0 0.4 0 1 −0.4 0
( I − Q) = 0 1 0 − =
0.3 0.1 0.6 −0.3 0.9 −0.6
0 0 1 0.1 0 0.9 −0.1 0 0.1
So, to obtain the average time to get from the different fire states to the
absorbing state (fire out) is given by adding along each row as follows:
u1 = Au0
u2 = Au1 = A2u0
u3 = Au2 = AA2u0 = A3u0
u100 = A100u0
Note: This type of difference equation uk+1 = Auk is seen to appear in many
practical applications such as Markov Chains in the area of probabilis-
tic risk. Also, uk could be used to represent the position in an evacuation
model on a floor grid during a fire.
4.5 Applications 113
1 a b c
Lx = Lx T −1 L y T −1 L y T −2
L x : 1 = a, L y : 0 = b + c, T : 0 = − a − b − 2c
Now these form a system of three equations in three unknowns that need
to be solved to find a, b, and c, and hence give the form of relationship for R.
These can be solved by different methods but here the Gaussian elimination
method of Section 4.4.2 is used as follows:
114 Determinants and Matrices
a + b + 2c = 0
b+c= 0
a =1
1 1 2 0
0 1 1 0
1 0 0 1
R3: R3 – R1 gives
1 1 2 0
0 1 1 0
0 −1 −2 1
R3: R3 + R2 gives
1 1 2 0
0 1 1 0
0 0 −1 1
a + b + 2c = 0
b+c= 0
−c = 1
Vx Vy
R∝
g
Problems
4.1 Calculate the following determinants.
4 1 3
5 1
a. b. −1 2 7
3 −2
1 5 0
Problems 115
1 k 2
4.2 Find the value of the constant k for which k −1 0 1 = 0
1 2 0
2 1 5 4 2 1
4.3 Let A = and B = , calculate the following.
1 3 −2 3 −1 7
a. A + B
b. A − B
4.4 Given
4 1
A = 3 2 and B = 2 −3
5 0 1 4
4.5 Solve the following set of simultaneous equations using the Gaussian
elimination method.
a. x + 2y + z = 7 b. x+ y−z = 2
3x + y + 4 z = 5 x + 2y = 5
2 x + 3 y − z = 14 2x + 3y − z = 7
4.6 An electrical network has currents i1, i2, and i3 given by the following set
of equations:
i1 + 3i2 + i3 = 9
2i1 − i2 + 5i3 = 7
4i1 + 2i2 − i3 = 11
Using the matrix inversion method, find the currents in the network.
2 1 3
A=1 2 3
3 3 20
4.8 Given the 2 × 2 matrix A = 2 1 , determine the following.
9 2
a. Diagonal Factorization of the matrix A.
4.9 Consider again Example 4.24 with the four states of fire: O, fire is out;
S, smoke development; F, flashover; and B, full burning fire. The time
steps are given in minutes and Figure 4.2 shows the transition diagram
modeling the fire with transition probabilities between the different
states. Calculate how long it will take the fire in the different states to
reach the absorbing state O (i.e., to go out).
B F
0.2 0.5
0.1
S
1.0
O
0.5
Figure 4.2 Transition diagram modeling a fire with the probabilities between
states.
5 Complex Numbers
5.1 Background
The origins of complex numbers came about when mathematicians were trying to
solve certain types of equations. Equations of the form x2 = 9, which had two real
solutions x = 3 and x = −3, were of no problem. But when the equation was of the
form x2 = −9, this has no real solutions since the square of a real number cannot
be negative. This was a problem. The idea to overcome this was to extend the real
numbers with the imaginary unit called j = −1, where j2 = −1, so that solutions
to equations such as x2 = −9 could now be found.
Originally, the symbol i was used to represent −1, but in many engineering
fields i was the symbol for electrical current and so the symbol j is used in engi-
neering instead.
− b ± b 2 − 4 ac
x= (5.1)
2a
Notes:
• If the number under the square root sign is positive, there are two real and
distinct solutions.
• If the number under the square root sign is zero, there is one repeated solution.
• If the number under the square root sign is negative, there are no real solutions.
117
118 Complex Numbers
In the next section, it will be seen that in the third case the solutions can in fact
be written as complex numbers.
Now, what if the quadratic equation is given as x2 + 6x + 10 = 0? Using the
formula given by Equation 5.1 with a = 1, b = 6, and c = 10 gives
−6 ± 36 − 40 −6 ± −4
x= =
2 2
Unfortunately, the calculator cannot work out −4 . In fact, with “normal” (or
real) numbers, there is not any number that when multiplied by itself can give a
negative result.
But suppose there was an invented “number,” call it j, such that j2 = −1. What
then? It can be seen that
(2 j)(2 j) = 4 j 2 = 4(−1) = −4
and
−6 ± 2 j
x= = −3 + j or − 3 − j
2
5.2.1 Some Properties of j
If j2 = −1, then j3 = −j, j4 = 1, j5 = j, and so on. You could also write
−4 = ±2 j, −9 = ±3 j, −16 = ±4 j, and so on.
5.2.2 Complex numbers:
A complex number is a number like 2 + 3j. The real part of 2 + 3j is 2. The imagi-
nary part of 2 + 3j is 3. (Note: The imaginary part doesn’t include j.)
In any equation with complex numbers, the real part of one side has to equal
the real part of the other side; similarly for the imaginary parts. There are two
equations rolled up as one. If it is given that x + yj = 3 − 4j, then it can be seen
straight away that x = 3 and y = −4.
5.3 Arithmetic Operations
5.3.1 Addition and Subtraction
It is easy to add and subtract complex numbers. It goes just like ordinary algebra;
add the real parts together and add the imaginary parts together.
5.3 Arithmetic Operations 119
Example 5.1
1. (3 + j) + (1 + 2j) = 4 + 3j
2. (2 − 3j) − (1 + j) = 1 − 4j
3. (1 + 4j) + 2(5 − j) = 11 + 2j
5.3.2 Multiplication
Multiplying complex numbers is also simple, as long as you remember that j2 = −1.
Example 5.2
1. j(3 + j) = 3j + j2 = −1 + 3j
2. (2 + j)(4 + 5j) = 2(4 + 5j) + j(4 + 5j) = 8 + 10j + 4j + 5j2 = 3 + 14j
5.3.2.1 Conjugate Numbers
When multiplying complex pairs such as (2 − j)(2 + j) = 5 and (1 + 6j)(1 − 6j) =
37 notice here that each time that the result is a real number. The “partner” of
the complex number is called its conjugate. So the conjugate of 2 − j is 2 + j. The
conjugate of 2 + j is 2 − j. The conjugate of −5 − 7j is −5 + 7j. The conjugate of
100 j is −100 j. The conjugate of 27 is 27. The conjugate of a + bj is a − bj. Also
notice that the result of multiplying a number by its conjugate is always real and
positive. Generally, the following is found:
(a + bj)(a − bj) = a 2 + b 2
5.3.3 Division
When it comes to division of complex numbers, it makes no sense to be dividing
by imaginary numbers. This idea of the complex conjugate helps out here as an
equivalent division can be done such that the denominator is now a real number
as seen in the next example.
Example 5.3
1 1(2 − j) 2− j 2− j 2 1
= = = = − j = 0.4 − 0.2 j
2 + j (2 + j)(2 − j) 22 + 12 5 5 5
Im
2 + 3j
Re
2 – 3j
5.4 Argand Diagram
5.4.1 Drawing a Diagram of Complex Numbers
It is often helpful to draw the complex plane, or Argand diagram, and represent
the complex number 2 + 3j, say, as a vector starting at the origin and finishing at
the point (2,3) as shown in Figure 5.1.
Note: Adding and subtracting complex numbers follows the rules of vector addi-
tion and subtraction.
Im
4 + 3j
5
3
36.9°
Re
4
is written in shorthand to 5 ∠ 36.9°. (The length is called the modulus. The angle
is called the argument.)
Note: The angle can be expressed in several ways, for example, –30° could be
written as 330°; 200° could be written as –160°.
a + bj
r
b
θ°
Re
a
Note: Most calculators can easily swap a number from one form of a complex
number to the other using the Pol and Rec function buttons.
Or when multiplying two numbers in polar form, just multiply their lengths
and add their angles.
122 Complex Numbers
Similarly,
5 ∠25° 5
= ∠(25° − 60°) = 2.5 ∠ (−35°)
6 ∠60° 2
Or when dividing two numbers in polar form, just divide their lengths and
subtract their angles.
5.5.2 Exponential Form
Using Euler’s identity, ejθ = cos θ + j sin θ. Then, z = r (cos θ + j sin θ) = r ejθ. In
this form r is the modulus and θ is the argument in radians. This is known as the
exponential form of a complex number.
Note: The proof of Euler’s identity uses the Taylor series expansions of the expo-
nential, cosine, and sine functions, and is omitted here.
Because of the way multiplication works with complex numbers in polar form,
this can be worked out easily as
(5 ∠25°)4 ≡ 54 ∠ (4 × 25°)
Or
The result works for both positive and negative whole number powers and is
used for finding the roots of equations in the next section.
5.6 Roots of Equations
De Moivre’s theorem can also be used with fractional powers to find square roots,
cube roots, and so forth of complex numbers. There’s a difference, however:
This time, there should be more than one answer. To see why, try calculating
z3 if z = 2 ∠60°.
From De Moivre,
z 3 = {2 ∠60°}3 = 8 ∠180°
and this can be expressed in rectangular form as –8. If it is drawn on the complex
plane, it will look like Figure 5.4.
But suppose one tries the same thing with the complex number w = 2 ∠300°.
Then z3 = 8 ∠900°, and turning this answer into rectangular form, it’s the same
as before: –8. Are there any other numbers that would give the answer –8 when
cubed? Without knowing anything about complex numbers, one can clearly say
that (−2)3 = −8. Therefore, –2 can be written as a complex number in polar form
as 2 ∠180°.
So there appears to be three different complex numbers that, when cubed, give
the answer –8. These are as follows: z1 = 2 ∠60°, z2 = 2 ∠180°, z3 = 2 ∠300°. These
1
are the three cube roots of –8, or the three values of (−8) 3 . If these are drawn on
the complex plane, they will look like Figure 5.5. Notice, how they are equally
spaced (120°) around a circle.
The procedure to find all the different values for the roots is very straightfor-
1
ward. Suppose one needs to find z 4 , if z is the complex number 81 ∠160°. This
can be done by completing the following steps:
Im
Re
–8 0
Im
Re
0
1 1
Step 1: z 4 = {81 ∠160°}4 Make sure the numbers are in polar form.
1
Step 2: = {81 ∠(160° + 360° n)}4 Add 360°n to the angle.
1
1
Step 3: = 814 ∠ (160° + 360° n) Using De Moivre.
4
= 3 ∠(40° + 90° n) Tidying up.
3 ∠40°
3 ∠130°
Step 4: = Taking the values of n = 0, 1, 2, 3…
3 ∠220°
3 ∠310°
Notes:
• Starting with a number that is not in polar form, the first task is to turn it
into polar form.
• Having found the set of answers in polar form, these can easily be turned
into rectangular form if required.
• The principal root is defined as the one closest to the positive real axis.
Example 5.4
Determine all solutions of the equation z3 − 125 j = 0, giving your answers
in rectangular form.
Solution:
z3 − 125 j = 0
∴ z3 = 125 j Rearranging into standard equation form.
1
1
∴ z = (125 j) 3
Taking both sides to the power .
1
3
= {125 ∠90°}3 Change the number into polar form.
1
= {125 ∠(90° + 360° n)}3 Add 360° n to the angle.
1
1
= 125 ∠ (90° + 360° n)
3
Using De Moivre’s theorem.
3
= 5 ∠(30° + 120° n) Tidying up.
5 ∠30°
= 5 ∠150° Taking enough values of n = 0, 1, 2, to get all
5 ∠270°
solutions.
4.33 + 2.5 j
= −4.33 + 2.5 j Change into rectangular form if required.
−5 j
5.7 Applications
The use of complex numbers occurs in many areas of science and engineering,
and one such application is in determining conditions for flashover fires and disas-
ters to occur. Another important area is in electrical circuit theory with alternat-
ing currents, resistor, inductor, and capacitor circuits.
dT
K = a2T 2 − a1T + 1 (5.3)
dt
parameters a1 and a2. From theory it turns out that if the roots of this qua-
dratic equation are complex in nature, then a flashover fire will occur.
As an example if a1 = 0.01 and a2 = 0.2 then the quadratic equation on
the right-hand side (RHS) becomes
0.2T 2 − 0.01T + 1 = 0
0.01 ± −0.7999
T=
0.4
which gives the two solutions as T = 0.025 ± j 2.236, which are two complex
conjugate roots. Hence for these parameter values the fire would develop
into a flashover state and have possible disastrous consequences.
1
XC =
wC
X L = wL
V
AC
i(t) L
Im
j
XL
+90°
Re
0
–90° R
XC
Figure 5.7 The different phases for the component on the complex plane.
Z = 4 + j10 + 3 + (− j6)
Z = 7 + j4
Now to find the magnitude and phase angle of the total impedance Z, this
complex impedance can be converted into polar form to give Z = 8.06 ∠ 29.7°,
which gives Z = 8.06 Ω.
Note: Other quantities like the current in the circuit can be found using
V
complex number division, that is, using the relationship = .
Z
128 Complex Numbers
4Ω j 10 Ω
AC i(t)
100 V 3Ω
–j6Ω
Figure 5.8 An electrical series circuit with resistors, inductor and capacitor.
Im
F2 = 60 N F1 = 50 N
45°
60°
Re
15°
F3 = 40 N
Im
FR = 37.95 N
26.72
44.75°
Re
–26.95
Problems
5.1 Solve the following quadratic equations:
a. x2 + 4x + 5 = 0
b. 2x2 − x + 7 = 0
c. 3x2 + 6x + 11 = 0
a. z1 + z2
b. z1 − 3z2
c. z1z2
z1
d.
z2
5.3 Solve the following equation to find the two pairs of values of a and b:
(a + 3j) (4 − bj) = 37 + 9j.
5.4 Find the roots of the following equations, giving your answers in rect-
angular form.
a. z3 = 5 + 4j
b. z 5 − 2 + 3 j = 0
130 Complex Numbers
c. z 2 = 18 3 − 18 j
d. z3 + 27 j = 0
b. Hence, find the size of the current i in the circuit, giving the answer
in milliamps.
R = 32 Ω
XC = 15.92 Ω
XL = 62.83 Ω
5.6 For a room fire the equation for the temperature T against time t is given
by Equation 5.3 as
dT
K = a2T 2 − a1T + 1.
dt
6.1 Differentiation
6.1.1 Definition of a Limit
For a straight line, to find the slope (or gradient) of the line, which is the same
across the whole line, it is the change in y against the change in x (Figure 6.1).
y
y = mx + c
Δy
Δx
change in y ∆y
m= = (6.1)
change in x ∆ x
What happens if we have some general curve as shown in Figure 6.2. What is
meant by the slope of this curve? Clearly, the slope is changing at different points.
131
132 Introduction to Calculus
y = f (x)
So, the slope at a given point P can be found by saying that this is also the slope
of the tangent line at that point, as shown in Figure 6.3.
y
y = f (x)
Tangent line
Figure 6.3 The slope of a curve as the slope of the tangent line.
At different points the slope would be different because of the different tangent
lines at those points. To find the slope at the different points derivatives and lim-
its are used. Consider the curve given in Figure 6.4. To find the slope at point P,
another point Q near to P is considered. The slope of the line joining P to Q, that
is, the secant line, is an approximation to the slope at P. So, the closer Q gets to
P the better the approximation of the slope at P. The slope of the line PQ is given
by using Equation 6.1 as
f ( x + h) − f ( x ) f ( x + h) − f ( x )
Slope of PQ = = (6.2)
x+h − x h
Now by letting h → 0, the line PQ becomes the tangent line at P to the curve,
which is the just slope of the curve at the point P. Therefore, the slope at the point
P is defined as
6.1 Differentiation 133
y
y = f (x)
P(x, y)
x
x x+h
dy f ( x + h) − f ( x )
Slope = = lim (6.3)
dx h→ 0 h
This is the definition of the derivative, that is, the slope of a curve at a particular
point P. This definition can be used to find slopes of different curves and to derive
a general formula.
Example 6.1
Review the curve y = x2 in Figure 6.5. Using the formula given by Equation
6.3, the slope of PQ is given as
dy ( x + h) 2 − x 2 2 xh + h 2
Slope = = lim = lim = lim (2 x + h)
dx h→0 h h→0 h h→0
y = x2
Q(x + h, (x + h)2)
P(x, x2)
x
x x+h
Therefore, as h → 0,
dy
Slope = = 2x
dx
for the curve y = x2.
This process can be generalized for any function of the form y = axn.
Similar analysis gives
dy a( x + h)n − ax n
Slope = = lim
dx h→0 h
n(n − 1) n− 2 2
( x + h)n = x n + nx n−1h + x h + O(h 3 + higher )
2!
n(n − 1) n− 2 2
a x n + nx n−1h + x h + O(h 3 + higher ) − ax n
dy
Slope = = lim 2 !
dx h→0 h
So,
dy n(n − 1) n− 2
= lim a nx n−1 + x h + O(h 2 + higher )
dx h→0 2!
Now taking the limit as h → 0, the second and higher terms become zero
and this gives for any curve given of the form y = axn the derivative as
dy
= anx n−1 (6.4)
dx
This is now a general formula for finding the slopes of curves for powers of x.
Example 6.2
Given
dy dy
y = x7 , find . Solution: = 7x 6
dx dx
dy dy
y = 3x 5 , find . Soolution: = 15 x 4
dx dx
dy dy
y = 2 x −3 , find . Solution: = −6 x −4
dx dx
dy dy
y = x, find . Solution: =1
dx dx
dy dy
y =1 find . Solution: =0
dx dx
6.1 Differentiation 135
Example 6.3
dy
If y = x5, find , then the gradient at the point (2, 32) on this curve.
dx
Solution:
dy
= 5x 4
dx
dy
At the point (2, 32), the gradient is = 5 × 24 = 80.
dx
dy
Note: Finding is called differentiating y with respect to x.
dx
Example 6.4
dy
If y = x6 + 3x4 − x2 + x, find.
dx
When differentiating, simply add and subtract the separate differentiated
dy
terms. So, in this case, becomes
dx
dy
= 6 x 5 + 12 x 3 − 2 x + 1
dx
Example 6.5
1
If y = 2 x 2 − 6 x −4
then the slope becomes
1
dy −
= x 2 + 24 x −5.
dx
1
When the function is written in terms of x or
, then to begin with it is
x2
probably best to rewrite it with powers that are fractions or negative num-
bers, then use the rule given by Equation 6.4.
136 Introduction to Calculus
1 1
= x −2 x = x2
x2
1 1 −
1
= x −3 =x 2
x3 x
1 1 −
m
= x −n =x n
xn n
m
x
dy
=0
dx
Example 6.6
Find the highest point on the graph of y = 5 + 2x − x2.
dy
= 2 − 2x
dx
6.1 Differentiation 137
x
0 4
dy
= 2 − 2x = 0
dx
Solving gives x = 1.
To determine the y coordinate, substitute the x value into y = 5 + 2x − x2
to give y = 6. So, the highest point on this graph is (1, 6).
The gradient will also be zero if our graph has a lowest point, like the
graph of y = x2 − 4x as sketched in Figure 6.7.
At a maximum or minimum point,
dy
=0 (6.5)
dx
6.1.2.1 Practical Test
The practical test looks at what’s happening to the gradient on either side of the
dy
point. Taking the previous example, y = 5 + 2x − x2 with = 2 − 2 x , you can
dy dx
construct Table 6.1. Here the slope is calculated on either side of the station-
dx
ary point x = 1. By looking at the shape of the slope formed, this indicates that the
point (1, 6) is a maximum point.
138 Introduction to Calculus
dy
Sign of 2 0 –2
dx
d dy d 2 y d2y
= is negative at a maximum,
, that is, <0 (6.6)
dx dx dx 2 dx 2
On the other hand, consider a minimum point shown in Figure 6.9. This can
be summarized as
d dy d 2 y d2y
= 2 is positive at a minimum, that is, >0 (6.7)
dx dx dx dx 2
What happens if the second derivative is zero? Well, there might be a point of
inflection, that is, neither a maximum or minimum point, but still the slope is zero
as shown in Figure 6.10.
The easiest way to check it out is to make a table, as in the first method of the
practical test.
6.1.3.1 Products
This type of function consists of one part multiplied by another; it is a product of
the two parts. The way we do it is to differentiate only one part at a time.
If y = (1 + 3x)(2x4 + 6x), then
dy
= (1 + 3x)(8x3 + 6) + (3)(2x 4 + 6x)
dx
plus
put down the first part differentiate the first part
dy dv du
=u +v (6.8)
dx dx dx
140 Introduction to Calculus
Example 6.7
dy
If y = (x2 − 2)(x2 + 1), find , simplifying your result.
dx
Solution:
dy
= ( x 2 − 2)(2 x ) + (2 x )( x 2 + 1) = 4 x 3 − 2 x
dx
6.1.3.2 Quotients
Sometimes there may be functions like
x2 + 1
y=
x2 − 1
dy
How do you find in this case?
dx
Well, there’s a formula that can be used similar to the one for the product rule.
It is a little more complicated than the one for products, but quite simple to use.
u
It is called the quotient rule. Expressed mathematically, if y = , where u and v
are functions of x, then v
du dv
dy v dx − u dx
= (6.9)
dx v2
minus
(put down the bottom part)(differentiate the top) — (put down the top)(differentiate the bottom)
(put down the bottom)2
Note: Start with the bottom part; remember it’s a minus sign linking the terms on
the top. The squared term on the bottom is not usually multiplied out.
Example 6.8
5x + 1 dy
If y = , find . Simplify your results (top only).
5x + 3 dx
6.1 Differentiation 141
Solution:
dy (5 x + 3)(5) − (5 x + 1)(5)
=
dx (5 x + 3)2
dy 10
=
dx (5 x + 3)2
6.1.4 Standard Functions
So far, the slope dy when y is some power of x, a product and a quotient have been
dx
considered. It is now time to consider how to differentiate such functions as e3x,
sin x, and cos x etc.
The following results can all be proved from first principles, but for the
moment it is better to just make use of them. There is no need to learn these,
but there will be a need to be able to look them up and use them, both with and
without substituting numbers. Table 6.2 shows some standard functions that are
commonly used.
Examples 6.9
dy
1. If y = x2 + 3ex, find when x = 2.
dx
Solution:
dy
= 2 x + 3e x = 26.2
dx
dy
2. If y = 4x + sin x, find when x = 1. (Be careful!
dx
Remember to
Solution: use radians
when working
dy these out.)
= 4 + cos x = 4.54
dx
142 Introduction to Calculus
Clearly, it’s not difficult to differentiate such functions. What about using
the rules for product and quotient? They apply just as well to these functions
as they did to powers of x. But do make sure the correct rule is being used.
To recap again, there are two arrangements to watch out for they are the
product and quotient.
dy dv du
Product: y = u.v then use =u +v
dx dx dx
du dv
v −u
u dy dx dx
Quotient: y = then use =
v dx v2
In the following, more examples using the product and quotient rule are
given.
Example 6.10
If y = x2 e3x, then
dy
= ( x 2 )(3e3 x ) + (2 x )(e3 x ) = 3 x 2e3 x + 2 xe3 x
dx
Example 6.11
ln x
If y = , then
x
1
( x ) − (1)(ln x )
dy x 1 − ln x
= =
dx x2 x2
Next is a more complex problem to show how to make use of these
derivatives.
Example 6.12
2x
Find any maximum or minimum points on the graph of y = .
ex
dy
Solution: Remember, the procedure is to first find , then put this equal to
dx
zero and solve it to find x. Next find the y-values that correspond with the
x-value(s), and last decide which sort of stationary point it is (maximum,
minimum, or point of inflection).
(This function is a quotient, so use the quotient rule when differentiating.)
2x dy
Starting with y = , find .
ex dx
dy (e x )(2) − (e x )(2 x ) 2(1 − x )
= =
dx (e x )2 ex
6.1 Differentiation 143
dy 2
At the stationary point, = 0 and so 1 − x = 0 giving x = 1 and y = .
dx e
So, the stationary point is at 1, 2 .
e
To check what kind of point it is, you could use the second derivative
test, but here it is simpler to use the practical test about the point x = 1 as
shown in Table 6.3.
Table 6.3 Slope of the Curve about the Stationary Point
Value of x L (x = 0) (x = 1) R (x = 2)
dy
Sign of 2 0 –0.27
dx
2
So, the point 1, is a local maximum point and the graph of this func-
e
tion is shown in Figure 6.11.
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.5 1 1.5 2 2.5 3 3.5 4
x
2x
Figure 6.11 Graph of the function y = .
ex
x2 + 5x + 1 x9
Example 6.13
dy
Find for the following functions.
dx
1. y = (4x − 5)10
dy
Solution: = 10(4 x − 5)9 (4) = 40(4 x − 5)9
dx
2. y = sin (x2)
Solution: dy
= cos( x 2 )(2 x ) = 2 x cos( x 2 )
dx
6.2 Integration
6.2.1 Introduction and the Riemann Sum
b
What does
∫ f (x) dx represent? An incomplete answer is to say it is the area under
a
the curve shown in Figure 6.13. Yes, it does give the area under the curve, but the
6.2 Integration 145
y = f (x)
x
a b
area under the curve is just an application of integration. There are many appli-
cations of integration, as will be shown in the later chapters. A more complete
b
Δx
a xp* b
Consider the interval from a to b. Take this interval and chop it up into pieces of
varying lengths. Then for each piece pick a point. Then evaluate the function f at
these points and then sum over all the pieces multiplied by the width of the piece.
So,
b
∫ f (x) dx lim
∆x → 0 ∑
all pieces
f ( x*p )∆ x (6.10)
a
Example 6.14
Find the area under the curve y = f(x) from a to b as shown in Figure 6.15.
y = f (x)
Total Area ≈ A1 + A2 + A3 + A4
A1 A2 A3 A4
x
a Δx b
lim
∆ x→0 ∑
all pieces
f ( x*p )∆ x =
∫ f (x) dx = Area
a
F (x + ∆ x ) − F (x)
F ( x )′ = lim
∆ x→0
∆x
F(x)
y = f (t)
t
a x b
x+∆ x x
F ( x )′ = lim
∫ a
f (t ) dt −
∫ a
f (t ) dt
∆ x→0 ∆x
x+∆ x x x+∆ x
∫
a
f (t ) dt −
∫ f (t) dt = Shaded area = ∫
a x
f (t ) dt
Therefore,
x+∆ x
1
F ( x )′ = lim
∆ x→0 ∆ x
x
f (t
∫) dt
The mean value theorem of definite integrals expresses that there exists a c
where (x ≤ c ≤ x + Δx) such that
x+∆ x
So,
x+∆ x
1
f (c) =
∆x ∫ x
f (t ) dt
F(x) y = f (t)
f (c)
t
a x c x + Δx b
F ( x )′ = lim { f (c)}
∆ x→0
Now, having this, we can compute ∫ f (x) dx. This is now given by
a
If y = xn, then
x n+1
∫ x n dx =
n +1
+C (6.12)
Note: In words, this particular rule becomes “add one to the power and divide
by the new power.”
Example 6.15
x5
1.
∫ x 4 dx =
5
+ C
3x 6 x6
2.
∫ 3 x 5 dx =
6
+ C =
2
+ C
x1.5
3.
∫ x 0.5 dx =
1.5
+ C
x −1 1
4.
∫ x −2 dx =
−1
+ C = − + C
x
6.2 Integration 149
These are called indefinite integrals as the value of the constant C is unknown.
For definite integrals, when there are limits of integration, use the for-
mula given by Equation 6.11.
b
Example 6.16
2 x dx = [ x 2 + C ]0 = (12 + C ) − (0 2 + C ) = 1 − 0 = 1
∫
1
Note: The C has canceled out. This will always happen with a definite inte-
gral, and so there is no need to write it.
2 x dx = [ x 2 ]0 = 12 − 0 2 = 1 − 0 = 1
∫
1
x
2 5
Example 6.17
1. Find the area under the curve y = 10x − x2 between the x-values 2
and 8.
8 8
x 2 x 3 8 2 83 2 2 23
∫
2
(10 x − x ) dx = 10 − = 10 − − 10
2
2 3 2 2 3 2
−
3
= 132
2. Find the area under the curve y = sin x between the x-values 0
and π.
Note: One, or both, of the limits may be negative, as long as the graph stays
above the x-axis.
So, the required area is 22.6667 square units (to four decimal places).
This area is shown in Figure 6.19.
20
18
16 It is found that the
area under the
14
graph, between
12 the two vertical
y 10 lines, is 22.67
8 square units.
6
4
2
0
–4 –2 2 x 4
What happens if the graph goes below the x-axis? Everything below the
axis results in a negative contribution to the integral, so if there are
parts above and parts below the axis, then there is a need to work out
the upper and lower bits separately. Figure 6.20 shows a cubic function.
4
2
–3 –2 –1 1 2 x 3 4 5
0
–2
–4
–6
y –8
–10
–12
–14
4. Find the total area enclosed by the graph and the x-axis between
x = −1 and x = 4.
Solution: First, work out the area above the x-axis between x = −1 and x = 0:
0 0
x4 x3 x2 3
∫
−1
( x − 3 x − 4 x ) dx =
3 2
4
− 3 − 4 = {0} − − = 0.75
3 2 −1 4
Now work out the area below the x-axis between x = 0 and x = 4:
4 4
x4 x3 x2
∫
0
( x 3 − 3 x 2 − 4 x ) dx =
4
− 3 − 4 = {−32} − {0} = −32
3 2 0
The integral is negative, but the area is positive; so this area is 32 square
units.
Adding together these two areas, the total area enclosed by the graph and
the x-axis over the interval −1 ≤ x ≤ 4 is 32.75 square units.
6.2.4 Improper Integrals
b
But sometimes there can be the situation as shown in Figure 6.21. f(x) is a
b
(i) f (x)
x
a ∞
f(x)
(ii)
b
∫a f (x) dx–interval is finite but
the function is not bounded
x
a b
f (x)
1
y=
x2
x
1 ∞
1
Figure 6.23 The function y = is finite but unbounded.
x2
Example 6.18
1
Consider the function given by the curve y = , as shown in Figure 6.23.
x2
Find the area under the curve from x = 1 to x = ∞.
∞
1
A=
∫x 1
2
dx
The following describes how to deal with these types of integrals (one of
the limits is infinite ∞).
Let x go up to some value x = t.
∞ t
1 1
∫
1
x2
dx = lim
t → ∞ ∫x
1
2
dx
t t t
1 x −1
= lim
t → ∞ ∫
1
x2
dx = lim
t → ∞ ∫
1
x −2 dx = lim
t → ∞ −1 1
t
−1 −1 −1 1
= lim = lim − = lim 1 −
x t 1 t →∞ t
t → ∞
1 t → ∞
∞
1 1
So, as t → ∞,
t
→ 0 . Therefore,
∫x
1
2
dx = 1
∞
1
∫x
1
2
dx
Example 6.19
+∞
Calculate
∫x e
−∞
2 − x3
dx .
+∞ 0 +∞
∫x e
−∞
2 − x3
dx =
∫x e
−∞
2 − x3
dx +
∫x e
0
2 − x3
dx
0 v
lim
t → −∞ ∫x e
t
2 − x3
dx + lim
v → ∞ ∫x e
0
2 − x3
dx
Start with
0 0
−e − x 3 −1 1 − t 3
lim
t → −∞ ∫x e
t
2 − x3
dx = lim
t → −∞ 3 t
= lim
t →
m
−∞
+ e
3 3
− t3
As t → −∞, e → ∞, so
∫x e
−∞
2 − x3
dx → ∞
+∞
∫x e
−∞
2 − x3
dx → ∞
6.3 Integration Techniques 155
Example 6.20
There is an infinite discontinuity in the following integral. There is a prob-
lem when x = 0 in the integrand.
9
1
Calculate
∫
0
x
dx.
Here, replace the zero with the symbol a and let this a → 0.
9 9
1 1
∫x
−
lim 2 dx = lim + 2 x 2 a = lim+ 2 9 − 2 a
a → 0+ a → 0 a→ 0
a
as a → 0, a → 0.
Therefore,
9
1
∫
0
x
dx = 6
6.3 Integration Techniques
6.3.1 Substitution
In this method, the idea is to replace the complicated part of the integral by letting
it equal a new variable, say, u. This then involves changing the integral over to a
u variable problem with respect to du.
2
Take the first example,
∫ x + 3 dx. Write a new letter, say, u, to stand for the
expression x + 3. Now, one must be very careful to keep track of the three vari-
ables x, y and u. The working will look like this after the substitution for u and
replacing for dx.
u = x+3
2 2
∫ ∫
du
dx = du = 2 ln(u) = 2 ln( x + 3) + C ∴ =1
x+3 u dx
∴ du = dx
The reader could check back by differentiating this result and the result
2
is indeed .
x+3
156 Introduction to Calculus
Example 6.21
Calculate
∫ 4 cos(100πx) dx .
First, make the substitution. Let u = 100πx and also replacing for dx
gives
du 4
∫ 4 cos(100πx) dx = ∫ 4 cos(u) 100π = ∫ 100π cos(u) du u = 100 πx
du
4 ∴ = 100π
= ( − sin(u)) + C dx
100 π
du
4 ∴ = dx
=− sin(100 πx ) + C 100π
100 π
Example 6.22
6x
∫x 2
+3
dx
6x 6 x du 3 u = x2 + 3
∫ x2 + 3
dx =
∫ .
u 2x
=
∫ u
du
∴
du
= 2x
= 3 ln(u) + C dx
du
= 3 ln( x 2 + 3) + C ∴ = dx
2x
Finally, definite integrals (those with limits) can be done by the sub-
stitution method. Of course, one can do it just as earlier, using the limits
at the end to calculate the answer. But it is much quicker to shortcut a
little by changing the limits so that one does not return to the variable
x at all.
Example 6.23
Consider the integral with limits as
5
2x
∫x
1
2
+7
dx
5 32
2x 2 x du u = x2 + 7
∫1
x +7
2
dx =
∫
8
.
u 2x
∴
du
= 2x
32 dx
du
∫ u = [ ln u ]
32
= du
8 ∴ = dx
8
2x
= ln(32) − ln(8) when x = 1, u = 8
= ln(4) when x = 5, u = 32
6.3.2 Partial Fractions
If one is faced with an integral that contains a fraction, first check to see if the top
is the derivative of the bottom of the fraction. For example:
2x 3x 2 + 4 x x
∫ x +3
2
dx ,
∫ x3 + 2x 2 − 5
dx ,
∫x 2
+4
dx
Note: In the case of the last example, the exact of the derivative on the top does
not exist, but it can be made to do so as follows,
x 1 2x
∫x 2
+4
dx =
2 ∫x 2
+4
dx
and now this can be done like the others using the method of substitution.
If, however, the top does not contain the derivative of the bottom, and if
the bottom part can be factorized, then try using integration using partial
fractions.
The different types of forms of partial fractions splits are summarized as
follows:
3x + 1 A B C
= + +
( x + 1)( x + 2)( x − 3) x + 1 x + 2 x − 3
158 Introduction to Calculus
1 A B C
= + +
( x + 2)( x − 1)2 x + 2 x − 1 ( x − 1)2
3x + 1 A Bx + C
= + 2
( x − 1)( x + 1) x − 1 x + 1
2
5x − 1 A B
= +
( x + 1)( x − 2) x + 1 x − 2
5x − 1 A B A( x − 2) + B( x + 1)
= + =
( x + 1)( x − 2) x + 1 x − 2 ( x + 1)( x − 2)
3. Put the tops of the right- and left-hand sides equal. This follows, since
the bottoms are equal.
5 x − 1 ≡ A( x − 2) + B( x + 1)
4. Substitute values for x that make each bracket equal to zero. This will
give the values for A, B, and so on.
In this case, the first bracket can be made zero if x = 2, and the second
bracket is zero if x = −1.
Letting x = 2:
Letting x = –1:
Now, the partial fraction split has finished. Next, write the partial frac-
tion out as
5x − 1 2 3
= +
( x + 1)( x − 2) x + 1 x − 2
1 A B C
= + +
( x + 2)( x − 1) 2
x + 2 x − 1 ( x − 1)2
Now proceed as before. First, make the right-hand side into one big fraction:
1 A B C A( x − 1)2 + B( x + 2)( x − 1) + C ( x + 2)
= + + =
( x + 2)( x − 1)2 x + 2 x − 1 ( x − 1)2 ( x + 2)( x − 1)2
Then make the top parts equal, since the bottom parts are the same:
1 = A( x − 1)2 + B( x + 2)( x − 1) + C ( x + 2)
And now need to think of which numbers we could substitute for x that
will make the brackets equal to zero. If x = 1, then
1
So, 1 = 3C, which gives C = .
If x = −2, then 3
1
So, 1 = 9A, which gives A = .
9
Now, to find B, either substitute any other number for x and make use of
the values of A and C, or look at the powers of on each side of the equation
to equate coefficients of the powers of x.
160 Introduction to Calculus
Start with
1 = A( x − 1)2 + B( x + 2)( x − 1) + C ( x + 2)
1
[ x 2 ]: 0 = A+ B = +B
9
1
∴ B=−
9
The work is done; just write out the partial fraction in full:
1 1 1
−
1
= 9 + 9 + 3
( x + 2)( x − 1)2 x + 2 x − 1 ( x − 1)2
This is more useful written with the constants at the front as follows:
1 1 1 1 1 1 1
= − +
( x + 2)( x − 1)2 9 x + 2 9 x − 1 3 ( x − 1)2
3x + 1 A Bx + C
= + 2
( x − 1)( x + 1) x − 1 x + 1
2
Now proceed as before, first making the right-hand side into one big
fraction:
3x + 1 A Bx + C A( x 2 + 1) + ( Bx + C )( x − 1)
= + =
( x − 1)( x 2 + 1) x − 1 x 2 + 1 ( x − 1)( x 2 + 1)
Now putting the top parts equal, since the bottom parts are the same:
3 x + 1 = A( x 2 + 1) + ( Bx + C )( x − 1)
6.3 Integration Techniques 161
This time, there will only be one number that can be substituted for x to
obtain A.
If x = 1, then
[ x 2 ]: 0 = A+ B
∴ B = −2
All three numbers A, B, and C are found and now the partial fraction can
be written out in full as
3x + 1 2 −2 x + 1
= + 2
( x − 1)( x + 1) x − 1 x + 1
2
3
1. ∫ x + 1 dx
2
2. ∫ ( x + 4) 2
dx
2x + 5
3. ∫x 2
+1
dx
The first two can be done by substitution, although the results are different.
The last integral has to be tackled in a different way. Another standard integral
must be used, which again can be looked up in a standard table of integrals as
shown in Table 6.2 and is reproduced here.
1 1 x
∫x 2
+a 2
dx = tan −1 + C
a a
(6.13)
2x + 5
∫x 2
+1
dx
162 Introduction to Calculus
what must be done here is split the top of the fraction and make two parts out of
the expression as follows:
2x + 5 2x 5
∫x 2
+1
dx =
∫x 2
+1
dx +
∫x 2
+1
dx
The first part can be done by substitution again; the second part can be
done using the standard integral given by Equation 6.13, where a2 = 1, and
so a = 1. Putting all this together gives the solution as
2x + 5 2x 5
∫x 2
+1
dx =
∫x 2
+1
dx +
∫x 2
+1
dx = ln( x 2 + 1)) + 5 tan −1 x + C
6.3.3 Integration by Parts
Sometimes there is a product of two functions that needs integrating as follows:
∫xe x
dx
There is a formula that is derived from the product rule and can be used to find
the integral of a product of two functions, that is,
dv du
∫ u. dx dx = u.v − ∫ v. dx dx (6.14)
dv
To use this formula, the term, which is , must be able to find its inte-
dx
gral easily. Now, if both terms are easy to integrate, then any powers of x is
chosen to be the u term.
Example 6.27
Evaluate
∫xe x
dx
dv
Since both functions are easy to integrate let u = x and then = ex.
dx
du
For the right-hand side of Equation 6.14, you need the and the v terms.
dx
du
= 1 and v = e x
dx
so,
∫ xe dx = xe − ∫ e dx
x x x
6.3 Integration Techniques 163
∫ xe dx = xe
x x
− e x + C = ( x − 1)e x + C
Example 6.28
Evaluate
∫ x ln(x) dx
Now clearly the ln(x) is not easy to integrate so it must be the u term, that
dv
is, u = ln(x) and then = x.
dx
du 1 x2
= and v =
dx x 2
x2 x2
∫ x ln( x ) dx =
2
ln( x ) +
4
+C
Area =
∫ {(Upper graph) − ( Lower graph)} dx
a
(6.15)
Example 6.29
Review Figure 6.24. To find the bounded area between the curves as shown
in Figure 6.24, simply find the points of intersection of the two curves by
equating the y values as follows:
4 − x 2 = x 2 − 2x
x2 − x − 2 = 0
so x = −1 and x = 2.
Performing the integration as given by Equation 6.15 gives
2 2
Area =
∫ (4 − x ) − ( x
−1
2 2
− 2 x ) dx =
∫ (−2x
−1
2
+ 2 x + 4) dx = 9 square units.
164 Introduction to Calculus
8
This shows the graphs of
6
y = 4 – x2
y = x2 – 2x
4
–2 –1 0 1 2 3
x
–2
–4
6.4 Applications
Example 6.30: Practical Optimization Problem
In some real practical situations, there may be some maximization or mini-
mization problems. Consider the following example of constructing an
open cylindrical tank using minimum material.
An open tank that has vertical sides and a circular base is to be constructed
from metal so as to use the minimum amount of material. If the capacity of
the tank is to be 8 m3, find the dimensions of the tank shown in Figure 6.25.
Solution: The volume of the tank gives πr2h = 8. The surface area of the
tank is given as S = πr2 + 2πrh. You need to find the value of r (and h),
which makes S a minimum.
From the first equation, h can be defined as
8
h=
πr 2
16
S = πr 2 + .
r
dS
Now the problem is one of minimizing S with respect to r. Find = 0.
Differentiating S with respect to r gives dr
dS 16
= 2πr − 2 = 0
dr r
8
r3 =
π
which gives
1
83
r = = 1.37
π
8
h=
π (1.37)2
d 2S 32
= 2π − 16(−2)r −3 = 2π + 3
dx 2 r
Q̇
˙ = 0.01 t2
Q
t
0 300
a
dQ
Q=
∫ dt dt
0
(6.16)
To calculate the heart released in the first 300 seconds of a fire with the
heat-release rate given as a t-squared function, Q (t ) = 0.01t 2, using the for-
mula given in Equation 6.16 gives
300 300
0.01 t 3 (0.01)(300)3 270000
Q=
∫
0
0.01 t dt =
2
3 0
=
3
=
3
= 90000 J = 90 KJ
P(T ≤ t ) =
∫ f (s) ds,
0
Problems 167
Now, the reliability of a component means the probability that the sys-
tem has not failed in the interval [0, t]. So,
R(t ) = 1 − P(T ≤ t )
t ∞
It can be seen that integrals can play an important role in calculating physi-
cal quantities that can be used to predict system behavior.
Another important application in the reliability of systems is that given
the failure probability density function f(t), then the expected time to failure
E(T) can be calculated using the formula
E (T ) =
∫ t f (t) dt
0
Now, since f(t) is some function of time and so is t, this integral will be a
product of two functions of time. For certain functions of f(t), this will need
to be done using the integration by parts method discussed in the earlier
sections.
The expected value is a measure of when something will typically fail.
This is an important measure as this allows for maintenance of systems to
take place in advance of this time, which will prevent a system from failing
and can also save lives and money.
Problems
dy
6.1 Find for each of the following:
dx
a. y = 4x3 − 5x2 + 7x − 11
b. y = 8 x ( x + 5)
c. y = x2 e3x
sin x
d. y =
x
e. y = (x3 + 11)5
6.2 Find the coordinates of the stationary points for the curve with equation
16
y = 8 − x2 −
x2
168 Introduction to Calculus
π 500
P = 2+ r +
2 r
2r
a.
∫ (4 x 3
+ 9 x 2 − 10 x + 4) dx b.
∫ x (3 x + 5) dx
x5 + 1
c. ∫ x2
dx
d.
∫ 6x x 2 + 5 dx
6
2x + 1
e.
∫ (x − 4)(x + 2) dx
5
f.
∫ xe 4x
dx
g.
∫ e sin x dx
x
h.
∫x n
ln x dx
6.5 Given that the initial fire growth is governed by the heat release rate
Q (t ) = 0.3t 2 , determine the heat released by the fire during the first 10
seconds of burning.
Problems 169
6.6 Given the failure probability density function f(t), the time t being mea-
sured in operator work years as follows:
1
(t − 1), 1≤ t ≤ 9
f (t ) = 32
0, otherwise
7.1 Background
In many physical situations, equations arise that involve differential coefficients
dy d 2 y dy
such as , , etc. These equations are called differential equations since
dx dx 2 dt
they contain differential coefficients. Differential equations arise naturally in the
modeling of real phenomena in engineering. The following example shows some
different areas in which they can occur.
Example 7.1
1. If a body, for example, a sprinkler droplet, falls freely under gravity
g, the distance traveled s is given by using Newton’s second law of
motion as
d 2s
=g (7.1)
dt 2
di
L + Ri = E (7.2)
dt
171
172 Ordinary Linear Differential Equations
E(t)
AC
i(t) L
d 2x dx
m 2
+r + sx = 0 (7.3)
dt dt
Note: In all of these cases the problem is to find the dependent variable in terms
of the independent one. For example, from Equations 7.1 to 7.3, respectively, find
s in terms of t, find i in terms of t, or find x in terms of t.
Example 7.2
dy
x = 3y (7.4)
dx
7.2 Types of Differential Equations 173
Example 7.3
d2y dy
+ 2 + y = sin x (7.5)
dx 2 dx
Example 7.4
d3y d2y dy
3
+ 8 y2 2 + 2 + y = 3 (7.6)
dx dx dx
d3y
The order is 3 because the highest derivative is the third derivative 3
term. dx
Example 7.5
3
dy
dx + y = x
2
(7.7)
dy
The order is 1 because the highest derivative is the first derivative term.
dx
7.2.3 Degree of a Differential Equation
The degree of a differential equation is the power to which the highest derivative
is raised.
In Example 7.5, Equation 7.7 has order 1 but the degree is 3, since the highest
dy
derivative is being raised to the power 3.
dx
7.2.4 Linearity
A differential equation is linear if it is linear in the dependent variable y and its
derivatives. The differential equations in Examples 7.2 and 7.3,
dy d2y dy
x = 3y and 2
+2 + y = sin x
dx dx dx
3
d3y d2y dy dy
3
+ 8 y2 2 + 2 +y=3 and dx + y = x
2
dx dx dx
3
d2y dy
are both nonlinear because of the y 2 and terms being present.
dx 2 dx
Consider the differential equation given in the following example.
Example 7.6
dy dy
x3 +4 + y = x5 (7.8)
dx dx
This equation is still linear since it is linear in y and its derivatives the x
powers are not a problem, because x is the independent variable not the
dependent variable.
Example 7.7
dy
If = 2x ⇒ y = x 2 + C
dx
These expressions are the same relationship in different forms.
Converting a differential equation into a direct equation between y and x is
called solving the differential equation.
7.3.1 Simplest Situation
dy
If the differential equation can be arranged in the form = f ( x ), then this type
dx
of equation can be solved by direct integration and there is generally no compli-
cations involved. The following example shows this simplest case.
Example 7.8
Solve
dy
x = 6x3 + 7 (7.9)
dx
dy 7
= 6x 2 + (7.10)
dx x
dy 7
∫ dx dx = ∫ 6x 2
+ dx
x
(7.11)
y = 2 x 3 + 7 ln x + C (7.12)
7.3.2 Separating Variables
dy
If the differential equation is not of the simple type, that is, = f ( x ), then it
dx
dy
may be the case that the differential equation is of the form = f ( x , y). Then the
dx
variable y on the right-hand side prevents solving by direct integration.
In this case, the function f(x,y) can be split into two separate functions as
follows:
It may be the case that f(x,y) can be made into a separable form as follows:
dy
= f ( x , y) = F ( x ).G ( y) (7.13)
dx
or
dy F (x)
= f ( x , y) = (7.14)
dx G ( y)
If the differential equation can be written as either Equation 7.13 or 7.14, then
it is possible to separate the variables x and y, and then rearrange the equations to
integrate separately as shown in the following examples.
Example 7.9
Solve
dy 2x
= (7.15)
dx y + 1
dy F (x)
= f ( x , y) = (7.16)
dx G ( y)
dy
( y + 1) = 2x (7.17)
dx
Equation 7.17 is now of the form in which the variables are separated.
Now integrating both sides of Equation 7.17 with respect to x gives
dy
∫ ( y + 1) dx dx = ∫ 2x dx (7.18)
dy
∫ ( y + 1) dx dx =
∫ 2x dx (7.19)
∫ ( y + 1) dy = ∫ 2x dx (7.20)
y2
+ y = x2 + C (7.21)
2
7.3 First-Order Differential Equations 177
Example 7.10
Suppose we require the equation of a curve that satisfies the following dif-
ferential equation:
dy cos x
2 = (7.22)
dx y
dy
∫ 2y dx dx = ∫ cos x dx (7.23)
dy
∫ 2y dx dx =
∫ cos x dx (7.24)
∫ 2y dy = ∫ cos x dx (7.25)
y 2 = sin x + C (7.26)
4 sin 0 + C ⇒ C = 4
y 2 = sin x + 4 (7.27)
Example 7.11
Solve
dy
= (1 + x )(1 + y) (7.28)
dx
1 dy
∫ (1 + y) dx dx = ∫ (1 + x) dx (7.29)
178 Ordinary Linear Differential Equations
1 dy
∫ (1 + y) dx dx =
∫ (1 + x) dx (7.30)
1
∫ (1 + y) dy = ∫ (1 + x) dx (7.31)
x2
ln(1 + y) = x + +C (7.32)
2
Example 7.12
Solve
dy y 2 + xy 2
= (7.33)
dx x 2 y − x 2
Separating the variables requires more work in this problem. First a com-
mon factor of y2 on the numerator and x2 on the denominator of Equation
7.33 can be factored out to give
dy y 2 (1 + x )
= (7.34)
dx x 2 ( y − 1)
( y − 1)
To separate the variables, multiply Equation 7.34 by on both sides
y2
and then integrating gives
( y − 1) dy (1 + x )
∫ y 2 dx
dx =
∫ x2
dx (7.35)
( y − 1) dy 1+ x
∫ y 2 dx
dx =
x2
dx
∫ (7.36)
y −2 −2 x
∫ y 2 − y dy = ∫ x + 2 dx
x
(7.37)
7.3 First-Order Differential Equations 179
1 1
∫ y − y −2
dy = ∫ x −2
+ dx
x
(7.38)
ln y + y −1 = − y −1 + ln x + C (7.39)
Tidying up both sides gives the solution of the differential Equation 7.33 as
1 1
ln y + = ln x − + C (7.40)
y x
dy
= f (x) (7.41)
dx
dy
= F ( x ).G ( y) (7.42)
dx
dy
+ P ( x ) y = Q( x ) (7.43)
dx
d
(……) = Q
dx
I.F. = e ∫
P ( x ) dx
(7.44)
dy ∫ P ( x ) dx
+ Pye ∫ = Qe ∫
P ( x ) dx P ( x ) dx
e (7.45)
dx
d
dx
(
ye ∫
P ( x ) dx
= Qe ∫)P ( x ) dx
(7.46)
ye ∫
∫
= Qe ∫
P ( x ) dx P ( x ) dx
dx (7.47)
The solution to the differential of the form given by Equation 7.43 can be writ-
ten in a form that is easier to remember as follows:
∫
y( I.F.) = Q( I.F.) dx (7.48)
where
I.F. = e ∫
P ( x ) dx
(7.49)
Example 7.13
Solve
dy
+ 5 y = e2 x (7.50)
dx
dy
+ P ( x ) y = Q( x ) (7.51)
dx
7.3 First-Order Differential Equations 181
I.F. = e ∫ = e∫
P ( x ) dx 5 dx
= e5 x
I.F. = e5 x (7.52)
ye5 x =
∫e 2x 5x
e dx (7.53)
ye5 x =
∫e 7x
dx (7.54)
e7 x
ye5 x = +C (7.55)
7
Now dividing both sides of Equation 7.55 by e5x gives the final solution as
e2 x
y= + Ce −5 x (7.56)
7
Example 7.14
Solve
dy
( x + 1) + y = ( x + 1)2 (7.57)
dx
dy
+ P ( x ) y = Q( x ) (7.58)
dx
dy 1
+ y = ( x + 1) (7.59)
dx x + 1
182 Ordinary Linear Differential Equations
1
This implies that P( x ) = and Q(x) = x + 1.
x +1
1
∫ dx
I.F. = e ∫
P ( x )dx
= e x +1 = e ln(x +1) = x + 1
∴ I.F. = x + 1 (7.60)
y( x + 1) =
∫ (x + 1) dx
2
(7.61)
( x + 1)3
y( x + 1) = +C (7.62)
3
( x + 1)2 C
y= + (7.63)
3 x +1
d2y dy
a( x ) 2
+ b ( x ) + c( x ) y = f ( x ) (7.64)
dx dx
dy d2y
y′ = and y′′ =
dx dx 2
a( x ) y′′ + b( x ) y′ + c( x ) y = f ( x ) (7.65)
If the right-hand side of Equation 7.65 is identically zero (i.e., f(x) = 0), then
a( x ) y′′ + b( x ) y′ + c( x ) y = 0 (7.66)
7.4 Second-Order Differential Equations 183
This is called the homogeneous equation as it only contains terms in y and its
derivatives.
Linear constant coefficient second-order differential equations are a spe-
cial case of Equation 7.65 where a(x), b(x) and c(x) are all constants (i.e., only
numbers). So, the differential equations to be solved in this section are of the
following type:
The task is to find these two functions yCF and yPI separately and then add
them together to get the final solution. The next section considers how to com-
pute the complementary function yCF.
Note: A second-order linear differential equation always has two and only two
linearly independent complementary solutions.
Then y1(x) and y2(x) are both solutions to Equation 7.70, that is,
Note: Equation 7.73 is called the general solution for the complementary
function.
y = e mx (7.75)
y′ = me mx , y′′ = m 2e mx (7.76)
am 2e mx + bme mx + ce mx = 0 (7.77)
e mx (am 2 + bm + c) = 0 (7.78)
Since emx does not equal zero implies for Equation 7.78 to be equal to zero, then
the quadratic equation must be equal to zero, that is,
am 2 + bm + c = 0 (7.79)
Example 7.15
Solve the following second-order linear constant coefficient differential
equation:
y′′ + 3 y′ + 2 y = 0 (7.80)
7.4 Second-Order Differential Equations 185
m 2e mx + 3me mx + 2e mx = 0 (7.81)
e mx (m 2 + 3m + 2) = 0 (7.82)
m 2 + 3m + 2 = 0 (7.83)
y1 = e −2 x and y2 = e − x (7.84)
y = α e −2 x + β e − x (7.85)
7.4.2 Types of Solutions
It has already been shown that by assuming an exponential solution of the form
y = emx to the constant coefficient equation ay″ + by′ + cy = 0 gives rise to the
characteristic or auxiliary equation am2 + bm + c = 0, which is solved for m
in order to fully determine the general complementary solution to the equation
ay″ + by′ + cy = 0.
The general solution to the characteristic equation am2 + bm + c = 0 is given by
− b ± b 2 − 4 ac
m= (7.86)
2a
m = m1 and m = m2
186 Ordinary Linear Differential Equations
where
− b + b 2 − 4 ac − b − b 2 − 4 ac
m1 = and m1 =
2a 2a
Example 7.16
Solve
y′′ + 5 y′ + 6 y = 0 (7.88)
m 2 + 5m + 6 = 0
⇒ (m + 2)(m + 3) = 0
m = −2 and m = −3
yCF = α e −2 x + β e −3 x
b
m1 = −
2a
Note: If one solution was q1 = e m1t , then the other would be q2 = te m1t.
Example 7.17
Solve
y′′ + 6 y′ + 9 y = 0 (7.90)
yCF = α e −3 x + β xe −3 x
m1 = p + jq and m2 = p − jq
b 4 ac − b 2
with p = − and q = .
2a 2a
Then the solution to the complementary function is
yCF = α e( p+ jq ) x + β e( p− jq ) x (7.91)
(
yCF = e px α e jqx + β e − jqx ) (7.92)
e jθ = cos θ + j sin θ
Therefore, the general solution for the complementary function can now be
written as
Example 7.18
Solve
y′′ + 4 y′ + 9 y = 0 (7.94)
m = −2 ± j 5
that is, where f(x) = 0 has been made. Now considering the full differential
equation:
ay′′ + by′ + cy = f ( x )
Table 7.1 shows which functions for the particular integral to try for the differ-
ent functions of f(x).
Example 7.19
If f(x) = sin 4x, then try the particular integral as yPI = C sin 4x + D cos 4x.
If f(x) = f(x) = x + 2ex, then try the particular integral as yPI = Cx + D + Eex.
Finally, putting all the theory together for the complementary function
and the particular integral, the next example shows how to solve a general
second-order differential equation.
Example 7.20
Solve
y′′ − 5 y′ + 6 y = 2 sin 4 x (7.99)
y′′ − 5 y′ + 6 y = 0 (7.100)
m 2 − 5m + 6 = 0
⇒ (m − 2)(m − 3) = 0
m = 2 and m = 3 (this is case 1 for different types of solutions) implies the solu-
tion is
yCF = α e 2 x + β e3 x
190 Ordinary Linear Differential Equations
′ = −4 A sin 4 x + 4 B cos 4 x
yPI
Now by substituting yPI, yPI′ and yPI″ back into Equation 7.99 gives
Collecting all the cos 4x and sin 4x terms together on the left-hand side gives
Here, the right-hand side has a 0 multiplying the cos 4x, so like terms can be
compared on both sides.
By equating the coefficients of the cosine terms on both sides gives
−10 A − 20 B = 0 ⇒ − 10 A = 20 B
⇒ A = −2 B (7.102)
−10 B + 20 A = 2 ⇒ − 10 B = 2 − 20 A
⇒ − 5 B = 1 − 10 A (7.103)
2 1
A= and B=− (7.104)
25 25
2 1 1
yPI = cos 4 x − sin 4 x = (2 cos 4 x − sin 4 x ) (7.105)
25 25 25
7.5 Applications 191
1
y(t ) = α e 2 x + β e3 x + (2 cos 4 x − sin 4 x ) (7.106)
25
Now to finally finish off the complete solution, the constants α and β need to be
27
found using the given initial conditions. The initial conditions are x = 0, y = ,
25
117
and y′ = 50
.
27
Using x = 0 and y = in Equation 7.106 gives
25
α +β =1 (7.107)
1
y′(t ) = 2α e 2 x + 3β e3 x − (8 cos 4 x + 4 sin 4 x ) (7.108)
25
117
Using x = 0 and y ′ = in Equation 7.108 gives
50
2α + 3β = 2.5 (7.109)
1
y(t ) = 0.5e 2 x + 0.5e3 x + (2 cos 4 x − sin 4 x ) (7.111)
25
The next section is of applications that show how differential equations arise
in real-world engineering problems and how using the different methods studied
so far can help solve these real problems.
7.5 Applications
Example 7.21: Calculating the Time for the Smoke Layer to Develop
In this problem, as the fire progresses, the dangerous smoke layer develops
as shown in Figure 7.2. The time for the smoke layer to reach a certain
height can be calculated by solving a differential equation. In Figure 7.2,
room height is H and lower layer height is z. The fire is treated as a point
source of heat Q. The mass flow rate of the lower layer to the upper layer is
given by m p, the plume mass flow rate. The plume is considered only as a
means of transporting mass from the lower layer to the upper layer.
192 Ordinary Linear Differential Equations
H
z
Fuel array
Applying the conservation of mass and energy to the lower layer gives
a simplified differential equation for smoke filling (in dimensionless form)
as follows:
1 5
+ 0.21(Q *) 3 y 3 = 0
dy
(7.112)
dτ
given y = 1 when τ = 0.
z
The dimensionless height given by y = varies from 0 to 1 and gives the
H
fraction of room height below the smoke layer. The dimensionless time is
H2 g
given by τ = t , where t is time in seconds, g equals 9.81 ms–2, H is
S H
the room height, and S the floor area.
Equation 7.112 is an example of a differential equation that can be solved
using the method of separating variables as follows.
1 5
Subtracting 0.21(Q *) 3 y 3 from both sides of Equation 7.112 gives,
1 5
= − 0.21(Q *) 3 y 3
dy
(7.113)
dτ
5
Dividing by y 3 and integrating both sides of Equation 7.113 with respect
to dτ gives
5 1
dy = − 0.21(Q *) 3 d τ
∫ ∫
−
y 3 (7.114)
2 1
y = − 0.21(Q *) 3 τ + C
3 −3
− (7.115)
2
2 1
y = − 0.21(Q *) 3 τ −
3 −3 3
− (7.116)
2 2
7.5 Applications 193
2 1
= 0.14 (Q *) 3 τ + 1
−
y 3 (7.117)
1 2
0.14 (Q *) 3 τ = y 3 − 1
−
(7.118)
This Equation 7.118 can now be used to calculate the time t taken for a
room to fill with smoke to any height.
If a pool of kerosene is ignited releasing 186 kW in a room with floor
area 5.62 m by 5.62 m and a height of 5.95 m, the time until the smoke layer
has filled half the room can be calculated as follows: First, calculate Q*
using the following formula:
Q 186
Q * = 5
= 5
= 0.002 (7.119)
1100 H 2 1100(5.95) 2
For half the room, y = 0.5, and so Equation 7.118 now gives τ as
1 2
−
0.14(0.002) 3 τ = (0.5) 3 −1
H2 g
τ =t = 1.44t
S H
this implies that the time is t ≈ 23 seconds. It therefore takes less than half
a minute for a 186 kW fire to fill half the room with smoke.
R F
Figure 7.3 Two-state dynamical process. The R state represents the system work-
ing and the F state represents not working.
dR
= −λ R + µF (7.120)
dt
where λR is the probability of the system operating at time t and not failing
in the time step dt, and μF is the probability of the system being in the failed
state at time t and being repaired in the time step dt.
Similarly, for the rate of change of the system being in state F is given by
dF
= −µF + λ R (7.121)
dt
Equations 7.120 and 7.121 are first-order differential equations that need
to be solved to find the probabilities of being in each state.
Solution: Starting with the equation for the R state (Equation 7.120) and
having the initial condition that R(0) = R0 (i.e., the initial probability of the
system working)
dR
= −λ R + µF (7.122)
dt
dR
= − λ R + µ (1 − R) = − λ R + µ − µ R (7.123)
dt
This tidies up as
dR
= µ − (λ + µ ) R (7.124)
dt
dR
+ (λ + µ ) R = µ with initial condition R(0) = R0 (7.125)
dt
7.5 Applications 195
dy
+ P(t ) y = Q(t ) (7.126)
dt
I.F. = e ∫
( λ + µ ) dt
= e( λ + µ )t (7.127)
R(t )e( λ + µ )t =
∫ µe ( λ + µ )t
dt (7.128)
µ
R(t )e( λ + µ )t = e( λ + µ )t + C (7.129)
(λ + µ )
µ
R(t ) = + Ce − ( λ + µ )t (7.130)
(λ + µ )
µ
C = R0 −
(λ + µ )
µ µ − ( λ + µ )t
R(t ) = + R0 − e (7.131)
(λ + µ ) (λ + µ )
λ λ − ( λ + µ )t
F (t ) = + F0 − e (7.132)
(λ + µ ) (λ + µ )
These solutions then give the probabilities for the system to be in a work-
ing or not working state.
E0 i(t) L
Using Kirchhoff’s voltage law gives the differential equation for the cir-
cuit as
di
L + Ri = E0 (7.133)
dt
Notes:
• The values of L the inductance, R the resistance, and E 0 the voltage
source are constants.
• Equation 7.133 is a type of differential equation in which there is a
choice of methods that can be used to solve it, either by separating
variables or the integrating factor method. The solution shows the
method of separating variables.
Solution: Starting with Equation 7.133 and subtracting the term Ri from
both sides yields
di
L = E0 − Ri (7.134)
dt
L di
=1 (7.135)
( E0 − Ri) dt
L
∫ E ∫
dt = 1 dt
0 − Ri
(7.136)
Problems 197
L
− ln( E0 − Ri) = t + C (7.137)
R
L
C=− ln E0 (7.138)
R
E0 R
ln = t (7.139)
E0 − Ri L
R
E0 t
= eL (7.140)
E0 − Ri
R
E0 − Ri − t
=e L (7.141)
E0
Rearranging this to make i(t) the subject gives the final solution for the
current in the circuit as
i(t ) =
E0
R
( − t
1− e L
R
) (7.142)
Problems
Solve the following differential equations using the appropriate method.
dy x
7.1 = with y(0) = 1
dx y
dy
7.2 = ( y − 3)( x + 5)
dx
dy 1
7.3 =
dx xy + x
198 Ordinary Linear Differential Equations
dy
7.4 x − y = x3
dx
dy
7.5 (1 − x 2 ) − 2 xy = x 4 with y(0) = 1
dx
d 2 y dy
7.6 − − 2y = x + 2
dx 2 dx
d2y dy
7.7 2
− 10 + 25 y = 10
dx dx
d2y dy
7.8 2
+4 + 5 y = 13e3 x with y(0) = 0 and y′(0) = 1.5
dx dx
dT
7.9 = 1 + 5T with T(0) = 1
dt
1
Ri +
C ∫ i dt = E 0
where R, C, and E 0 are all constants. By differentiating the above
equation, derive a differential equation for the current i(t) in the circuit
t
−
and hence show that i(t ) = ke RC , where k is a constant.
E0 i(t) C
The Laplace Transform is an integral transform named after its founder Pierre-
Simon Laplace. It takes a function of continuous time t (t > 0) to a function of a
complex variable s (frequency). As discussed earlier in Chapter 7, when modeling
real world problems, the formulation of differential equations naturally arises in
many different fields of engineering.
d2y dy
a 2
+ b + cy = E (8.1)
dx dx
199
200 Laplace Transforms
∑a x
0
n
n
= a0 + a1x + a2 x 2 +…
And for some an this series can be written in closed form as, say, A(x):
∑a x 0
n
n
= A( x )
Using a slightly different notation for the coefficients an = a(n) this becomes
∑ a(n) x
0
n
= A( x ) (8.2)
So, different values of a(n) can produce a different closed form sum A(x). Some
examples of using specific a(n)’s are given next.
Example 8.1
If a(n) = 1, then the series just becomes
∑x
0
n
= 1 + x + x 2 +…
This is just the geometric series with first term a = 1 and common ratio
r = x.
This has a sum to infinity given by
a
S∞ =
1− r
1
A( x ) =
1− x
Example 8.2
1
If a(n) = , then the series is
n!
∞
∑ 0
xn
n!
x
= 1+ +
x2
1! 2!
+…
∫ a(t)x
0
t
dt = A( x ) (8.3)
Equation 8.3 could be left in this form, but for integration purposes it is
not good to have x as the base. It is usually better to have e as the function.
This can be done by letting x = eln(x), so xt = (eln(t))t
Now for this integral to converge, as t → ∞, x has be less than 1 and x > 0
to avoid any imaginary numbers appearing. So, 0 < x < 1 implies that ln(x) <
0 in this range for x.
Let –s = ln(x) and replace a(t) = f(t) in Equation 8.3:
∫ f (t)e
0
− st
dt = F (s) (8.4)
This is called the Laplace transform of the function f(t). This is just the
continuous analog of the summation of a power series.
∞
ℒ f (t ) =
∫
0
f (t )e − st dt = F (s) (8.5)
L
f (t) F (s)
L–1
t-domain s-domain
The operator ℒ is used to represent the Laplace transform and the operator
ℒ –1 is used to represent the inverse Laplace transform.
Ordinary
differential Algebraic Inverse
Laplace
laplace Solution
equations transform equation transform
with initial
conditions Y (s) y (t)
8.3.2 Standard Transforms
In the aforementioned process, the differential equation needs to be Laplace
transformed and so this requires Laplace transforming different functions of time
dy d2y
as well as derivative terms like and 2 etc.
dt dt
Most of the Laplace transforms and subsequently the inverse Laplace trans-
forms are generally taken from a standard table of results. Therefore, it is use-
ful to first show some of the more common Laplace Transforms of functions in
a table format and then to see how they are derived from first principles using
the formula definition. Table 8.1 gives some functions and their corresponding
Laplace transforms. All the functions in the table can be proved using the basic
definition given by Equation 8.5 and some further manipulation. Next, a few of
the results are proved to show how they are derived using the formula definition.
8.3 Introduction and Standard Transforms 203
Example 8.3
Laplace transform f(t) = 1.
∞
∞ ∞
e − st 1 1
∫
= 1e
0
− st
dt =
− s 0
= (0) − =
−s s
1
F (s ) =
s
Example 8.4
Laplace transform f(t) = t.
∞
dv du
∫ u. dt dt = u.v − ∫ v. dt dt
dv du e − st
Let u = t = e − st gives =1 and v=
dt dt −s
This gives
∞ ∞ ∞ ∞
e − st e − st e − st e − st 1
F (s ) = t
− s 0
− ∫
0
−s
dt = 0 +
∫
0
s
dt = 2 = 2
− s 0 s
1
F (s ) = 2
s
204 Laplace Transforms
Example 8.5
Laplace transform f(t) = eat.
∞
∞ ∞
e − t( s − a ) 1
F (s ) =
∫e
0
− t( s − a )
dt =
−(s − a) 0
= 0−
− (s − a)
1
F (s ) =
(s − a)
Note: Other Laplace transforms such as sin at and cos at can be done with-
out the need of integration by parts in a much more convenient manner
using Euler’s formula and complex numbers.
Euler’s formula is ejθ = cosθ + j sinθ, so it implies that ejat = cos at + j sin at.
Since this formula contains a real part cos at and imaginary part sin at, then
if one Laplace transforms ejat the real part will be the Laplace transform of
cos at while the imaginary part will be the Laplace transform of sin at. This
is shown in Example 8.6.
Example 8.6
∞
∫e (
− s − ja )t
F (s ) = dt
0
∞
e − ( s− ja )t
Therefore, F (s ) =
−(s − ja) 0
1
F (s ) = 0 − −
s − ja
1
F (s ) = ,
s − ja
Now this needs to be written as a real and imaginary part, that is, a + jb.
One can multiply the top and bottom by the complex conjugate of the bot-
tom (s – ja), that is, by (s + ja), to give
(s + ja) s + ja s a
F (s ) = = 2 = 2 +j 2
(s − ja)(s + ja) s + a 2
s +a 2
s + a2
8.3 Introduction and Standard Transforms 205
So this gives
s
ℒ cos at =
s + a2
2
and
a
ℒ sin at =
s2 + a2
as required.
This follows from the property that the integral of the sum of two functions is
equal to the sum of the two separate integrals. This property of linearity allows
the Laplace transforms of sums of functions to be found easily. The next example
shows how this is can be applied.
Example 8.7
Given f(t) = t and g(t) = e3t, then using Equation 8.6 gives
5 11 −1 1 −1 1
5t − 11e3t = ℒ−1 2 − = 5ℒ 2 − 11ℒ s − 3
s s − 3 s
8.3.4 Basic Relations
There are some basic relations that can be used when solving differential equation
problems of which the most important ones are usually the Laplace transforms
of the first and second derivative of functions. These are properties 6 and 7 in
the Table 8.2 where a is an arbitrary constant. These relationships can be proved
206 Laplace Transforms
∫ f (t) dt
1
8. Integration F (s )
s
0
using the basic definition given by Equation 8.5. The following examples show
how properties 1 and 6 of Table 8.2 are derived.
Example 8.8
ℒ f (t − a) =
∫ f (t − a)e
0
− st
dt (8.7)
∞ ∞
f (u)e ( ) du =
∫ ∫ f (u)e
− s u+ a − su − as
ℒ f (t − a) = e du
0 0
The e−as term can be taken out in front of the integral since it does not
depend on u.
∞
ℒ f (t − a) = e − as
∫ f (u)e
0
− su
du
But
∫ f (u)e
0
− su
du = F (s) , so the following result is obtained:
ℒ f (t − a) = e − as F (s)
8.4 Inverse Transforms 207
Example 8.9
Show that the ℒ[ f ′(t )] = sF (s) − f (0) (Property 6).
∞
dv
u = e − st and = f ′(t )
dt
gives
du
= − se − st and v = f (t )
dt
Therefore,
∞ ∞
∫ f (t)(−se ∫ f (t)e
∞
ℒ f ′(t ) = e − st f (t ) 0 − − st
) dt = 0 − f (0) + s − st
dt
0 0
= − f (0) + s F (s )
Note: This result together with the second derivative transform (property 7)
are used extensively when solving differential equations.
8.4 Inverse Transforms
In the final part of the process of solving differential equations (see Figure 8.1),
one needs to obtain the original function back in the time domain. This will
require once again using the standard table of transforms and inverses. Usually,
the expression that is required to be inverse Laplace transformed cannot be read-
ily obtained from the table as it stands. However, with the use of partial fraction
decomposition (see Chapter 6, Section 6.3.2) expressions can be obtained that can
be inverse Laplace transformed easily.
Next, examples are given on how to find the inverse Laplace transform of some
given functions.
208 Laplace Transforms
Example 8.10
Find
2
ℒ−1 2 2 .
s (s + 4)
This is not readily available as it stands in the standard tables. The frac-
2
tion 2 2 has to be split into simpler fractions using partial fraction
s (s + 4)
decomposition:
2 A B Cs + D
= + + (8.8)
s (s + 4)
2 2
s s2 s2 + 4
1 1
A = 0, B = , C = 0, and D = −
2 2
2 1 1
= −
s (s + 4)
2 2
2s 2
2(s + 4)
2
2 −1 1 −1 1 1
ℒ −1 2 2 = ℒ 2 −ℒ 2 2
s ( s + 4 ) 2 s ( s + 4)
2 1 −1 1 1 −1 1
ℒ −1 2 2 = ℒ 2 − 2ℒ 2
s ( s + 4 ) 2 s (s + 4)
Finally, the inverse transforms are obtained from using the table of stan-
dard transforms:
2 1 1
ℒ −1 2 2 = 2 t − 4 sin(2t )
s ( s + 4 )
8.4 Inverse Transforms 209
Example 8.11
s+4
If F (s) = , find f(t).
(s + 3)(3s − 2)
Solution: First using the partial fractions method gives
s+4 A B
= + (8.10)
(s + 3)(3s − 2) s + 3 3s − 2
s + 4 = A(3s − 2) + B(s + 3)
2
Using easy values of s, that is, s = –3 and s = , gives
3
1 14
A=− and B =
11 11
s+4 −1 1 14 1
= +
(s + 3)(3s − 2) 11 s + 3 11 3s − 2
and so
f (t ) = ℒ −1 [ F (s) ]
s+4 −1 1 14 1
f (t ) = ℒ −1 = ℒ −1 + ℒ −1
(s + 3)(3s − 2) 11 (s + 3) 11 (3s − 2)
1
−1 −1 1 14 −1 3
f (t ) = ℒ + ℒ
11 (s + 3) 11 s − 2
3
Here in the term (3s – 2) the 3 has been taken out of the bracket first and
then using the standard table of transforms gives
2
−1 −3t 14 3 t
f (t ) = e + e
11 33
210 Laplace Transforms
So far, it has been shown how to Laplace transform and inverse Laplace
transform certain continuous time functions. However, in some engineering
problems, the input to the system is discontinuous in time and hence con-
sideration is now given on how to Laplace transform some of these kinds
of input functions.
8.5 Discontinuous Functions
8.5.1 Heaviside Unit Step Function
The unit step function has the effect of switching on or switching off at some pre-
described value of the time t. The function is shown in Figure 8.2. This function
is the Heaviside unit step and is denoted by f(t) = H(t – c), where
f (t ) = 0 :t <c
f (t ) = 1 :t ≥c
ℒ H (t − c) =
∫ H (t − c)e
0
− st
dt
Generally,
e − sc
ℒ H (t − c) = (8.11)
s
f (t)
c t
e− s 0 1
ℒ H (t ) = =
s s
f (t ) = H (t − a) − H (t − b)
e − sa e − sb
ℒ f (t ) = L H (t − a) − H (t − b) = −
s s
f (t)
0 t
f (t)
0 a b t
δ (t − a) = 0, t≠a
δ (t − a) = ∞, t=a
∫
ℒ δ (t − a) = δ (t − a)e − st dt
0
∞
=e − sa
∫ δ (t − a) dt
0
But
∫ δ (t − a) dt = 1 (just the area of the delta function).
0
Therefore, the Laplace of the delta function operating at t = a is
ℒ δ (t − a) = e − sa (8.12)
f (t) δ (t–a)
0 a t
δ (t ) = 0, t≠0
δ (t ) = ∞, t=0
8.6 Shift Theorems 213
δ (t)
0 t
It follows that the Laplace transform of the delta function at the origin with
a = 0 is
ℒ δ(t ) = 1
Note: Use of these types of discontinuous input functions can arise in engineering
problems. For example, in fire these can be used to determine what happens if a
window breaks or a suppression system activates. Hence, knowing their Laplace
Transforms enables solutions to problems as is shown later in the applications
section (Section 8.8).
8.6 Shift Theorems
When working with differential equations with discontinuous input functions, the
final part of inverse Laplace transforming sometimes requires the use of the first
and second shift theorems and so these are stated next:
f (t ) ↔ F (s), then e at f (t ) ↔ F (s − a)
Again, these shift theorems can be derived using the definition formula of the
Laplace transform. These are useful in some cases for finding inverse Laplace
transforms, that is, taking the inverse Laplace transform of Equation 8.13 gives
the following important result:
The following examples show how to make use of the second shift theorem to
find inverses.
Example 8.12
e −4 s
Find a function f(t) whose transform is F (s) = .
s2
Solution: The numerator e–4s corresponds to e–cs (i.e., c = 4 ) in Equation
8.14. Therefore, this indicates that there is a H (t – 4) term present and as
1
F (s ) =
s2
e −4 s
ℒ −1 2 = (t − 4) H (t − 4)
s
Step 4: Finally, determine the inverse Laplace transform to obtain the solution
(with the help of partial fractions and the table of standard transforms).
The following are key relationships that are used frequently in step 1:
ℒ x (t ) = X (s)
dx
ℒ = sX (s) − x (0)
dt
d 2x
ℒ 2 = s 2 X (s) − sx (0) − x (0)
dt
The next few examples show how to apply the preceding methodology to a range
of different differential equations.
8.7 Method for Solving Linear Differential Equations 215
Example 8.13
Find the solution to the first-order differential equation
dx
+ 3 x = e −2t (8.15)
dt
Solution:
dx
ℒ + ℒ[3 x ] = ℒ [ e −2t ]
dt
1
sX (s) − x (0) + 3 X (s) =
s+2
1
sX (s) − 2 + 3 X (s) =
s+2
1
(s + 3) X (s) = +2
s+2
2s + 5
X (s ) =
(s + 2)(s + 3)
2s + 5
x (t ) = ℒ −1 X (s) = ℒ −1
(s + 2)(s + 3)
2s + 5
(s + 2)(s + 3)
1 1 −1 1 −1 1
x (t ) = ℒ −1 X (s) = ℒ −1 + = ℒ (s + 2) + ℒ (s + 3)
( s + 2) ( s + 3)
Example 8.14
Solve the following second-order differential equation:
d2y dy
2
+ 5 + 4 y = 3δ(t − 2) (8.16)
dt dt
Solution:
d2y dy
ℒ 2 + 5ℒ + 4ℒ[ y] = 3 ℒ δ(t − 2)
dt dt
(s 2 + 5s + 4)Y (s) − 2s − 8 = 3e −2 s
3 e −2 s + 2s + 8
Y (s ) =
(s 2 + 5s + 4)
3 e −2 s + 2(s + 4)
Y (s ) =
(s + 1)(s + 4)
3 e −2 s 2
Y (s ) = +
(s + 1)(s + 4) (s + 1)
3
(s + 1)(s + 4)
1 1 −2 s 2
Y (s ) = − e +
(s + 1) (s + 4) (s + 1)
and hence
8.7 Method for Solving Linear Differential Equations 217
Therefore, using the second shift theorem and standard table gives the
solution as
y(t ) = [ e − (t − 2) − e −4 (t − 2) ] H (t − 2) + 2e − t
Example 8.15
This is a more complicated example in which there are two-coupled second-
order differential equations to solve.
Solve the following second-order simultaneous equations:
d 2x
+ 2x − y = 0 (8.17)
dt 2
d2y
+ 2y − x = 0 (8.18)
dt 2
Solution:
d 2x
ℒ 2 + 2ℒ[ x ] − ℒ[ y] = L[0]
dt
d2y
ℒ 2 + 2ℒ[ y] − ℒ[ x ] = L[0]
dt
s 2 X (s ) − 4 s + 2 X (s ) − Y (s ) = 0
s 2Y (s) − 2s + 2 Y(s) − X(s) = 0
218 Laplace Transforms
These are two simultaneous equations in X(s) and Y(s) and we can
eliminate Y(s) to obtain X(s) first.
Multiplying Equation 8.19 by (s2 + 2) and then adding the two equa-
tions and after simplifying an expression for X(s) can be found. Hence,
solving the preceding two equations for X(s) gives
4 s 3 + 10 s
X (s ) =
(s 2 + 1)(s 2 + 3)
3s s
X (s ) = +
(s 2 + 1) (s 2 + 3)
Hence,
3s −1 s
x (t ) = ℒ −1 2 +ℒ 2
( s + 1) + 3)
( s
s −1 s
x (t ) = 3ℒ −1 2 + ℒ
( s + 1 ) ( ( ))
s 2 + 3
2
Now y(t) can be found using the original Equation 8.17. Using
Equation 8.17 and rearranging gives
d 2x
y(t ) = + 2x (8.22)
dt 2
dx
= −3 sin t − 3 sin 3t
dt
d 2x
= −3 cos t − 3 cos 3t
dt 2
8.8 Applications 219
8.8 Applications
Example 8.16: Temperature Variation of the Hot Gas and Smoke Layer
A simple two-zone model of a room splits the room into an upper smoke
layer and a lower layer with the fire as shown in Figure 8.7.
A simple equation for a two-zone model of a fire in a room can be derived
from using the heat balance equation:
Heat in the room = Heat generated from the fire − Heat losses in the room
dT
mc p = Q − Ah∆T (8.23)
dt
assuming that the initial temperature in the room is T0 (i.e., T(0) = T0) and if
radiation is not a dominant phenomenon (i.e., nonflashover fires).
Fuel array
dT Q Ah
= − (T − T0 ) (8.24)
dt mc p mc p
Ah Q
a= and b = + aT0
mc p mc p
into Equation 8.24. This gives the following simple first-order differential
equation:
dT
= b − aT (8.25)
dt
dT
+ aT = b with T (0) = T0 (8.26)
dt
dT
ℒ + aℒ[T ] = ℒ[b]
dt
b
sT (s) − T (0) + a T (s) =
s
b
sT (s) − T0 + aT (s) =
s
b
sT (s) + aT (s) = + T0
s
b + sT0
(s + a)T (s) =
s
sT0 + b
T (s ) =
s (s + a)
8.8 Applications 221
sT0 + b A B
= + (8.27)
s (s + a) s s + a
sT0 + b = A(s + a) + Bs
b b
Using s = 0 gives A = and using s = –a gives B = T0 − . This gives
Equation 8.27 as a a
sT0 + b b 1 b 1
= + T0 −
s (s + a) a s a s + a
So,
b1 b 1
T (s ) = + T0 − (8.28)
as a s + a
T (t ) = ℒ −1 T (s)
b1 b 1 b −1 1 b −1 1
T (t ) = ℒ −1 + T0 − = ℒ + T0 − ℒ
as a s + a a s a s + a
b b
T (t ) = T0 − e − at
a a
Substituting back for the constants a and b and collecting terms gives
T(t):
Ah
−
mc p
t Q Ah
1 − e
−
T (t ) = T0 e + T0 + mc p
t
(8.29)
Ah
T (t)
Q̇
T0 +
Ah
T0
0 t
Example 8.17: Basic Fire Growth Model Using Unit Step Functions
As an example, in fire combustion a basic crude model of fire growth is that
it can be represented as a pulse wave similar to that given in Figure 8.4 with
the heat release rate Q (t) against time t shown in Figure 8.9.
Here in the first 2 minutes it is assumed that the fire is just getting started
and then there is a constant fire for around 20 minutes before it dies out. The
heat release above could be represented in terms of the unit step functions
with the units here being MW as follows:
Q (t ) = H (t − 2) − H (t − 22)
The next example shows how the unit step function can be used to represent
more general functions and their applications.
˙ (t)
Q
1 MW
0 2 22 t (mins)
f (t ) 0 ≤ t < t1
1
f (t ) =
f (t ) t1 ≤ t ≤ t2
2
f (t ) = 1 − H (t − t1 ) f1 (t ) + H (t − t1 ) − H (t − t2 ) f2 (t ) (8.30)
Since when 0 < t < t1, then H (t − t1) = 0 and H (t − t2) = 0. Then f(t) = f1(t).
When t1 ≤ t < t2, then H (t − t1) = 1, H (t − t2) = 0, and f(t) = f 2(t).
And when t ≥ t2, then now H (t – t1) = 1 and H (t – t2) = 1, so f(t) = 0 again
as required.
This representation of f(t) is true for all times.
Figure 8.10 can now be used to model a more realistic development of
fire growth then that given in Figure 8.9 in which the first function f1(t) is
a t-squared function and the second function f 2(t) is a constant K. Now, the
value of t1 would vary for the different types of fires from 10 minutes for a
slow fire, 5 minutes for a medium, 2.5 minutes for a fast, and 1.25 minutes
for an ultrafast fire.
If f(t) is now the heat release rate Q (t), then this can be represented by a
single function as follows using Equation 8.30:
Q (t ) = 1 − H (t − t1 ) t 2 + H (t − t1 ) − H (t − t2 ) K
f (t)
f2 (t)
K
f1 (t)
0 t1 t2 t
In other engineering fields, the use of the unit step function can offer a
method to derive solutions for problems in which the boundary conditions
depend upon time. An example of this is found in oil and gas exploration
where there is a pressure buildup following constant pressure production.
The boundary conditions for the oil well are now different for the constant
pressure and for the pressure buildup regions. Representing the boundary
conditions as a single function using the unit step functions as in Equation
8.30 allows solutions to be found subsequently using the Laplace Transform
method. This scenario often arises in drill stem testing of gas wells.
di q
L + Ri + = V (8.31)
dt C
dq
i(t ) = (8.32)
dt
V(t)
AC
i(t) L
di d 2q
this implies that = . So Equation 8.31 can be written in terms of the
dt dt 2
di
charge q(t) by replacing for and for i(t) as follows:
dt
d 2q dq q
L +R + =V (8.33)
dt 2 dt C
This differential equation has initial conditions for the current and the
0) and q(0), respectively.
charge in the circuit, which can be written as q(
This electrical circuit problem can now be solved using the Laplace
transform method to first find the charge q(t) and then subsequently the cur-
rent in the circuit i(t) using Equation 8.32.
Solution:
d 2q dq q
ℒ L 2 + R ℒ + ℒ = ℒ[V ]
dt dt
c
1
L s 2Q(s) − s q(0) − q (0) + R s Q(s) − q(0) + Q(s) = V (s)
C
2 1
L s + Rs + Q(s) = V (s) + L s q(0) + L q (0) + R q(0)
C
Note: An important part of fire studies is fire investigation. After the fire has
been put out by firefighters, fire investigators have to determine what was the
original cause of the fire. Most fire investigation studies carried out are insur-
ance related, where the cause of the fire can often have an electrical cause as
the suspicion. For example, with a voltage source V = 230 V and a device that
has been running for say 2 hours, the temperatures reached can be calculated
approximately using the appropriate energy equations, that is, VIt = mcΔT. So,
understanding well the basic concepts of electrical flow are essential to the fire
investigators’ knowledge base.
H(s)
R(s) Y(s)
Input Output
Control
system
Y ( s ) = G ( s ) E ( s ) = G ( s ) [ R( s ) − Z ( s ) ] (8.36)
Z(s)
H(s)
But Z(s) = H(s) Y(s). Substituting this in above for Z(s) gives
G (s )
Y (s ) = R( s ) (8.37)
[1 + G(s) H (s)]
This can be written as
Y ( s ) = T ( s ) R( s )
where
G (s )
T (s ) =
1 + G (s) H (s)
Thus, the transfer function for the control system is T(s) in which the
feedback transfer function H(s) can be manipulated for design purposes as
required.
The solution for Equation 8.37 can be found by the Laplace transforms
and inverse Laplace transforms method once G(s), H(s), and R(s) are known.
G (s )
y(t ) = ℒ −1 [Y (s) ] = ℒ −1 R( s )
1 + G (s) H (s)
Note: Although not studied extensively through this approach, flashover room
fires can also be understood as thermal feedback problems.
Problems
8.1 Using the formula definition, find the Laplace transform F(s) of f(t) = t2.
(Hint: Use integration by parts twice.)
228 Laplace Transforms
b. T ′(t ) − 6 T (t ) = 3; T (0) = 1
di
L + Ri = V (t )
dt
Find an expression for the current i(t) in the circuit given that i(0) = 0 for the
following cases:
V(t)
AC
i(t) L
x ′(t ) − k y(t ) = C
y′(t ) − k x (t ) = 0
The Fourier series is named after the French mathematician Jean-Baptiste Joseph
Fourier. A Fourier series is a mathematical way of representing a wavelike func-
tion (or signal) as a sum of simple sine and cosine waves. It decomposes a periodic
function or periodic signal into a sum of an infinite set of sinusoidal functions.
When these sines and cosines are expressed as complex exponentials this gives
the Fourier series in complex form. As seen in Section 9.4 the Fourier transform
is the generalization to nonperiodic functions. Since the Fourier series deals with
periodic phenomena, it is important to first understand what is meant by a func-
tion being periodic in nature.
9.1 Periodic Functions
A periodic function (or signal) f(x) is said to have a period T or be periodic with
period T if for all values of x
f (x + T ) = f (x) (9.1)
where T is a positive constant. The function then just repeats itself with period T
over the whole interval −∞ < x < ∞.
1
Note: The number of oscillation per second, that is, frequency (Hz) f = , and
2π T
angular frequency (radians per second) is w = = 2π f .
T
The function sin x is an example of a periodic function with period 2π. The
function repeats itself again after an interval of 2π (Figure 9.1).
Similar trigonometric functions that are periodic are the cosine function cos x
with period 2π, and the tangent function tan x with period π.
229
230 Fourier Series and Fourier Transforms
f (x)
Period
0 2π x
f (x)
Period
f (x)
Period
Other examples of periodic functions are the square wave and the sawtooth
function shown in Figures 9.2 and 9.3, respectively.
9.2 Fourier Series
9.2.1 Periodic Functions of Period T
Repeating functions can have different values for the period. For example, the
sine function has period 2π A more general treatment of the Fourier series is
9.2 Fourier Series 231
f (t)
Period T
given where the period can be of any value, say, T, and considering the interval
as being time t, a general periodic function f(t) can be represent as in Figure 9.4.
2π
Here, f(t + T) = f(t). Generally, w = , where w is the angular frequency (radi-
T
ans per second) and T is the period (seconds).
Using the idea that a periodic function can be represented as an infinite sum of
sinusoidal functions, it can be shown that the Fourier series to represent f(t) can
be written as follows:
f (t ) =
a0
2
+ ∑ a cos nwt + b sin nwt
n =1
n n (9.2)
Notes:
• When the period T = 2π, that is, w = 1, the terms in Equation 9.2 are
made up of cos nt and sin nt, which are periodic on the interval 2π for
any integer n.
• The coefficients an and bn measure the strength of the contribution from
each “harmonic” in the series.
The task is to see if f(t) can be written in the form given by Equation 9.2, then
what the coefficients a 0, an, and bn have to be.
Before these can be found, it is important to have some understanding of
orthogonality of functions as well as other general properties. These are covered
in the next section.
π π
∫ sin nt dt = ∫ cos nt dt = 0,
−π −π
for all integers n (9.3)
232 Fourier Series and Fourier Transforms
These last two results can easily be shown by integrating the preceding two
functions over the limits of integration but can also be seen from the graphs of
these functions. The cos nt and sin nt functions being periodic with period 2π,
then the integral over half the period cancels out the integral over the other half
of the period, as can be seen in Figure 9.1.
The functions cos nt and sin nt can now be used to help find the coefficients in
Equation 9.2 because they satisfy the following orthogonality properties:
∫ sin mt cos nt dt = 0,
−π
for all m and n (9.4)
∫ sin mt sin nt dt = 0,
−π
for m ≠ n
(9.5)
=π (half the period), for (m = n) > 0
∫ cos mt cos nt dt = 0 ,
−π
for m ≠ n
The orthogonality properties given in Equations 9.4, 9.5, and 9.6 can all be
proved by either considering the graphs of the product functions or by expressing
the product functions in terms of the sums of individual sine and cosine functions
and integrating out (see trigonometric functions and integration sections). These
results will all be very useful when calculating the Fourier coefficients in the next
section.
Another important result for the Fourier series in complex form is that in the
exponential notation the orthogonality conditions where m and n are integers
become
∫e
−π
jnt − jmt
e dt = 0; m≠n
(9.7)
= 2π (period); m=n
Again, the proof of this is evident from direct integration and putting in the limits.
9.2.3 Fourier Coefficients
Starting with Equation 9.2 for the Fourier series:
f (t ) =
a0
2
+ ∑ a cos nwt + b sin nwt
n =1
n n
9.2 Fourier Series 233
The coefficients a0, an, and bn can be found by multiplying f(t) by 1, cos mwt, and
sin mwt, and integrating over a period, respectively, to give the following results:
2
a0 =
T ∫ f (t) dt
T
(9.8)
2
an =
T ∫ f (t) cos nwt dt
T
(9.9)
2
bn =
T ∫ f (t) sin nwt dt
T
(9.10)
Proofs:
For the coefficient a 0, starting with f(t)
f (t ) =
a0
2
+ ∑ a cos nwt + b sin nwt
n =1
n n
∫ f (t ) dt =
∫
a0
2
dt + ∑ a ∫ cos nwt dt + b ∫ sin nwt dt
n n
T T n =1 T T
a0
∫ 2 dt = ∫ f (t) dt
T T
(9.11)
since
a0 T
2
=
∫ f (t) dt
T
and so a 0 is
2
a0 =
T ∫ f (t) dt
T
as required.
234 Fourier Series and Fourier Transforms
Similarly, proofs for an and bn can be shown using the orthogonality properties.
Note: For a general period T, it is better to first sketch the function f(t), then
T T
choose the appropriate periodic interval, that is, from 0 to T or − to .
2 2
Example 9.1
Determine the Fourier series for the periodic function defined by:
Solution: To see what the function given by Equation 9.12 looks like, it is
first sketched as shown in Figure 9.5.
Starting with the general Fourier series for f(t),
∞
a
f (t ) = 0 +
2 ∑ a cos nwt + b sin nwt
n =1
n n (9.13)
2π
Since the period of f(t) is equal to T = 2, this implies that w = = π.
Calculating the coefficient for a 0 using Equation 9.8 gives T
1 1
2 2
a0 =
T ∫T
f (t ) dt =
2 ∫ f (t) dt = ∫ f (t) dt
−1 −1
f (t)
–3 –2 –1 0 1 2 3 t
Now this integral has to be split into two regions since the function f(t)
is different in the two regions −1 < t < 0 and 0 < t < 1, as can be seen from
Figure 9.5.
1 0 1
a0 =
∫
−1
f (t ) dt =
∫
−1
f (t ) dt +
∫ f (t) dt
0
(9.14)
Putting in the function values for the function in the different regions
into Equation 9.14 gives
0 1
a0 =
∫
−1
2(1 + t ) dt +
∫ 0 dt
0
0
a0 =
∫ 2(1 + t) dt = 1
−1
2
an =
T ∫ f (t) cos nwt dt
T
Using w = π gives
1 0
2
an =
2 ∫ f (t) cos nπ t dt = ∫ f (t) cos nπ t dt
−1 −1
an =
∫ 2(1 + t) cos nπ t dt
−1
(9.15)
2
an = [1 − cos nπ ] (9.16)
n 2π 2
236 Fourier Series and Fourier Transforms
where n = 1, 2, 3, ….
This coefficient an has different values according to whether n is odd or
even as follows:
4 (9.17)
If n = odd (1, 3, 5, ), an = because cos nπ = −1
n π22
These are the values of an that will be used in Equation 9.13 for f(t).
Similarly, calculating for the coefficient for bn using Equation 9.10 gives
2
bn =
T ∫ f (t) sin nwt dt
T
with w = π, gives
1 0
2
bn =
2 ∫ f (t)sin nπ t dt = ∫ f (t)sin nπ t dt
−1 −1
bn =
∫ 2(1 + t)sin nπ t dt
−1
(9.19)
Now this integral is again a product of two functions and has to be done
dv
by using integration by parts, that is, let u = 2(1 + t) and then = sin nπ t
gives dt
2
bn = − (9.20)
nπ
where n = 1, 2, 3, ….
Putting together all the coefficients found gives
a0 = 1
4
an = , n = odd
n 2π 2
an = 0, n = even (9.21)
2
bn = −
nπ
9.2 Fourier Series 237
Having all the coefficients, the Fourier series for f(t) in Equation 9.2 can
be written as
1 4 cos 3π t cos 5π t
f (t ) = + cos π t + + +…
2 π2 9 25
(9.22)
2 sin 2π t sin 3π t
− sin π t + + +…
π 2 3
∞ ∞
4
f (t ) =
1
+ ∑ cos nπ t +
2 n= odd n 2π 2 ∑ − n2π sin nπ t
n=1
(9.23)
The above sum continues to an infinite number of terms. It can be seen how
it converges to the original function by plotting a truncated sum of a finite num-
ber of terms. If the sum containing n-trigonometric terms is defined as fn(t), then
1
f0 (t ) = (which is just the average value of the function over the period)
2
1 4 2
f1 (t ) = + 2 cos π t − sin π t
2 π π
1 4 2 2
f2 (t ) = + cos π t − sin π t − sin 2π t , etc.
2 π2 π 2π
The graphs of these functions can now be plotted to see how the series
converges to the original function, as shown in Figure 9.6.
As more and more terms in the series are taken, the resulting function
approximates the original signal more closely. This can be seen in Figure
9.6 with n = 40. The series is starting to approximate the original function
given by Equation 9.12 and Figure 9.5 reasonably well.
Note: The oscillations seen in Figure 9.6b and c do become smaller and smaller
as n gets larger and larger but do not disappear altogether since a discontinuous
function is being represented by smooth sinusoidal functions. This is known as
Gibbs phenomenon after J.W. Gibbs.
1.2 1.8 2
1.0 1.6
1.4
0.8 1.5
1.2
1
0.6 1
0.8
0.4 0.6
0.4 0.5
0.2
0.2
–3 –2 1 2 3 –3 –2 –1 0 1 2 3
–0.2 t –3 –2 –1 0 1 2 3
t t
(a) (b) (c)
Figure 9.6 Graphs of f n(t) for values of n: (a) n = 1, (b) n = 6, and (c) n = 40.
238 Fourier Series and Fourier Transforms
f (t ) =
a0
2
+ ∑ a cos nwt + b sin nwt
n =1
n n (9.24)
can be written in a more compact form known as the complex form, which simpli-
fies the calculations and is more useful especially when the Fourier transform is
considered later.
Starting with Euler’s formula:
then
By adding and subtracting Equations 9.25 and 9.26 gives the following formu-
lae for the cosine and sine function terms of exponentials:
e jθ + e − jθ
cos θ = (9.27)
2
e jθ − e − jθ
sin θ = (9.28)
2j
e jnwt + e − jnwt
cos nwt = (9.29)
2
e jnwt − e − jnwt
sin nwt = (9.30)
2j
In Equation 9.24 for the trigonometric Fourier series, the term inside the ∑,
which is an cos nwt + bn sin nwt, now becomes
1 ( jnwt
+ e − jnwt ) +
1 ( jnwt
an e bn e − e − jnwt )
2 2j
(9.31)
1
2
( an − jbn ) e jnwt + 12 ( an + jbn ) e− jnwt
9.3 Complex Form of the Fourier Series 239
Letting
1
cn =
2
( an − jbn ) and kn = 12 ( an + jbn ) (9.32)
then
f (t ) = c0 + ∑c e
n =1
n
jnwt
+ kn e − jnwt (9.34)
a0
where now c0 is given as c0 = .
2
Also, using the definitions of cn and kn given in Equation 9.32, it can be shown
further that kn = c−n, where cn is given by
1
cn =
T ∫ f (t) e
T
− jnwt
dt (9.35)
f (t ) = c0 + ∑c e
n =1
n
jnwt
+ c− n e − jnwt (9.36)
If the summation for n goes from −∞ to ∞, this then gives the most compact
form for the Fourier series as
f (t ) = ∑c e
n =−∞
n
jnwt
(9.37)
where,
1
cn =
T ∫ f (t) e
T
− jnwt
dt (9.38)
Again, this result for cn follows directly from the orthogonality condition for
exponential notation, that is, from Equation 9.7.
Note: The real coefficients an and bn can be obtained from Equation 9.32 using
cn = 1
2
( an − jbn ) and c− n = 12 ( an + jbn ) .
240 Fourier Series and Fourier Transforms
These allow for converting back to the real form of the Fourier series if required.
The next example shows how to find the Fourier series of a periodic function
using the complex form.
Example 9.2
Determine the complex Fourier series for the following periodic function:
f (t ) = ∑c e
n =−∞
n
jnwt
(9.41)
1
cn =
T ∫ f (t) e
T
− jnwt
dt (9.42)
2π
Since f(t) = f(t + 2π) gives T = 2π and so w = =1.
Calculating the coefficients cn using T
1
cn =
T ∫ f (t) e
T
− jnwt
dt (9.43)
f (t)
–3π –2π –π 0 π 2π 3π t
π
1
=
2π ∫t e
−π
− jnt
dt (9.44)
dv
u=t and = e − jnt
dt
gives
e − jnt
du = dt and v=
− jn
− jnt π π
− jnt
1 t e − e dt
cn =
2π − jn − π
−π
− jn
∫
− jnt
π
cn =
1
(
−π π e − jnπ − (−π ) e jnπ − e )
2
2π jn (− jn) − π
1 −π ( − jnπ
+ e jnπ ) + 2 ( e − jnπ − e jnπ )
1
cn = e
2π jn n
1 −π
cn = 2(−1)n
2π jn
This gives
j(−1)n
cn =
n
j ( −1) jnwt
∞ n
f (t ) = ∑
n =−∞
n
e (9.45)
Converting back to the real form of the Fourier series can be done using
Equation 9.39, that is, an = cn + c−n and jbn = c−n − cn if required.
242 Fourier Series and Fourier Transforms
9.4 Fourier Transforms
So far, the Fourier series has been representing periodic functions by a com-
bination of infinite sinusoidal functions. What happens when the function is
not periodic in nature? Can it still be made up of a combination of simpler
functions?
The idea is to transfer from periodic phenomena to nonperiodic phenomena.
This can be achieved by viewing a nonperiodic function as a limiting case of a
periodic function as the period tends to infinity.
9.4.1 Nonperiodic Functions
As stated earlier, Fourier series are applicable to periodic functions only, but non-
periodic functions can also be decomposed into Fourier components. This pro-
cess is called a Fourier transform of a function or signal.
First consider a function that is of finite extent but much less than its periodic-
ity, T, as shown in Figure 9.8. If the period T becomes very large, that is, it tends to
infinity, then the above function shown in Figure 9.8 becomes an isolated aperiodic
function as required. This limiting process is used to develop a heuristic approach
to finding the equations for the Fourier transform from the Fourier series.
f (t ) = ∑c e
n =−∞
n
jnwt
(9.46)
1
cn =
T ∫ f (t) e
T
− jnwt
dt (9.47)
It would be nice to just let T →∞ for cn in Equation 9.47, but this will not work
since it can be shown that Cn → 0 in this case.
f (t)
t
–T –T/2 0 T/2 –T
What is considered first is to multiply the equation for cn by the period T and
look at the integral given by
∫ f (t) e
T
− jnwt
dt (9.48)
T T
The period can be any period, so consider going from − to and replacing
2π 2 2
for w = into Equation 9.48 giving
T
T
2 n
n
∫
−2π j t
T
F = f (t ) e dt (9.49)
T
T
−
2
Then Equation 9.46 for the Fourier series can now be written as
∞ n
n 2π j t 1
f (t ) = ∑
n =−∞
F e T
T T
(9.50)
n
Now as T →∞, the “discrete variable” gets closer together and is replaced
T
by the continuous variable s, where −∞ < s < ∞.
Therefore, Equation 9.49 can now be called the Fourier transform and can be
written as
∞
F (s ) =
∫ f (t ) e
−∞
−2π jst
dt (9.51)
And the Fourier series Equation 9.49 is given by the summation changing to
1
an integral, and the gets smaller and smaller ≈ ds in integration for the limiting
T
process giving the following:
f (t ) =
∫ F (s)e
−∞
2π jst
ds (9.52)
Therefore, the equations for the Fourier transform for a nonperiodic function are
∞
F (s ) =
∫ f (t) e
−∞
−2π jst
dt (9.53)
f (t ) =
∫ F (s ) e
−∞
2π jst
ds (9.54)
Equation 9.53 is called the Fourier transform of f(t) and Equation 9.54 is called
the inverse Fourier transform of F(s).
244 Fourier Series and Fourier Transforms
Note: The definitions and notations for the Fourier transform and its inverse
are not rigidly fixed; they can vary by factors of 2π or 2π in their equations.
It is useful to see how Equation 9.53 can be used to find the Fourier transform
of functions, as shown in the next example.
Example 9.3
Find the Fourier transform of the following rectangular function
f (t ) = ∏(t ) given by
1 1
f (t ) = 1 − <t<
2 2
(9.55)
1
=0 t ≥
2
F (s ) =
∫ f (t) e
−∞
−2π jst
dt
1
1
2
e −2π jst 2 e − π js − eπ js
1
∫
= e −2π jst dt = =
−2π js − 1 −2π js
− 2
2
e − π js − eπ js 1 eπ js − e − π js
= =
−2π js πs 2j
sin π s
F (s) =
πs
f (t)
–1/2 0 1/2 t
F(s)
–2π –π 0 π 2π s
Example 9.4
See the general waveform f(t) in Figure 9.11. To find the Fourier transform
of f(t) is to ask what combination of sinusoids added together will give f(t).
Suppose that f(t) is made up of two functions f1(t) and f 2(t), then the Fourier
f (t)
–T/2 0 T/2 t
Transform of f(t) has been found. If the two functions are shown as in
Figure 9.12, then a diagram can be constructed that displays the amplitude
and frequency of each of the sinusoids f1(t) and f 2(t).
From Figure 9.11, it can be seen that
1
f1 (t ): amplitude = 1; frequency =
T
1 3
f2 (t ): amplitude = ; frequency =
2 T
Putting this information onto a single diagram gives Figure 9.13.
In terms of the delta (or impulse) function this can be written as
1 3 1 3 1 1 1 1
F (s) = δ s− + δ s+ − δ s− − δ s+
4 T 4 T 2 T 2 T
A summary of this is to say that every signal has a spectrum and the spec-
trum then determines the signal.
f1(t) f2(t)
1 +
1/2
F (s)
1/2
1/4
–1
Figure 9.13 Amplitude and frequency components of functions f1(t) and f 2(t).
9.4 Fourier Transforms 247
[ f (t )] = F (s)
9.4.4.1 Linearity Property
If f(t) and g(t) are functions of time and a and b are constants, then
a f (t ) + b g(t ) = a F f (t ) + b F g(t )
= a F (s ) + b G (s )
There are other important general properties of the Fourier transform and
some of these are given in Table 9.1. The proof of the properties can be shown
using the definition formula of the Fourier transform given by Equation 9.53. The
next example shows the proof for property 3, that is, multiplying a function in the
time domain by an exponential produces a shift in the frequency domain.
Example 9.5
Show that the [e2πjat f(t)] = F(s − a).
F (s ) =
∫ f (t) e
−∞
−2π jst
dt (9.56)
e 2π jat f (t ) =
∫e
−∞
2π jat
f (t )e −2π jst dt (9.57)
248 Fourier Series and Fourier Transforms
=
∫ f (t) e
−∞
−2π j ( s − a ) t
dt = F (s − a) (9.58)
f (t ) + g(t ) = f (t ) + g(t )
This is just modifying [f(t)] by adding the spectrum of [g(t)] to it. Therefore,
what about multiplying the Fourier transform of two functions does this have a
similar correspondence in the time domain of just multiplying the functions? The
answer to this is not as simple as just simply multiplication of functions in the
time domain but of a convolution of two functions in the time domain instead.
To show this process, start with the multiplication of the Fourier transform of
two functions:
g(t ) × f (t ) (9.59)
Replacing for the definitions of the Fourier transforms for both g(t) and f(t)
gives
∞ ∞
∫
= g(t ) e
−∞
−2π jst
−∞
∫
dt f ( x ) e −2π jsx dx
(9.60)
∞ ∞
=
∫ ∫e
−∞ −∞
−2π jst −2π jsx
e g(t ) f ( x ) dt dx (9.61)
∞ ∞
=
∫ ∫e
−∞ −∞
−2π js ( t + x )
g(t ) f ( x ) dt dx (9.62)
9.4 Fourier Transforms 249
∞ ∞
=
∫ ∫
e −2π js (t + x ) g(t ) dt f ( x ) dx
−∞ −∞
(9.63)
∞ ∞
∫ ∫
= e −2π jsu g(u − x ) du f ( x ) dx
−∞ −∞
(9.64)
∞ ∞
=
∫ ∫
g(u − x ) f ( x ) dx e −2π jsu du
−∞ −∞
(9.65)
h(u) =
∫ g(u − x) f (x) dx
−∞
(9.66)
[h(u)] =
∫ h(u) e
−∞
−2π jsu
du (9.67)
(g * f ) x =
∫ g(x − y) f ( y) dy
−∞
(9.69)
Then finally, it can be stated that the Fourier transform of the convolution of
two functions is equal to the product of the Fourier transform of the two functions
as shown next:
g * f = g(t ) × f (t ) (9.70)
This is again an important property for signal processing purposes as seen in the
applications section.
250 Fourier Series and Fourier Transforms
9.5 Applications
Example 9.6: Designing of Fourier Transform Infrared (FTIR)
Smoke Detectors
Fire detector systems should have the ability to discriminate between real
fire sources and nonfire sources. Smoke detectors that can respond quickly
may suffer from the inability to discriminate between a real fire smoke
and smoke from other sources. In high-value installations where there is
expensive and sensitive equipment, it is clear that reliable fire detection
systems are needed. The detection systems are generally used to activate
fixed fire suppression systems like sprinklers, so any false alarms are an
undesirable outcome causing valuable time loss and having potential cost
implications.
Fourier transform infrared (FTIR) smoke detectors make use of Fourier
Transform techniques to analyze combustion products and so aid new more
reliable detection systems to be built. Using FTIR measurements of con-
centrations of gases (e.g., CO2, CO, H2O, and CH4) given off during differ-
ent modes of combustion, the detection system can classify input data as
a flaming fire, smoldering fire, nuisance, or other environmental sources.
Therefore, FTIR spectroscopy can give multiple gas concentrations that
enable an advanced fire detection system to be built.
c
λ=
f
φ( t ) = x ( t ) 1 e 2π jω t + e −2π jω t
2
9.5 Applications 251
If the Fourier transform of x(t) is X(s) and the Fourier Transform of φ(t)
is Φ(s) then making use of property 3 in Table 9.1 gives
1
Φ (s ) = X (s − ω c ) + X (s + ω c )
2
X(s)
Φ(s)
1/2
–wc wc s
φ (t)
0 t (months)
Φ(s)
0 10 20 30 40 50
Frequency
Figure 9.16 Graph showing the frequency spectrum of the signal Φ(s).
–ν 0 ν
Frequency
φ' (t)
0 t (months)
The result φ′(t) is a much smoother function with the jaggedness taken
out, as shown in Figure 9.18.
Examples can also be found where the use of other types of filtering,
such as, high-pass filters and band-pass filters that allow signals to be
manipulated to produce the required outputs.
Problems
9.1 Determine the trigonometric Fourier series for the following functions:
a. f (t ) = 0 −2 < t < 0
=t 0<t<2
f (t ) = f (t + 4)
9.3 Prove the orthogonality property for the following exponential func-
tions, where m and n are integers.
∫e
−π
jnt − jmt
e dt = 0 m≠n
= 2π m=n
9.4 Given the triangular function Λ(t) is shown in Figure 9.19 and defined
below as
Λ(t ) = 1 − t t ≤1
=0 t >1
Λ(t)
–1 0 1 t
2
9.5 Given the Gaussian function f (t ) = e − π t , show that its Fourier trans-
2
form is given by F (s) = e − π s . (Harder problem)
(Hint: Start with definition, consider F′(s) and then integration by parts.)
10 Multivariable
Calculus
10.1 Partial Derivatives
10.1.1 Introduction and Definition
A function of two or more variables such as
f ( x , y) = xy − ye x
has two inputs x and y and it produces one output. A many-input function could
have many outputs and is usually termed as a vector function.
Recall the definition from 1-dimensional differentiation (or derivative) as seen
in Chapter 6,
df ( x ) f ( x + h) − f ( x )
= f ′( x ) = lim (10.1)
dx h→0 h
255
256 Multivariable Calculus
y
y = f (x)
Tangent line
Slope f ʹ(a)
a x
Figure 10.1 The derivative as the slope of the tangent line at some point.
If there is a change in the input by an amount h, the output changes by the rate
of change output f′(a) multiplied by h. So f′(a) represents the instantaneous rate
of change of f(a).
Now what is the situation when there is two or more variables? Here there is
still a definition of derivatives, but since there is more than one variable these are
called partial derivatives. Considering a function with two variables, the partial
derivatives can be defined in a similar manner to Equation 10.1.
∂f f (a + h, b) − f (a, b)
Partial with respect to x: (a, b) = f x lim
∂x h→0 h
∂f f (a, b + h) − f (a, b)
Partial with respect to y: (a, b) = f y lim
∂y h→ 0 h
So, with partial derivatives the derivatives are taken with respect to one vari-
able while keeping the other variables fixed.
∂f
Note: The shorthand notation for partial derivatives is (a, b) = f x .
∂x
In the same way as with 1-D derivatives, the interpretation of the partial deriv-
atives can be seen as an instantaneous rate of change as follows.
10.1 Partial Derivatives 257
∂f
f (a + h, b) ≈ f (a, b) + h (a, b)
∂x
∂f
f (a, b + h) ≈ f (a, b) + h (a, b)
∂y
∂f ∂f
f (a + h, b + k ) ≈ f (a, b) + h (a, b) + k (a, b) (10.3)
∂x ∂y
g( x ) = f ( x , b) with y = b
then
∂f
g′(a) = (a, b)
∂x
at the point x = a.
z = f (x, y)
(a, b)
Fix x = a
Figure 10.2 Shows the slope of the tangent line keeping one variable fixed.
258 Multivariable Calculus
This is generally the best way to understand partial derivatives and is the way
that is used to compute them. The next examples illustrate this process.
Example 10.1
Given f(x, y) = x + xy2, compute the following.
∂f
1. (1, 3)
∂x
∂f
2. (2, 4)
∂y
Solution:
∂f
1. For (1, 3), here x is varying and so consider y as fixed or constant.
∂x
f(x, y) = x + xy2 differentiating with respect to x keeping y fixed
∂f
gives = 1 + y 2.
∂x
Now substituting in the values of the point (1,3) gives
∂f
(1, 3) = 1 + 32 = 10
∂x
∂f
2. For (2, 4), here y is varying and consider x as fixed or constant.
∂y
∂f ∂f
= 2 xy , so (2, 4) = 2(2)(4) = 16 .
∂y ∂y
Example 10.2
Given f(x, y) = x2y + xey, compute (1) fx(1, 0) and (2) f y(1, 1).
Solution:
∂f
1. f x = = 2 xy + e y
∂x
f x (1, 0) = 2(1)(0) + e 0 = 1
∂f
2. f y = = x 2 + xe y
∂y
10.1.2 Higher Derivatives
As with the one variable case higher derivatives can be calculated. The only dif-
ference now is that with the multiple variable case, there are many higher deriva-
tives to consider. With a single variable case, there is only one second derivative
but with a two-variable problem. It turns out that there are four second derivatives
that can be calculated. Some of the notation used is given next.
The following notation is used:
∂2 f ∂ ∂f
= = f xx
∂x 2 ∂x ∂x
∂2 f ∂ ∂f
= = f yy
∂y 2 ∂y ∂y
∂2 f ∂ ∂f
= = f xy
∂y∂x ∂y ∂x
∂2 f ∂ ∂f
= = f yx
∂x ∂y ∂x ∂y
Example 10.3
Given f(x,y) = xey + xy3, find the following second-order derivatives:
Solution:
1. fx = ey + y3, fxx = 0
2. fx = ey + y3, fxy = ey + 3y2
3. fy = xey + 3xy2, f yx = ey + 3y2
4. fy = xey + 3xy2, f yy = xey + 6xy
It can be seen that fxy = f yx. Generally, this property is observed and the
following theorem known as Clairaut’s theorem expresses this.
10.1.2.1 Clairaut’s Theorem
Most of the time fxy(a, b) = f yx (a, b). The order of the differentiation does not mat-
ter. With more variables, that is, x, y, z, the order does not matter, for example,
fxyzz = fzxzy. However, there are conditions on f for this to be true.
fx = 2 x + 3 y
f y = 3x + e y
260 Multivariable Calculus
Now the function f(x, y) has to fit both Equation 10.4 and Equation 10.5.
Therefore,
f ( x , y) = x 2 + 3 xy + e y
will satisfy both the required conditions. This will be an important idea for
working with conservative vector fields as seen later in Chapter 11.
Example 10.5
Find f(x, y) such that
fx = 4 x − 3 y
fy = x + y
Solution:
f x = 4 x − 3 y implies f = 2 x 2 − 3 xy + g( y)
y2
fy = x + y implies f = xy + + h( x )
2
Now in this case there cannot be found any function f(x, y) that fits both
profiles. Therefore, it is not possible to find any f(x, y) that has these required
fx and f y functions simultaneously.
10.1.3 Chain Rule
10.1.3.1 Chain Rule with One Variable
The chain rule in one variable is a process of calculating derivatives in situations
where there is a function of a function involved. This can be stated in different
ways, including
d
f g( x ) = f ′ g( x ) g′( x ) (10.6)
dx
10.1 Partial Derivatives 261
f = f ( y) and g = g( x )
df df dy
= × (10.7)
dx dy dx
Sometimes it is easier to see how the variables are related to each other by
drawing a dependency chart as shown in Figure 10.3. Equations 10.6 and 10.7 are
the same process.
Using Equation 10.7 as the way of thinking about how variables depend on
each other is more useful as it leads to a natural extension in higher dimensions
as shown in the next sections.
g=y
x = x (u, v)
y = y(u, v)
Then what is the change in the function due to a change in the variable u, that
∂f
is, ?
∂u
First, drawing a dependency chart helps to show the situation of how the vari-
ables are related to each other, as in Figure 10.4.
So,
∂f ∂f ∂x ∂f ∂y
= × + ×
∂u ∂x ∂u ∂y ∂u
This gives the net chain in f due to the change u through the variables x
and y.
262 Multivariable Calculus
x y
u v
Example 10.6
∂f ∂f ∂z ∂x ∂f ∂w ∂y ∂f ∂z ∂y
= + +
∂u ∂z ∂x ∂u ∂w ∂y ∂u ∂z ∂y ∂u
∂f ∂f ∂z ∂y ∂f ∂w ∂y
= +
∂v ∂z ∂y ∂v ∂w ∂y ∂v
Consider the chain rule by using the diagram to find the relationships
between the different variables.
z w
x y
u v
Example 10.7
∂f
Given that f = f (u,v,t) where u = u(t) and v = v(t), find.
∂t
Solution: First draw the dependency chart as in Figure 10.6.
df ∂f ∂u ∂f ∂v ∂f
= + +
dt ∂u ∂t ∂v ∂t ∂t
u v
a
x = x0 + t (10.8)
a + b2
2
b
y = y0 + t (10.9)
a + b2
2
These makes us travel along 〈a, b〉 with unit speed so unit time is now equal to
unit distance in the direction 〈a, b〉. A dependency chart for this situation is shown
in Figure 10.8.
Now the change in f due to a change in t is given by
∂f ∂f dx ∂f dy
= +
∂t ∂x dt ∂y dt
264 Multivariable Calculus
<a, b>
(x0, y0)
x y
Using Equations 10.8 and 10.9 and differentiating with respect to t gives
∂f ∂f a ∂f b
= + (10.10)
∂t ∂x a +b
2 2 ∂y a + b2
2
∂f ∂f 〈 a, b 〉
D‹ a ,b › f = 〈 , 〉⋅ (10.11)
∂x ∂y a2 + b2
∂f ∂f
Using some further notation as 〈 , 〉 = ∇f (called the gradient of f). Also,
∂x ∂y
∇ is sometimes called “del” or “nabla.” So, Equation 10.11 can be written in sim-
pler notation as follows.
10.1 Partial Derivatives 265
u
Du f = ∇f ⋅ = ∇f . uˆ (10.12)
u
Now it can be checked to see what the directional derivatives would be in the x
and y directions from the formula given in Equation 10.12 as follows.
In the x-direction, the unit vector is uˆ = 〈1, 0 〉. This then gives Du f as
D〈 0 , 1〉 f = ∇f .〈0, 1〉 = 〈 f x , f y 〉.〈0, 1〉 = f y
Example 10.8
Given f(x,y) = xy2 – 10x:
1. Compute ∇f.
2. What is the directional derivative of f in the direction 〈2,5〉 at the
point (1, 1).
Solution:
〈2, 5〉
2. D〈 2,5〉 f (1,1) = ∇f (1,1).
29
〈2, 5〉 8
= 〈−9, 2〉 ⋅ =−
29 29
Example 10.9
Given f(x,y) = x – xy2, find:
1. D〈2,−1〉 f(1,0)
2. In what direction û is Du f the biggest?
Solution:
Therefore,
〈2, −1〉 2
D〈 2,−1〉 f (1, 0) = 〈1, 0 〉. =
5 5
2.
To make this the biggest, cos θ needs to be the largest, which implies
θ = 0. Therefore, angle between ∇f and û should be in the same direction. So,
û should be in the same direction as ∇f but the unit vector version of it, that is,
∇f
û =
∇f
f ( x ) ≈ f (a) + ( x − a) f ′(a)
And if y ≈ f (a) + ( x − a) f ′(a) this is just the tangent line, similar to the argu-
ments for the tangent plane in higher-order problems.
What do you consider if it is a maximum? At a maximum, f ′(a) = 0 . So, the
second-order approximation is shown in Figure 10.10.
1
f ( x ) ≈ f (a) + ( x − a) f ′(a) + ( x − a)2 f ′′(a)
2
y = f (x)
a x
y = f (x)
a x
This is the Taylor series expansion of a function at a point. This is the parabola
of best fit at x = a. Now, if f′(a) = 0 at a maximum, then f(x) becomes
1
f ( x ) ≈ f (a) + ( x − a)2 f ′′(a)
2
So, this parabola is shaped if f ″(a) > 0 (i.e., a minimum) and this parab-
ola is shaped if f″(a) < 0 (i.e., a maximum).
f f
xx xy
∇2 f =
f yx f yy
Note: The determinant of the above matrix is also sometimes referred to as the
Hessian.
f f
xx xy
D = det
f yx f yy
f f
xx xy
2. Let D = det = f xx (a, b) f yy (a, b) − f xy2 = Discriminant
f yx f yy
Example 10.10
Find and identify the critical points of the following function f(x, y):
f ( x , y) = x 3 − 12 xy + 8 y3
Solution:
0 = f x = 3 x 2 − 12 y gives x 2 = 4 y
0 = f y = −12 x + 24 y 2 gives x = 2 y 2
Using x = 2y2 into x2 = 4y gives (2y2)2 = 4y. Solving for y gives 4y(y3 – 1) = 0.
This gives y = 0 or y = 1 as the solutions to this equation which, then gives
the critical points as (0,0) and (2,1).
To see if the critical points are a maximum or minimum, find the Hessian D:
f f
xx xy
D = det = f xx (a, b) f yy (a, b) − f xy2
f yx f yy
f xx = 6 x f xx (0, 0) = 0 f xx (2,1) = 12
f yy = 48 y f yy (0, 0) = 0 f yy (2,1) = 48
So, D(0,0) = (0)(0) – 144 gives D = – 144 < 0, which implies a saddle
point.
D(2,1) = (12)(48) – 144 gives D = 432 > 0, but fxx(2,1) > 0, so it is a
minimum.
10.2 Higher-Order Integration 269
Example 10.11
Find three numbers that sum to 100 and have the largest product.
0 = f x = y(100 − x − y) + x (− y) = y(100 − 2 x − y)
Since the product of two things equals zero implies y = 0, the product = 0,
so ignore this.
Therefore, 100 – 2x – y = 0
and x = 0, then again the product = 0, so ignore this.
Therefore, 100 – x –2y = 0.
Solving these two equations simultaneously gives
2 x + y = 100
x + 2 y = 100
100
−3 y = −100 y=
3
10.2 Higher-Order Integration
10.2.1 Double Integrals and Fubini’s Theorem
Instead of integrating over an interval, integration is carried out as shown in
Figure 10.11 over a 2-D region.
First, chop the region into small pieces and pick a point (x*, y*). Then,
∫∫ f (x, y) dA lim ∑
D
∆A→0
all pieces
f ( x*, y*)∆A
(x*, y*)
∆A
x
a b
∫∫ f (x, y) dA lim ∑
D
∆A→0
all pieces
f ( x*, y*)∆A
Split the rectangular region into horizontal and vertical strips as shown in
Figure 10.13.
lim lim
∆x →0 ∆y→0 ∑ ∑ f (x*, y*)∆y∆x
i j
= lim
∆x →0 ∑ ∑ lim f (x*, y*)∆y ∆x
∆y→0
i j
d
= lim
∆x →0 ∑ ∫ f (x*, y) dy∆x
i c
d
=
∫ lim ∑ f (x*, y)dy∆x
c
∆x →0
i
b d
∫ ∫ f (x, y) dy dx
a c
10.2 Higher-Order Integration 271
d
(x*, y*)
∆y
c
∆x
x
a b
Example 10.12
3 1
Evaluate ∫ ∫ (2xy − 4 y) dy dx
0 0
Example 10.13
π
3 2
Evaluate
∫ ∫ x cos y dy dx.
0 0
2
Solution:
3 π 3 3
x3
∫
0
x 2 sin y 2 dx =
0 ∫
0
x 2 dx = = 9
3 0
Example 10.14
1 1
Evaluate
∫ ∫ xy
0 0
x 2 + y 2 dy dx (more tricky problem).
272 Multivariable Calculus
1 1
( )
1 3
1 3 2 5 5
x (x + y ) 2 x ( x + 1) − x dx = ( x + 1) − x = 1 2 2 − 2
2 2 2 4 5
∫ ∫
2 2
3 dx =
3 3
15 15 0 15
0
0 0
b d d b
∫ ∫ f (x, y) dy dx ≡ ∫ ∫ f (x, y) dx dy
a c c a
(10.13)
This is just saying that it does not matter if you sum x first or y first.
Vol ≈ lim
∆A→0 ∑
all pieces
f ( x*, y*)∆A
Volume =
∫∫ f (x, y) dA
D
(10.14)
z
z = f (x, y)
D y
x ∆A
Example 10.15
Find the volume of the solid in the first octant (see Figure 10.15) bounded
by z = 9 – x2 and y = 4.
Solution: What surface does z = 9 – x2 look like? Since y is not present it can
take on any value. This region is shown in Figure 10.15.
Volume =
∫∫ f (x, y) dA = ∫∫ ( 9 − x ) dA
D D
2
3 4
=
∫ ∫ (9 − x ) dy dx
0 0
2
z = 9 – x2
9
4
D
3
y
x = r cos θ
y = r sin θ
274 Multivariable Calculus
P
r
y
θ
x
x
r = x 2 + y2
y
Note: Sometimes θ = tan −1 , but in general must be careful if the point is in a
x
different quadrant, hence the need to use trigonometry to find the angle θ accord-
ing to the problem being solved.
r = r2 (θ)
β r = r1 (θ)
α
x
∫∫ f (x, y) dA lim ∑
D
∆A→0
all pieces
f ( x*, y*)∆A
= lim lim
∆r →0 ∆θ →0 ∑ ∑ f (r* cosθ*, r* sinθ* ) r∆r∆θ
i j
i j i j
The r in r∆r∆θ is needed here for the change of variable (see later section defin-
ing Jacobian).
β r2
Area =
∫ ∫ f (r cosθ , r sinθ )r dr dθ
α r1
(10.15)
Example 10.16
Consider the disc x2 + y2 ≤ 4 has charge density σ(x,y) = 3x + x2 + y2.
Compute the total charge on the disc. The diagram of this region is shown
in Figure 10.18.
A=
∫∫ (3x + x
D
2
)
+ y 2 dA
0 x
–2 2
–2
2π 2
∫ ∫ (3r cosθ + r )r dr dθ
0 0
2
2π 2
3 1 4
=
∫
0
r cos θ + 4 r dθ
0
2π
=
∫ [8 cosθ + 4] dθ = 8π
0
Since the integral of both the cos θ and sin θ over a period are both equal
to zero.
Example 10.17
Solution:
Volume =
∫∫ Upper surface − ∫∫ Lower surface
z
D
2
2
Volume =
∫∫ (D
4 − x 2 − y2 − )
x 2 + y 2 dA (10.16)
The region D is a circle, which is the intersection of the sphere with the
cone as shown in Figure 10.20.
x 2 + y 2 + z 2 = 4 with z = x 2 + y2
This gives
( 2)
2
x 2 + y 2 = 2 or x 2 + y2 =
2π 2
∫ ∫(
0 0
)
4 − r 2 − r r dr dθ
2π 2
∫ ∫ (r
0 0
)
4 − r 2 − r 2 dr dθ
Which is just a standard integral with respect to r first then with respect
to θ giving the volume as
2π 2 2π 5
1 3
1 3 8 − 2
∫ ∫
2
− (4 − r ) 2 − r dθ =
2 dθ
3 3 0 3
0 0
0 x
√2
Figure 10.20 The circle that is the intersection of the sphere and cone.
278 Multivariable Calculus
Volume =
2π
3
(
8 − 22
5
)
10.2.3 General Regions
Previously, calculations of double integrals over a rectangular region were consid-
ered as shown in Figure 10.21.
The area of the region D is given by
b d d b
∫∫
D
f ( x , y) dA =
∫∫
a c
f ( x , y) dy dx =
∫ ∫ f (x, y) dx dy
c a
Suppose the region D has the general form as shown in Figure 10.22.
The area of the region D can be found as
b g2 ( x )
∫∫ f (x, y) dA = ∫ ∫
D a g1 ( x )
f ( x , y) dy dx
x
a b
y = g2(x)
y = g1(x)
D
x
a b
y
x = g1(y) x = g2(y)
D
c
∫∫ f (x, y) dA = ∫ ∫
D c g1 ( y )
f ( x , y) dy dx
Example 10.18
Integrate f(x,y) = 4xy – x2 over the region D given in Figure 10.24.
∫∫ f (x, y) dA = ∫ ∫ (4 xy − x ) dy dx
D 0 0
2
Note: Usually, the limits are the most difficult part of the integral to sort
out correctly.
1 y = –x + 1
0 x
1
Here if x limits are from 0 to 1 then the y limits are from 0 to the line y = 1 − x.
1 1
∫ ∫ 2x(1 − x) − x (1 − x) dx
1− x
= 2 xy 2 − x 2 y dx = 2 2
0
0 0
1 1
5x 3 3x 4 1
=
∫
0
(2 x − 5 x + 3 x ) dx = x 2 −
2
3
3
+ =
4 0 12
Example 10.19
Find the volume enclosed by the paraboloid z = x2 + 3y2 and the planes
x = 0, y = 1, y = x, and z = 0 shown in Figure 10.25.
Solution: The volume below the surface within the region D is required. Again
if the limits for x are from 0 to 1, then the limits for y are from 0 to y = x.
1 1
∫∫
D
( x 2 + 3 y 2 ) dA =
∫ ∫ (x
0 x
2
+ 3 y 2 ) dy dx
1 1
∫ ∫ (x
1
= x 2 y + y3 dx = 2
+ 1) − 2 x 3 dx
x
0 0
1
x3 x4 5
= +x− =
3 2 0 6
The next section looks at triple integrals and some of their applications.
y=1
1
D
1
y
x y=x
10.2.4 Triple Integrals
Consider a 3-D region E in space as shown in Figure 10.26.
What is meant by triple integration? This is exactly the same concept as double
integration, that is, splitting the region into smaller volumes and summing.
Again, there are similar applications for triple integrals as double integrals.
Vol (E ) =
∫∫∫ 1. dV
E
with f ( x*, y*, z*) = 1
Mass(E ) =
∫∫∫ d (x, y, z) dV
E
In 2-D space the double integrals were calculated using either Cartesian or
polar coordinates. These concepts can be extended to 3-D space.
To calculate triple integrals in 3-D space, different coordinate systems can
be used, including Cartesian, cylindrical, or spherical coordinates depending on
which will make the integration easier to compute.
The next section starts with evaluating triple integrals in Cartesian coordinates.
10.2.4.1 Cartesian Coordinates
Express the triple integrals as
This integral can be done six different ways since the order of the integration
can be taken with respect to the different variables x, y, or z as required.
Example 10.20
Let E be the region shown in Figure 10.27 which is bounded by the planes
x + y + z = 1, x = 0, y = 0 and z = 0. Suppose E has mass density d(x,y,z) = x2y.
Compute the mass of E of the region bounded by the above planes.
1 1–x1–x–y
Mass (E) = ∫∫∫ x. dV = ∫∫ ∫ x2y dz dy dx
E 0 0 0
=
∫ ∫ x y[z]
0 0
2 1− x − y
0 dy dx =
∫ ∫ x y(1 − x − y) dy dx
0 0
2
y
z
y=1–x
1
1 x
(0, 0, 1)
(b)
E
(0, 1, 0)
(1, 0, 0) y
x
(a)
Figure 10.27 (a) The bounded volume E. (b) The triangular region in the x-y plane.
10.2 Higher-Order Integration 283
1
1 1 1
=
∫ 2 x (1 − x) − 2 x (1 − x) − 3 x (1 − x) dx
0
2 2 3 2 2 3
1
1 1 3 1 4 1 5
=
∫ 6 x
0
2
−
2
x + x − x dx
2 6
x = r cos θ
y = r sin θ
r = x 2 + y2
z=z
θ
r
y
Now consider spherical coordinates (θ, φ, ρ). Here there are two angles and
a distance to travel from the origin as shown in Figure 10.29. Relationships are
given by the following
284 Multivariable Calculus
z P
φ
ρ
z
θ
r
y
z = ρ cos ϕ
r = ρ sin ϕ
x = ρ sin ϕ cos θ
y = ρ sin ϕ sin θ
ρ= x 2 + y2 + z 2
∫∫∫ f (x, y, z) dV =
∫∫∫
Limits!
f (r cos θ, r sin θ, z) r dz dr dθ
E
Jacobian
Spherical:
∫∫∫ f (x, y, z) dV =
∫∫∫
Limits!
f (ρ sin φ cos θ, ρ sin φ sin θ, ρ cos φ) ρ2 sin φ dρ dφ dθ
E
These formulae can be used to find and prove the volumes of standard shapes.
Example 10.21
Find the volume of a sphere (E) of radius r and center the origin as shown
in Figure 10.30.
10.2 Higher-Order Integration 285
r
r
Vol ( E ) =
∫ ∫ ∫ 1.dv
E
=
∫ ∫ ∫ 1ρ sin ϕ d ρ dϕ dθ
0 0 0
2
2r 3 4
= × 2π = π r 3
3 3
So,
4 3
Vol (E ) = πr
3
This is just the usual expression for the volume of a sphere of radius r.
286 Multivariable Calculus
b g −1 ( b ) upper limit
∫ f (x) dx = ∫
a −1
g (a)
f g(u) g′(u) du =
∫
lower limit
f g(u) g′(u) du
For the two-variable case, making a change of variables to u and v gives the
transformation shown in Figure 10.31:
where
∂x ∂y
∂( x , y)
∂u ∂u
= det
∂(u, v) ∂x ∂y
∂v ∂v
y v
T
D
T(D)
x u
Figure 10.31 How a region D in the x-y space gets transformed into the u-v space.
v y
<xv dv, yv dv>
T
(u,v + dv) (u + du, v + dv)
<xu du, yu du>
(x(u,v), y(u,v))
(u, v) (u + du, v)
u x
Figure 10.32 Area gets mapped to a new area under a general transformation.
10.2 Higher-Order Integration 287
x du yu du
= det u
x v dv yv dv
x yu
= det u dvdu
x v yv
∂( x , y)
= dvdu
∂(u, v)
Example 10.22
Evaluate
∫∫ (64xy) dA, where D is the parallelogram with vertices, (–1,3),
D
(1,–3), (3,–1), and (1,5) as shown in Figure 10.33.
Solution: The region is complicated using just the x and y coordinates sys-
tem. So changing variables can make things easier. Let
u = y− x
v = y + 3x
y v
y =x+4 y = –3x + 8 T 8
T(D)
x u
–4 0 4
y = –3x
y =x– 4
1 1
x= ( v − u) and y = (3u + v)
4 4
y=x–4 maps to u = –4
y=x+4 maps to u=4
y = –3x + 8 maps to v=8
y = –3x maps to v=0
4 8
1 1 ∂( x , y)
∫∫ (64 xy) dA = ∫ ∫ 64 4 (v − u) 4 (3u + v) ∂(u, v) dv du
D −4 0
So
∂x ∂x 1 1
∂( x , y) − 1
∂u ∂v = det 4 4 = −
= det
∂(u, v) ∂y ∂y 3 1 4
∂u ∂v 4 4
4 8
1
∫ ∫ 4( v
−4 0
2
+ 2uv − 3u 2 ) −
4
dv du
4 8 4
v3 512 1024
=
∫
−4
+ uv − 3u v du =
3
2 2
0
∫
−4
3
+ 64u − 24u 2 du =
3
∂( x , y) 1
=
∂(u, v) ∂(u, v)
∂( x , y)
∂( x , y, z )
∫∫∫ f (x, y, z) dz dy dx = ∫∫∫ f ( x(u, v, w)……) ∂(u, v, w) dw dv du
E
x x x
u v w
∂( x , y, z )
= det yu yv yw .
∂(u, v, w) z z z
u v w
10.3 Applications
There are many physical applications of double and triple integration, most of
which depend on the idea of splitting a region into smaller pieces and summing
over all the pieces. Start by looking at applications of double integrals, which
shows the principle involved.
∫∫ f (x, y) dA lim ∑
D
∆A →0
all pieces
f ( x*, y*)∆A
The region is chopped into pieces. Then pick a point (x*, y*) and sum over all
the pieces with f(x*, y*) multiplied by the width of all the pieces ΔA.
This forms the basis of all applications as shown in the next few examples.
(x*, y*)
∆A
Volume ≈ lim
∆A→0 ∑
all pieces
f ( x*, y*)∆A
Volume =
∫∫ f (x, y) dA
D
(10.18)
z
z = f (x, y)
D y
x ∆A
D
(x*, y*)
∆A
Solution: Here, if f(x*,y*) = 1, then one has just the area of the small piece
and then summing gives
Area ≈ lim
∆A→0 ∑ 1.∆A
all pieces
Area =
∫∫ 1dA
D
(10.19)
Solution:
Mass ≈ lim
∆A→0 ∑
all pieces
d ( x*, y*)∆A
Mass =
∫∫ d (x, y) dA
D
(10.20)
D
(x*, y*)
∆A
nj
θj
dAj
ni Aj
θi
dAi
Ai
by Fij and defined as the fraction of the radiation leaving the surface i that is
intercepted then by the surface j. This can be represented as
The mathematical formula for this view factor needs some work to
derive it and is given as
1 cos θ i cos θ j
Fij =
Ai ∫∫
Ai A j
π R2
dA j dAi (10.21)
Mass(E ) = M =
∫∫∫ d (x, y, z) dV
E
(10.22)
10.3 Applications 293
1
x=
M ∫∫∫ x d (x, y, z) dV
E
(10.23)
1
y=
M ∫∫∫ y d (x, y, z) dV
E
(10.24)
1
z=
M ∫∫∫ z d (x, y, z) dV
E
(10.25)
Solution: See Figure 10.40. By radial symmetry, the center of mass must be
on the z axis with more mass lower down so it should be closer to the z = 0
plane. By symmetry, x = 0 and y = 0 are already known. Now to work out
z use the formula given by Equation 10.25 with d = 1 gives
1 1
z=
M ∫∫∫ z d (x, y, z) dV = M ∫∫∫ zd dV
E E
294 Multivariable Calculus
z = 1− x2 − y2
1
1
D
1
y
First, work out the mass M of the region using the mass formula given
by Equation 10.22:
M=
∫∫∫ 1dV
E
1− x 2 − y 2
=
∫∫ ∫
D 0
1 dz dy dx
=
∫∫ (1 − x
D
2
− y 2 ) dy dx
2π 1
π
∫ ∫ (1 − r ) r dr dθ = 2
0 0
2
1
z=
M ∫∫∫ zd dV
E
1− x 2 − y 2
2
=
π ∫∫ ∫
D 0
z dz dy dx
2 1
=
π ∫∫ 2 (1 − x
D
2
− y 2 )2 dy dx
2π 1 2π 1
1 1 1 2 3
=
π ∫∫
0 0
(1 − r ) r dr dθ =
2 2
π ∫0
− 6 (1 − r ) dθ
0
2π
1 1 1
=
π ∫ 6 dθ = 3
0
1
∴ center of mass is at ( x , y , z ) = 0, 0, .
3
Problems
10.1 Given that f(x,y) = x3y + y3, evaluate the following.
∂f ∂f
a. b.
∂x (1,1) ∂y ( 2,1)
a. fx b. f y c. fxy d. f yx
10.3 Given that f(x,y) = xy2 –5xy, calculate the directional derivative of f in
the direction of 〈3,1〉 at the point (1,1).
3 1
∫ ∫ (x y − 5x) dy dx
0 0
2
296 Multivariable Calculus
10.5 Find the volume of the solid tetrahedron enclosed by the plane
2x + y + z = 4 and the coordinate planes.
10.6 In Figure 10.41, a cylinder of height H and radius R is shown. Show that
the volume of the cylinder is given by the formula V = πR2 H.
R
R
r (t ) = 〈 f (t ), g(t ), h(t )〉
The derivative of r(t) with respect to time t is defined as r ′(t ) and is given by
r (t + ∆t ) − r (t )
r ′(t ) = lim (11.1)
∆t →0 ∆t
This is a tangent vector to a space curve. Also, this can be thought of as the instan-
taneous velocity, as shown in Figure 11.1.
As, Δt → 0 the vector PQ tends to the tangent vector at the point P.
f (t + ∆t ) i + g(t + ∆t ) j + h(t + ∆t ) k − f (t ) i + g(t ) j + h(t ) k
= lim
∆t → 0 ∆t
297
298 Vector Calculus
r(t)
r(t + ∆t)
So,
Therefore, to find the derivative of r(t) with respect to time t, just differentiate
each of the components of r(t), that is, given that
r (t ) = 〈 f (t ), g(t ), h(t )〉
then
Example 11.1
If r (t ) = 〈t , t 2 , t 3 〉, then
This is the directional vector of the tangent line space curve at any point
t for the vector function r (t ).
11.1 Differentiation and Integration of Vectors 299
Example 11.2
∫ (t i + t
0
2
)
j + t 3 k dt
1
1 1 1
= t2 i + t3 j + t4k
2 3 4 0
1 1 1 1 1 1
= i + j + k or 〈 , , 〉
2 3 4 2 3 4
Example 11.3
1 3t t3 e 3t
∫ i + t j + e k dt = ln t i + j +
t
2
3 3
k+a
(
Where a is a constant vector, that is, a = c1 i + c2 j + c3 k . )
Example 11.4
Solution:
r (t ) =
∫ ( 2 i + 4tj − 6t k ) dt 2
r (t ) = 2t i + 2t 2 j − 2t 3 k + c
r (0) = 0 + 0 + 0 + c = j + k
Therefore,
r (t ) = 2t i + 2t 2 j − 2t 3 k + j + k or r (t ) = 2t i + (2t 2 + 1) j + (1 − 2t 3 )k
300 Vector Calculus
11.2 Vector Fields
Previously, the functions had different input variables but only one output variable.
For example,
F ( x , y, z ) = 〈 M ( x , y, z ), N ( x , y, z ), P( x , y, z )〉 3 inputs, 3 outputs
a vector field is a more general function. So to every point in the x-y plane or 3-D
space it assigns a vector.
Some simple vector fields can be seen by graphing them as follows.
Example 11.5
What does the vector field F ( x , y) = 〈 x , y〉 look like? See Figure 11.2.
What is the magnitude of the vector field, that is, F ( x , y) ? The magni-
tude of a vector is given by the square root of all the components squared
and added together. Therefore,
(x, y)
1 2 x
Look at points and see what vector is given. Take a general point (x, y)
in space and this vector from the origin is the vector you put at the
location (x, y).
F ( x , y) = x 2 + y 2 = r
So the magnitude of the vector field at any position is just the dis-
tance from the origin to that point, pointing radially outward as shown
in Figure 11.2.
Example 11.6
Sketch the vector field given by
〈 x , y〉
F ( x , y) =
x 2 + y2
Solution: Again this is the same vector field as in Example 11.5 but now the
magnitude is always equal to unity as shown in Figure 11.3 (the arrows are
all of length 1 unit).
The magnitude of the vector field is given by
〈 x , y〉 x 2 + y2
F ( x , y) = = =1
x 2 + y2 x 2 + y2
1 2 x
〈 x , y〉
Figure 11.3 The vector field for F ( x , y) = .
x 2 + y2
Example 11.7
Sketch the vector field given by, F ( x , y) = 〈 y, 0 〉.
302 Vector Calculus
Solution: This vector field shown in Figure 11.4 is known as shear flow in
fluid dynamics and models flow of water near boundaries.
y
Example 11.8
Consider the vector force field for gravity. For gravitational attraction the
magnitude of the force varies as reciprocal of the distance squared, that is,
1
Force of gravity = F ( x , y, z ) ~ (distance squared)
r2
It turns out that the vector field for the force due to gravity can be arrived
at by starting with the basic vector field F ( x , y, z ) = 〈 x , y, z 〉, multiplying by
a constant and a negative sign, and making the modulus proportional to the
reciprocal of the distance squared.
The sketch is shown in Figure 11.5.
z
− k 〈 x , y, z 〉
F ( x , y) = 3
( x 2 + y2 + z 2 ) 2
11.3 Line Integrals
The basic idea of a line integral is to integrate along a curve C in 2-D or 3-D, as
shown in Figure 11.6.
∫ f (x, y) ds = lim ∑
C
∆s → 0
all pieces
f ( x*, y*)∆s
Length = 1.ds
∫
C
(11.3)
2. If d(x, y) is the mass density (mass/length) of the wire, then the total mass
of the wire is given by
Mass of wire =
∫ d (x, y) ds
C
(11.4)
Note: These formulae can be used to calculate how long or heavy are the steel
cables needed in the design of suspension bridges.
∆s
To compute the ds-type of integral, first, the curve C, as shown in Figure 11.7,
needs to be parameterized.
Here, x = x(t), y = y(t), and a ≤ t ≤ b. Then ds can be found using the Pythagorean
theorem as
ds = dx 2 + dy 2
Then
∫ f (x, y) ds = ∫ f (x, y)
C C
dx 2 + dy 2
b 2 2
dx dy
∫
C
f ( x , y) ds =
∫
a
f ( x (t ), y(t )) + dt
dt dt
(11.5)
∆s ds
dy
dx
C
Example 11.9
If the density of a wire is d(x,y) = x, compute its mass when the wire is the
quarter circle of radius 2 as shown in Figure 11.8.
Mass =
∫ d (x, y) ds = ∫ x ds
C C
(11.6)
To parameterize the curve, since the wire is in the form of a circle, then
the natural parameter is in terms of the angle made, that is, t.
x (t ) = 2 cos t x (t ) = −2 sin t
2
C
x
2
π
0≤t ≤
2
Mass =
∫ 2 cos t
0
(−2 sin t )2 + (2 cos t )2 dt
π
2 π
=
∫ 0
4 cos t dt = 4[sin t ]02 = 4
Here a vector field acts in the region given by F ( x , y). Again, the line is split into
small segments as shown in Figure 11.9.
The component of the vector field in the direction of the line is given by
F ( x*, y*).∆r , so summing along the whole line gives
Work =
∫ F. dr
C
(11.8)
306 Vector Calculus
=
∫ F (x, y).d r
C
=
∫ M (x, y) dx + N (x, y) dy
C
b
dx dy
=
∫ M (x(t), y(t)) dt + N (x(t), y(t)) dt dt
a (11.9)
y
Vector field F(x, y) = M(x, y), N(x, y)
dr
dy
C dx
Example 11.10
Work =
∫ F. d r
C
=
∫ x dx + y dy
C
2
dx
x=t =1
dt
dy
y = −t 2 + 1 = −2t
dt
y
C y = –x2 + 1
x
–1 1
=
∫ (t) (1) + (−t
−1
2 2
+ 1)(−2t ) dt
∫ [ 2t + t 2 − 2t ] dt =
2
= 3
3
−1
308 Vector Calculus
11.3.3 Summary of Results
A general path in 3-D space is shown in Figure 11.12.
The two types of line integrals can be summarized by the formulae as follows:
ds–type:
b 2 2 2
dx dy dz
∫
C
f ( x , y, z ) ds =
∫a
f ( x (t ), y(t ), z (t )) + + dt
dt dt dt
dr − type:
b
dx dy dz
∫
C
F ( x , y, z ).d r =
∫ M (x(t), y(t), z(t)) dt + N (x(t), y(t), z(t)) dt + P(x(t), y(t), z(t)) dt dt
a
x = x(t)
y = y(t)
z = z(t)
a≤t≤b
C
11.4 Gradient Fieds
A special type of vector field is known as a gradient field.
Given f(x,y), defining a vector field, the F ( x , y) = ∇f
∇f is called its “gradient field.”
Example 11.11
1. Given the function f(x,y) = xy + y3, the gradient field is
F ( x , y) = ∇f = 〈 f x , f y 〉 = 〈 y, x + 3 y 2 〉
11.4 Gradient Fieds 309
1
2. Given the function f ( x , y) = ( y − x 2 ), then the gradient field is
2
given by
1
F ( x , y) = ∇f = 〈 f x , f y 〉 = 〈− x , 〉
2
1
Now f ( x , y) = ( y − x 2 ). The level curves for this function, that is,
2
1
f =0= (y − x2) y = x2
2
1
f =1= (y − x2) y = x2 + 2
2
1
f = −1 = (y − x2) y = x2 − 2
2
and so on as shown in Figure 11.13. These show the level curves of f(x,y).
Also the gradient vector ∇f is always perpendicular to the level curves
pointing to the greatest increase.
1
∇f = 〈 f x , f y 〉 = 〈− x , 〉
2
f=1
f=0
f = –1
2
–2
y
Vector field F(x, y)
End
Start
C
Example 11.12
∫ F (x, y). d r
C
∫ F. d r
C
= (32 + 0) − (0 2 + 0) = 9
11.4 Gradient Fieds 311
y = 9 – x2
x
3
∫ F. d r = ∫ F. d r
C C
(11.11)
C End
Start
Figure 11.16 Two paths with the same start and end points.
Note: Systems that are reversible are conservative and those that are non-
conservative in nature mean that energy is lost during the process as waste
or to entropy.
312 Vector Calculus
Figure 11.17 A closed curve with the same start and end points.
∫ F. d r = 0
C
(conservativeness) (11.12)
Can it be said that all vector fields are conservative? No, not all vector fields
are conservative. Consider the following vector field (also see Figure 11.18)
〈− y, x 〉
F ( x , y) =
x 2 + y2
〈− y, x 〉
Figure 11.18 The vector field given by F ( x , y) = .
x 2 + y2
11.4 Gradient Fieds 313
∫ F. d r > 0
C
F ( x , y) = 〈 M ( x , y), N ( x , y)〉 = ∇f
then we need to find some function f such that we have the following as shown in
Figure 11.19.
∂ ∂
∂x ∂y
M N
∂ ∂
∂y ∂x
Same
Figure 11.19 Showing that the order of differentiating does not matter.
Example 11.13
∂N
= Nx = 0
∂x
and
∂M
= My = 1
∂y
Example 11.14
F ( x , y) = ∇f
314 Vector Calculus
Example 11.15
Compute the following for line integral
∫ F. d r
C
∫ F. d r = f (end) − f (start)
C
= f (4, 0) − f (0, 2) = 4 3 − 0 = 64
(0, 2)
C
x
(4, 0)
Notation: Given curves C1, C2, C3, and C4 as shown in Figure 11.21.
The whole curve is partitioned as C = C1 + C2 + C3 + C4.
Also, shown in Figure 11.22 is the negative of a curve.
The same curve as C but in the opposite direction is called –C.
Property 3: If F ( x , y) = 〈 M ( x , y), N ( x , y)〉 is conservative, then My = Nx.
11.4 Gradient Fieds 315
Note: Connected means that you can get from a point to another without
leaving the region.
C4
C3
C1
C2
–C
Example 11.16
Some examples of regions D that are connected and simply connected are
shown in Figure 11.23.
Theorem: If F ( x , y) = 〈 M ( x , y), N ( x , y)〉 is defined on a simply connected
domain and My = Nx, then it can be said that F is conservative.
D
D D
Connected: Y Y N
Simply connected: Y N Y
Example 11.17
An example of a famous conservative vector field is the gravitational vector
force field as shown in Figure 11.24. Here the vector field is given by
− k 〈 x , y, z 〉
F ( x , y) = 3
( x 2 + y2 + z 2 ) 2
k
f ( x , y, z ) = 1
( x 2 + y2 + z 2 ) 2
and so F ( x , y) = ∇f .
11.5 Green’s Theorem
Having calculated line integrals for different paths as shown in Figure 11.25,
using the formula
∫ M (x, y) dx + N (x, y) dy
C
∫∫ D
f ( x , y) dA
11.5 Green’s Theorem 317
If the curve C is a simple, closed and positively oriented curve that surrounds
a two-dimensional region D.
The following definitions are used about the curve.
Positively oriented – As you travel along the curve C the region D is toward
the left.
∫ M (x, y) dx + N (x, y) dy = ∫∫ (N
C D
x − M y ) dA (11.13)
This states that the line integral along the curve C is identical to a double inte-
gral over the region D.
Note: Just as the fundamental theorems of calculus and line integrals turn line
integrals into a calculation of the function at end points, Green’s theorem turns
an area integral into an integral around a boundary line of the area. It is a
dimension reducing method.
Example 11.18
∫ F.d r
C
Now, since F is not a conservative vector field, then it is the case that
∫ F.d r ≠ 0
C
So, this problem can be done either directly by parameterizing the three
curves and solving the F .dr integral along each one.
Alternatively, you can use Green’s theorem and change this problem into
a double integral using Equation 11.13:
∫ F.d r = ∫∫ (N
C D
x − M y ) dA
C = C1 + C2 + C3
C2
C3
x
C1 5
∫∫ (N
D
x − M y ) dA =
∫∫ y dA
D
π
2 5
∫∫ (N
D
x − M y ) dA =
∫∫ y dA = ∫ ∫ (r sinθ )r dr dθ
D 0 0
π π
2 2
1 125 125
=
∫ 3 [r sinθ ] dθ = ∫
0
3 5
0
0
3
cos θ dθ =
3
With a closed curve, Green’s theorem can be used to save some effort in the
calculations.
∫ F.d r = ∫ ∫ (N
C D
x − M y ) dA = 0
This ties into the previous result for conservative closed vector fields (Equation 11.12).
Property 2: The above only relies on My= Nx.
〈− y, x 〉
If F ( x , y) = 2 as shown in Figure 11.28, then it can be shown that My = Nx,
x + y2
but the vector field is not conservative.
So,
∫ F. d r = ∫ ∫ (N
C1 D1
x − M y ) dA = 0
Also, what about along C2, since Nx and My are not defined at (0, 0).
∫ F. d r = ∫ ∫ (N
C2 D2
x − M y ) dA ≠ 0
C1
D1
C2
D2 x
C2
D1
C1
〈− y, x 〉
F ( x , y) = with My = Nx
x 2 + y2
So,
∫ F.d r − ∫ F. d r = ∫
C2 C1 C2 −C1
F. d r =
∫ ∫ (N
D2
x − M y ) dA = 0
11.6 Divergence and Curl of Vector Fields 321
∫ F.d r = ∫ F.d r
C1 C2
Now let’s look at some examples that illustrate these concepts with vector
fields.
Example 11.19
div F = M x + N y = 1 + 1 = 2 > 0
curl F = N x − M y = 0 + 0 = 0
x
1 2
Example 11.20
div F = M x + N y = 0 + 0 = 0
Example 11.21
div F = M x + N y = 0 + x = x
curl F = N x − M y = y − 1 = y − 1
11.6.2.1
Curl Form
Using the definition of Green’s theorem given by Equation 11.13,
324 Vector Calculus
∫ M (x, y) dx + N (x, y) dy = ∫ ∫ (N
C D
x − M y ) dA
This can now also be written as using the definition of the curl:
∫ F. d r = ∫ ∫ curlF dA
C D
(11.16)
So, the curl F represents how much counterclockwise rotation there is at each
point and if all of these are added up, this then gives the total rotation around the
boundary.
Also note that now if F is a conservative (My = Nx), then this implies that
curl F = 0.
11.6.2.2
Divergence Form
Figure 11.35 shows a bounded region within a vector field.
∫ ( F.nˆ ) ds = ∫ M dy − N dx
C C
y
n
F(x, y) = M(x, y), N(x, y)
dr
dy
D C dx
∫ M dy − N dx = ∫ ∫ (M
C D
x − − N y ) dA =
∫ ∫ (M
D
x + N y ) dA =
∫ ∫ div F dA
D
∫ ( F.nˆ ) ds = ∫ ∫ div F dA
C D
(11.17)
This shows that the net flow out at each point all added up together gives the
total flux out of the curve shown in Figure 11.35.
11.6.3 3-Dimensional Definitions
It is easier to remember the definitions for grad F , div F , and the curl F in
terms of the “del or nabla notation.”
Let
∂ ∂ ∂
∇ 〈 , , 〉
∂x ∂y ∂z
grad F = ∇F (11.20)
curl F = ∇ × F (11.22)
Note: The div F still measures the net flow out at a point. However, the curl
F is more difficult to see. Generally, the curl F points in the direction of the
“broom handle” with the largest counterclockwise rotation of the paddle wheel.
The magnitude of curl F is the amount of counterclockwise rotation.
326 Vector Calculus
Example 11.22
Given the vector field,
F ( x , y, z ) = 〈 M ( x , y, z ), N ( x , y, z ), P( x , y, z )〉 = 〈 x 2 , xy, z 〉,
find the divergence and curl of the vector field. Using the definitions gives
div F = M x + N y + Pz = 2 x + x + 1 = 3 x + 1
i j k
∂ ∂ ∂
curl F = ∇ × F =
∂x ∂y ∂z
x2 xy z
= 〈 0 − 0, 0 − 0, y − 0 〉
= 〈 0, 0, y 〉
Proofs:
i j k
∂ ∂ ∂
curl (∇f ) =
∂x ∂y ∂z
fx fy fz
i j k
3. curl F = ∇ × F = ∂ ∂ ∂ = 〈 Py − N z , M z − Px , N x − M y 〉
∂x ∂y ∂z
M N P
11.7 Surface Integration
11.7.1 Parametric Surfaces
For a line there was one parameter. Generally for surfaces there are more param-
eters starting with two, say, u and v.
z
v
(x(u, v), y(u, v), z(u, v))
S
(u, v)
D
u y
Example 11.23
A graph of a function in terms of an equation given the surface z = 4 – x2
– y2 and z ≥ 0 is shown in Figure 11.37.
To parameterize the surface, let x = u, y = v, then z = 4 – u2 – v2, where
(u, v) ∈ D (Figure 11.38).
z = 4 – x2 – y2
2
y
2 u2 + v2 ≤ 4
D
u
2
Example 11.24
A spherical shell of radius 5 is shown in Figure 11.39. In spherical coordi-
nates, this a sphere with radius ρ = 5.
11.7 Surface Integration 329
x = 5sin v cos u
y = 5sin v sin u
z = 5cos v
5
5
π
D
u
0
2π
z = f (x, y)
S “u = x”, “v = y”
x=u
y=v
z = f (x, y)
(u, v) D
D y
x
10
4
4
π
0≤u≤ and 0 ≤ v ≤ 10
2
11.7.2 Surface Integrals
The idea of a surface is similar to a region as shown in Figure 11.43.
Splitting the surface into small pieces and summing gives the definition of a
surface integral as
∫ ∫ f (x, y, z) dS lim ∑
S
∆S →0
all pieces
f ( x*, y*, z*)∆S (11.23)
Area ( S ) =
∫ ∫ 1dS
S
(11.24)
Mass ( S ) =
∫ ∫ d (x, y, z) dS;
S
(11.25)
y
x
∂( x , y, z )
∫ ∫ g(x, y, z) dS = ∫ ∫ g[x(u, v), y(u, v), z(u, v)]
S D
∂(u, v)
dv du (11.26)
332 Vector Calculus
x xv
u
∂( x , y , z )
Now, the Jacobian is = det yu yv
∂(u, v) z
u z v
The generalized determinant of a rectangular matrix is given by
x xv 2 2 2
u
x xv x xv y yv
u u u
det yu yv det + det + det (11.27)
yu yv zu z v zu z v
z z v
u
z
v
S
T
D
u y
Example 11.25
Find the area of x + y + z = 1 in the first octant as shown in Figure 11.45.
Parameterizing the surface gives x = u, y = v, and z = 1 − u − v. So, the
area is given by the formula
∂( x , y, z )
Area( S ) =
∫ ∫ 1dS = ∫ ∫ 1
S D
∂(u, v)
dv du
x x v 2 2 2
u
∂( x , y, z ) 1 0 1 0 + det 0 1
= det yu yv = det + det
∂(u, v) z 0 1 −1 −1 −1 −1
u z v
Area( S ) = 3
∫ ∫ dv du
D
3
= 3 Area( D) =
2
11.7 Surface Integration 333
v
z
D
1 u
(0, 0, 1)
(0, 1, 0)
D
(1, 0, 0) y
f ( x , y) ≈ f ( a , b ) + ( x − a ) f x ( a , b ) + ( y − b ) f y ( a , b )
z = f (x, y)
z
y
(a, b)
Now, this is a plane accurate at (a, b) but approximate near (a, b) and is called
the tangent plane (or best linear approximation).
Example 11.26
Find the tangent plane to the surface z = x + 5xy2 at point (1, 2, 3).
Solution: Let the function f(x, y) = x + 5xy2. Therefore, using Equation 11.28
gives
Let
z = f (1, 2) + ( x − 1) f x (1, 2) + ( y − 2) f y (1, 2)
f (1, 2) = 21
fx = 1 + 5 y2 → f x (1, 2) = 21
f y = 10 xy → f y (1, 2) = 20
z = 21 + 21( x − 1) + 20( y − 2)
f ( x , y) = x 2 + y 2 f (−1, 2) = 5
fx = 2 x → f x (−1, 2) = −2
fy = 2 y → f y (−1, 2) = 4
z = 5 − 2( x + 1) + 4( y − 2)
2 x − 4 y + z = −5
(–1, 2, 2)
Figure 11.47 Surface and its normal vector at the given point.
z = f (a, b) + ( x − a) f x (a, b) + ( y − b) f y ( a, b)
z
z = f (x, y)
n
(a, b) y
This can be rewritten as x, y, and z terms on one side with all the constants on
the other side as
Therefore, the normal vector also normal to the surface at the given point is
n
Normal vectors can be pointing
up or down
“Oriented upwards”
n
r
“Oriented outwards”
E n
r
r
There are different ways to find the normal vectors to different surfaces. For a
general graph as shown in Figure 11.52 the normal vector is simply given by using
the formula for z = f(x,y) as
〈− f x , − f y , 1〉
nˆ = (11.30)
f x2 + f y2 + fz2
z = f (x, y)
S
The case of a sphere is shown in Figure 11.53. The normal vector is given by
using the formula given by Equation 11.29 as
〈 x , y, z 〉 (11.31)
nˆ =
x + y2 + z 2
2
338 Vector Calculus
n r n
r
r
In the case of a cylindrical surface is shown in Figure 11.54. The normal vector
is given by using the formula given in Equation 11.31 as
〈 x , y, 0 〉 (11.32)
nˆ =
x 2 + y2
Generally, for a surface of the form r = (u, v), the normal vector is
ru × rv
n̂ = (11.33)
ru × rv
z
n
n F
S
∆S
Example 11.28
Compute the flux of the vector field F = 〈 x , y, z 〉 across the surface S, where S is
given by z = 4 – x2 – y2 and z ≥ 0 and oriented upward as shown in Figure 11.56.
〈− f x , − f y , 1〉
Solution: The normal vector n̂ is given by the formula nˆ = ,
which gives f x2 + f y2 + fz2
〈2 x , 2 y, 1〉
nˆ =
4 x 2 + 4 y2 + 1
〈2 x , 2 y, 1〉
Flux =
∫∫ (F.nˆ) dS = ∫∫ 〈x, y, z 〉.
S S
4 x 2 + 4 y2 + 1
dS
2 x 2 + 2 y2 + z
=
∫∫ S
4 x 2 + 4 y2 + 1
dS
340 Vector Calculus
4
n
S z = 4 – x2 – y2
2
y
2 x 2 + 2 y2 + z (2u 2 + 2v 2 ) + (4 − u 2 − v 2 ) ∂( x , y, z )
=
∫∫
S
4 x 2 + 4 y2 + 1
dS =
∫∫
D
4u 2 + 4 v 2 + 1 ∂(u, v)
dv du
2 u2 + v2 ≤ 4
D
u
2
x xv
u
∂( x , y, z )
= det yu yv
∂(u, v) z
u z v
1 0
= det 0 1
= (1)2 + (2 v)2 + (2u)2 = 1 + 4u 2 + 4 v 2
−2u −2 v
11.7 Surface Integration 341
2 x 2 + 2 y2 + z (4 + u 2 + v 2 )
=
∫∫
S
4 x 2 + 4 y2 + 1
dS =
∫∫
D
4u 2 + 4 v 2 + 1
1 + 4u 2 + 4 v 2 dv du
=
∫∫ (4 + u
D
2
+ v 2 ) dv du
∫ ∫ (4 + r )r dr dθ = 24π
0 0
2
Example 11.29
Find the surface area of the cylinder shown in Figure 11.58.
Area is given by the formula
2π H
∂( x , y, z )
Area( S ) =
∫∫S
1 dS =
∫ ∫1
0 0
∂(θ , z )
dz dθ
x = R cos θ
y = R sin θ
z=z
S 0 ≤ θ ≤ 2π
0≤z≤H
R
R
Now,
x xz − R sin θ
θ
0
∂( x , y, z )
= det yθ yz = det R cos θ 0
∂(θ , z )
z z z 0 1
θ
342 Vector Calculus
∂( x , y, z )
=R
∂(θ , z )
2π H
=
∫ ∫ R dz dθ
0 0
2π
=
∫ RH dθ = 2π RH
0
Example 11.30
Compute the flux out of a sphere given by x2 + y2 + z2 = 4 for the vector field
given by F = 〈 x , y, z 〉 and shown in Figure 11.59.
2
n
2
2
〈 x , y, z 〉
Flux =
∫∫ (F.nˆ) dS = ∫∫ 〈x, y, z 〉.
S S
x + y2 + z 2
2
dS
x 2 + y2 + z 2
=
∫∫
S
x +y +z
2 2 2
dS =
∫∫
S
x 2 + y 2 + z 2 dS
11.8 Stokes’ Theorem 343
x = 2 sin ϕ cos θ
y = 2 sin ϕ cos θ
z = 2 cos ϕ
0 ≤ θ ≤ 2 π and 0 ≤ φ ≤ π
2π π
∂( x , y, z )
=
∫ ∫ρ
0 0
∂(ϕ ,θ )
d ϕ dθ
x xθ 2 cos ϕ cos θ
ϕ −2 sin ϕ sin θ
∂( x , y, z )
= det yϕ yθ = det 2 cos ϕ sin θ 2 sin ϕ cos θ = 4 sin ϕ
∂(ϕ , θ )
zϕ zθ 2 sin ϕ 0
2π π 2π
=
∫ ∫ 2(4 sin ϕ ) dϕ dθ = ∫ 16 dθ = 32π
0 0
11.8 Stokes’ Theorem
There are many theorems that relate the integral over a region to the integral over the
boundary of the region. Some that have already been considered are the following.
In 1-D, Figure 11.60 shows a line interval from [a, b].
x
a b
∫ ∇f .d r = f (end) − f (start)
C
C End
Start
∫∫ curlF dA = ∫ F.d r
D C
This is Green’s theorem. This can be extended to a 3-D surface and its bound-
ary, as shown in Figure 11.63.
Suppose C is a simple, closed and positively oriented curve with respect to the
surface S as earlier. Stokes’ theorem states
∫∫
S
^ dS =
(curl F . n)
∫
C
F .dr (11.35)
n
S
Example 11.31
Verify Stokes’ theorem for the vector field F = 〈 x , z , − y〉 and the surface S,
where S is defined by x + y + z = 1 and x2 + y2 ≤ 1.
The surface S is given by the plane being effectively chopped by cylinder
of radius 1 as shown in Figure 11.65.
To verify Stokes’ theorem the following has to be shown:
∫∫ (curlF.nˆ) dS = ∫ F.d r
S C
(11.36)
346 Vector Calculus
S
1 n
D
1
1
∫∫ (curlF.nˆ) dS
S
i j k
∂ ∂ ∂
curl F = ∇ × F = = 〈−2, 0, 0 〉
∂x ∂y ∂z
x z −y
〈1, 1, 1〉
A unit normal vector to the surface is given by nˆ =
3
〈1, 1, 1〉 −2
∫∫ (curlF.nˆ) dS = ∫∫ 〈−2, 0, 0〉.
S S
3
ds =
∫∫
S
3
dS
y= y
z = 1− x − y
where (x, y) ∈ D
−2 ∂( x , y, z )
=
∫∫
D
3 ∂( x , y)
dy dx
11.9 Divergence Theorem 347
x xy
x 1 0
∂( x , y, z )
= det yx yy = det 0 1
= (1)2 + (−1)2 + (−1)2 = 3
∂( x , y)
z x z y −1 −1
=
∫∫ −2 dy dx = −2Area(D) = −2π
D
∫ F.d r = ∫ x dx + z dy + (− y) dz = ∫ x dx + z dy − y dz
C C C
x = cos t
y = sin t
z = 1 − cos t − sin t
0 ≤ t ≤ 2π
=
∫ x dx + z dy − y dz
C
2π
=
∫ [(cos t)(− sin t) + (1 − cos t − sin t)(cos t) − (sin t)(sin t − cos t)] dt
0
2π
=
∫ (−1 + cos t − cos t sin t) dt
0
Using trigonometric identities and properties of the cosine and sine func-
tions gives
2π
1
=
∫ −1 + cos t − 2 sin 2t dt = −2π
0
So the left-hand side equals the right-hand side and Stokes’ theorem is
verified.
11.9 Divergence Theorem
In Figure 11.66 there is a 3-D region (E) bounded by a 2-D surface-oriented out-
ward. If all the flux at each point in the region (E) are added together, as shown
348 Vector Calculus
in Figure 11.67, then this will give the net flux out of the surface. Mathematically,
this is stated using the divergence theorem as
Example 11.32
Verify the divergence theorem for the vector field F = 〈 x , y, z 〉 for the follow-
ing sphere of radius 1 and with a boundary surface S as shown in Figure 11.68.
4
∫∫∫ divF dv = ∫∫∫ 3 dV = 3Vol(E) = 3 3 π (1)
E E
3
= 4π
1
n
1
1
〈 x , y, z 〉
∫∫ (F.nˆ) dS = ∫∫ 〈x, y, z 〉.
S S
x + y2 + z 2
2
dS
x 2 + y2 + z 2
∫∫
S
x 2 + y2 + z 2
dS
0 ≤ θ ≤ 2π and 0 ≤ φ ≤ π
x2 + y2 + z2 = ρ 2 and ρ = 1
This gives the following:
∂( x , y, z )
=
∫ ∫1S
∂(ϕ ,θ )
dS
2π π
=
∫ ∫ (sin ϕ ) dϕ dθ
0 0
2π
∫ − cosϕ
π
= 0
dθ
0
2π
=
∫ 2 dθ = 2 × 2π = 4π
0
11.10 Applications
Example 11.33: Fluid Dynamics of Smoke Flow
In fluid dynamics one of the main tasks is to find the velocity field describ-
ing the flow in a given region. One can make use of the laws of mechanics,
that is, the conservation of mass to derive the continuity equation.
Any “small” fluid element can be assigned a velocity v(x,t) and average
density ρ(x,t).
Considering a volume V bounded by a surface S, the mass inside the
volume is given by
∫ ρ dV
V
d ∂ρ
V =−
dt ∫ ρ dV = − ∫ ∂t dV
V V
(11.39)
This must be equal to the total rate of mass flux out of V if mass is con-
served. The rate of mass flux out of any small element dS of S is given as ρv.dS.
Integrating over the whole surface gives the rate of mass flux out of V as
=
∫ ρv. dS
S
Now using the divergence theorem this can be written as a volume integral as
11.10 Applications 351
∫ ∇.(ρv) dV
V
(11.40)
This then implies that for mass to be conserved, Equations 11.39 and
11.40 must be equal for any volume V giving the continuity equation as
∂ρ
+ ∇.( ρ v) = 0
∂t
In any flow region, the flow equations are solved to a set of conditions
that act at the boundary.
Applications in smoke control situations occur when the smoke temperature
of the smoke reservoir does not change. In order to maintain the smoke layer
height, the amount of smoke extracted from the smoke reservoir equals the
amount of smoke flow into the reservoir from the fire plume, that is, ∇.(ρv) = 0.
V = 5 cos(−t ) + 50
P = 5 sin(−t ) + 30
0 ≤ t ≤ 2π
The formula for the work done is given by using Equation (11.8),
Work =
∫ P. dV
C
2π
=
∫ (−5 sin(t) + 30)(−5 sin(t)) dt
0
2π
=
∫ (25 sin (t) − 150 sin(t)) dt = 25π J
0
2
352 Vector Calculus
30 5
50 V
i j k
curl v = ∇ × v = ∂ ∂ ∂ = i ∂w − ∂v + ∂u ∂w
j −
∂v ∂u
+ k −
∂x ∂y ∂z
∂y ∂z ∂z ∂x ∂x ∂y
u v w
Since, in general, the horizontal velocities are much larger than the ver-
tical velocities and vertical scales are much smaller than the horizontal
scales, in the x and y components of the above expression we can neglect
the terms in the vertical velocity, giving
∂v ∂u ∂v ∂u
curl v ~ i − + j + k −
∂z ∂z ∂x ∂y
∂v ∂u (10 ms −1 )
The first two terms and have typical magnitudes ~ 10 −3 s −1.
∂z ∂z 4
(10 m)
∂v ∂u (10 ms −1 )
The third term − has a typical magnitude ~ 10 −5 s −1.
∂x ∂y (10 6 m)
But for irrotational flows, the vorticity is 0 (i.e., ∇ × v = 0) and then the
velocity can be represented as the gradient of a scalar function. This is
known as the velocity potential f such that v = ∇f .
∫ M (x, y) dx + N (x, y) dy = ∫∫ (N
C D
x − M y ) dA
∫ x dy = ∫∫ (N
C D
x − M y ) dA =
∫∫ (1 − 0) dA = ∫∫ dA = Area (D)
D D
or it can be given as
− ∫ y dx = ∫∫ (N
C D
x − M y ) dA =
∫∫ (0 − −1) dA = ∫∫ dA = Area (D)
D D
Also, it can be shown that a combination of the above two, that is,
1 1 1
2 ∫ x dy − y dx = ∫∫ 2 − − 2 dA = ∫∫ dA = Area (D)
C D D
This principle has been used to find areas of shapes as one draws a curve
around the region. It uses
∫ x dy and keeps a track of this quantity and adds
C
it up to calculate the area.
Some software applications have been developed that allows the calcula-
tions of areas of regions on maps using this type of principle.
Problems
11.1 If r (t ) = 3t i + t 2 j − 2t 3 k , calculate r ′(t ) and r ′′(t ).
11.3 Suppose the density of a wire is given by d(x,y) = x2. Compute its mass
when the wire is the semicircle of radius 1 as shown in Figure 11.71.
354 Vector Calculus
1
C
x
1 1
4 y = x2
0 x
2
n
5
S3
n
E
S2
4
S1
n y
Chapter 1
3.6V F −G h − v2
1.1 a. A = b. H= c. k = αρc d. t=
L 7 g
1
I 4 sτ H 2E
e. d = 2 α t f. T = g. t= h. v =
εσ A H2 g m
3
3 + 5g U
i. h = j. Q = H
g −1 0.96
3
H (Tj − T0 ) 2
1.2 a. Q = r b. 26649.5 kW
5.38
1.3 a. a = 3 b = –1 b. x = 1.5 y = 0.5 c. x = 0.54 or – 5.54
d. x = 3.16 or – 1.16
Chapter 2
1 2 7
2.1 a. b. c.
4 13 13
5
2.2
14
2.3 a. 0.63 b. 0.97
1 3 3 1
2.4 P(0) = , P(1) = , P(2) = , P(3) =
8 8 8 8
2.5 P(B\D) = 0.62
357
358 Answers
13
2.6 E ( X ) = , σ = 0.68
6
2.7 E(t) = 2.7 years, σ = 0.84 years
1
2.8
407
Chapter 3
1
3.1 uˆ = 〈2, − 1, 4 〉
21
3.2 θ = 129.05°
3.3 α = –2
3.5 −3 i − 5 j − 7 k
3.6 4 i + 9 j + 5k
3.7 –3x + 5y + 2z = –3
3.8 a. 1
Chapter 4
4.1 a. –13 b. –154
6
4.2 k =
5
6 3 6 −2 −1 4
4.3 a. b. −2 4 −9
4 2 5
9 −8 17 14 20
−16 −75
4.4 a. 8 −1 b.
c.
14 13 15
25 34
10 −15 20 15 25
4.5 a. x = 2 y = 3 z = –1 b. x=3y=1z=2
4.6 i1 = 2 i 2 = 2 i3 = 1
−1 −3 1
4.7 λ1 = 1 e1 = 1 , λ2 = 2 e2 = −3 , λ3 = 21 e1 = 1
0 1 0
Answers 359
1 1 1 1 0 3 −1 b. A10 = 4882813 1627604
4.8 a.
6 −3 3 0 5 3 1 14648436 4882813
Chapter 5
1 55 33
5.1 a. –2 ± j b. ± j c. −1 ± j
4 4 3
13 1
5.2 a. 4 + 7j b. –j c. –7 + 11j d. − j
5 5
1 1
5.3 a = 9, b = ; a = , b = 12
3 4
5.4 a. 1.81 + 0.42j, –1.27 + 1.36j, –0.55 – 1.78j
b. 1.21 – 0.17j, 0.54 + 1.09j, –0.87 + 0.85j, –1.08 – 0.57j, 0.21 – 1.20j
Chapter 6
x cos x − sin x
d. e. 15x2(x3 + 1)4
x2
xe 4 x e 4 x 1 x
f. − +C g. e (sin x − cos x )
4 16 2
x n+1 x n+1
h. ln x − +C
n +1 (n + 1)2
360 Answers
6.5 100 J
Chapter 7
7.1 y2 = x2 + 1
x2
7.2 ln( y − 3) = + 5x + C
2
y2
7.3 + y = ln x + C
2
x3
7.4 y = + Cx
2
x5 1
7.5 y = +
5(1 − x ) 1 − x 2
2
x 3
7.6 y = α e − x + β e 2 x − −
2 4
10
7.7 y = α e5 x + β xe5 x +
25
1
7.9 T = (6e5t − 1)
5
Chapter 8
2
8.1
t3
1 10 1
8.2 a. y = − e −2t + e 7t b. T = (1 + e −6t )
9 9 2
c. y = 1 + et d. 1
x= (109 + 18t − e18t )
54
( )
R
− t
V0
R
− t V wLe L V
8.3 a. i = 1− e L b. i = 02 + 2 0 2 2 ( R sin wt − wL cos wt )
R R +w L R +w L
2 2
C kt − kt C kt − kt C
8.4 x = (e − e ) y = (e + e ) −
2k 2k k
Answers 361
Chapter 9
2
0 n = even − n = even
9.1 a. a0 = 1, an = 4 , bn = nπ
− n 2π 2 n = odd 2 n = odd
nπ
0 n = even
b. a0 = 1, an = 0 all n, bn = 2
nπ n = odd
4
− n = even
b. a0 = 0, an = 0, bn = nπ
1
9.2 a. cn = 3 − 5(−1)
n
jnπ 16 n = odd
nπ
Chapter 10
∂f ∂f
10.1 = 3, = 13
∂x ∂y
10.2 a. xy cos xy + sin xy
10.3 − 15
10
10.4 –18
10.5 16
3
Chapter 11
11.2 r (t ) = (1 + t ) i + (1 + 1.5t 2 ) j + (1 − 2t 3 ) k
π
11.3
2
11.4 10
http://taylorandfrancis.com
Index
A D
Algebra, 9–17 Decimal places, 3–4
Alpert’s equation, 16, 41 Delta function, 212–213
Applications, 41–44, 60–63, 78–79, Determinants, 81–87
110–114, 125–129, 164–167, 2 x 2, 83
191–197, 219–227, 250–253, 3 x 3, 85–87
289–295, 339–343, 350–353 Cramer’s rule, 82
definition, 82
B properties, 83–84
Bayes’ theorem, 49–51, 62–63 Differential equations, 171, 172
complementary function, 183
C continuous time Markov process,
Centre of mass, 292–294 193–195
Change of scale, 39 degree, 173–174
Complex numbers, 117 first order, 174–182
addition and subtraction, 118 integrating factor, 179–182
Argand diagram, 120 order, 173
conjugate, 119 particular integral, 188–191
De Moivre’s theorem, 122 second order, 182–191
division, 119 separating variables, 175–179
electrical circuits, 127 smoke layer, 191–193
exponential form, 122 types of, 172–174
forces, 128 Differentiation, 131
imaginary j, 117–118 chain rule, 143–144
multiplication, 119 definition of a limit, 131–136
polar form, 120–122 general formula, 133–134
roots of equations, 123–125 gradient, 131
Conditional probability, 48–49 practical test, 137–138
Conservative vector fields, 310–313 products, 139–140
Continuous random variables, quotients, 140–141
58–60 second derivative test, 138–139
Cosine rule, 25, 27, 28 standard functions, 141–143
Cumulative distribution function, stationary points, 136–137,
56–57 266–267
Curl, 321–327 Dimensional analysis, 113
363
364 Index
Q V
Quadratic equations, 20–22, 117, 118, 184, Variance, 39, 54–55
185 Vector calculus, 297–355
conservative vector field, 310–313
R derivatives of vector functions, 297–299
Radian measure, 30 divergence theorem, 347–350
Range, 36–37 gradient field, 308–316
Relative frequency, 46 integrating vector functions, 299
Reliability theory, 166–167 path independence, 311
Resultant forces, 28 Stokes’ theorem, 343–348
Right angled triangles, 23–25 Vectors, 65
Round-to-even method, 2–3 addition and subtraction, 68
Rounding numbers, 1 definition, 65
flux of vector, 339
S magnetic induction, 79
Scalene triangles, 25–28 magnitude of, 66
Scientific notation, 6–9 normal vectors, 333–338
Significant places, 4–6 projection of, 72
Sine rule, 25–26 scalar product, 69–70
Smoke flow, 41–42, 78, 350–351 tangent planes, 333–336
Solving linear equations vector equation of a line, 74
Gaussian elimination, 104–107 vector equation of planes, 76
inverse matrix method, 108 vector fields, 68, 260, 297, 300,
Standard deviation, 37–39, 42, 54–55, 310–313, 319, 321–327
58–60 vector product, 70
Standard form, 6–9, 180–181, 195 View factors, 291
Statistics, 31–39 Vorticity, 352
Stokes’ theorem, 343–347
Substitution method, 20 Z
Surface integration, 327–336 Zone model, 192, 219