AVery Applied First Course in PDE PDF
AVery Applied First Course in PDE PDF
I /I
f T ar
CIF
TINS
1:
MICHAEL K.. KEANE
A Very Applied First Course
in
Partial Differential Equations
MICHAEL K. KEANE
Department of Mathematical Sciences
United States Air Force Academy
hail
Prentice Hall
Upper Saddle River, New Jersey 07458
Library of Congress Cataloging-in-Publication Data
Keane, Michael, K.
A very applied first course in partial differential equations / Michael K. Keane
p. cm.
ISBN: 0-13-030417-4
1. Differential equations, Partial. I. Title
QA377.K38 2002
515'.353-dc21 2001040032
10 9 8 7 6 5 4 3 2 1
ISBN: 0-13-030417-4
Preface xvii
1 Introduction 1
5 Separation of Variables:
The Homogeneous Problem 131
5.1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
5.2 OPERATORS: LINEAR AND HOMOGENEOUS
EQUATIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
5.2.1 Linear Operators . . . . . . . . . . . . . . . . . . . . . . . . . 132
5.2.2 Linear Equations . . . . . . . . . . . . . . . . . . . . . . . . . 136
5.3 SEPARATION OF VARIABLES: HEAT EQUATION . . . . . . . . 139
5.3.1 Spatial Problem Solution . . . . . . . . . . . . . . . . . . . . 141
5.3.2 Time Problem Solution . . . . . . . . . . . . . . . . . . . . . 144
5.3.3 The Complete Solution . . . . . . . . . . . . . . . . . . . . . 145
5.4 SEPARATION OF VARIABLES: WAVE EQUATION . . . . . . . . 153
5.4.1 Spatial Problem Solution . . . . . . . . . . . . . . . . . . . . 155
5.4.2 Time Problem Solution . . . . . . . . . . . . . . . . . . . . . 156
5.4.3 The Complete Solution . . . . . . . . . . . . . . . . . . . . . 156
5.5 THE MULTIDIMENSIONAL SPATIAL PROBLEM . . . . . . . . . 164
5.5.1 Spatial Problem for X (x) . . . . . . . . . . . . . . . . . . . . 167
5.5.2 Spatial Problem for Y(y) . . . . . . . . . . . . . . . . . . . . 168
5.5.3 Time Problem . . . . . . . . . . . . . . . . . . . . . . . . . . 168
5.5.4 The Complete Solution . . . . . . . . . . . . . . . . . . . . . 168
5.6 LAPLACE'S EQUATION . . . . . . . . . . . . . . . . . . . . . . . . 180
5.6.1 An Electrostatics Derivation of Laplace's Equation . . . . . . 180
5.6.2 Uniqueness of Solution . . . . . . . . . . . . . . . . . . . . . . 181
5.6.3 Laplace's Equation in Cartesian Coordinate System . . . . . 182
Contents vii
11 Fourier Integrals
and Transform Methods 403
11.1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . 403
11.2 THE FOURIER INTEGRAL . . . . . . . . . . . . . . . . . . . . . . 404
11.2.1 Development of the Fourier Integral . . . . . . . . . . . . . . 405
11.2.2 The Fourier Sine and Cosine Integrals . . . . . . . . . . . . . 414
11.3 THE LAPLACE TRANSFORM . . . . . . . . . . . . . . . . . . . . 421
11.3.1 Laplace transform Solution Method of ODEs ..........422
11.3.2 The Error Function ........................424
11.3.3 Laplace Transform Solution Method of PDEs ..........425
11.4 THE FOURIER TRANSFORM . . . . . . . . . . . . . . . . . . . . . 439
11.4.1 Fourier Cosine and Sine Transforms . . . . . . . . . . . . . . 442
11.4.2 Fourier Transform Theorems . . . . . . . . . . . . . . . . . . 444
11.5 FOURIER TRANSFORM SOLUTION METHOD
OF PDES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 450
Appendices 462
Bibliography 501
Index 504
List of Figures
2.1 One-dimensional rod of length L. 6
2.2 One-dimensional rod of length L . . . . . . . . . . . . . . . . . . . . . 10
2.3 Arbitrary slice [a, b] of one-dimensional rod. . . . . . . . . . . . . . . 10
2.4 Temperature distribution of slice [a, b] . . . . . . . . . . . . . . . . . . 12
2.5 Heat flow from end x = L .. . . . . . . . . . . . . . . . . . . . . . . . 18
2.6 Heat flow at end of rod x = 0 .. . . . . . . . . . . . . . . . . . . . . . 19
2.7 The domain of u(x, t) in the rectangle 0 < x < L and 0 < t < T. 24
2.8 Steady-state temperature distribution with u(0) = a and u(L) = b. 32
2.9 Graph of 5 cos x + 20. . . . . . . . . . . . . . . . . . . . . . . . . . . 35
xi
xii List of Figures
4.36 Fourier cosine series representation of 7(x) = x for the partial sum
S5o (x) .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
4.37 Fourier cosine series representation of 7(x) = x) on the interval [-8, 8].124
4.38 f (x) = x on the interval [-2, 2] .. . . . . . . . . . . . . . . . . 124 . . . .
4.39 Fourier sine series representation of 1(x) = x for the partial sum Si(x). 125
4.40 Fourier sine series representation of 1(x) = x for the partial sum
Sio (x) .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
4.41 Fourier sine series representation of f (x) = x for the partial sum
S25 (x) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
4.42 Fourier sine series representation of f (x) = x for the partial sum
S5o (x) .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
4.43 Fourier sine series representation of f (x) = x on the interval [-8,8]. 126
6.1 The graph of the Fourier series representation off (x) . . . . . . . . . 193
6.2 The graph of the Fourier series representation of g(x). . . . . . . . . 193
6.3 The graph of 1(x) = x2. . . . . . . . . . . . . . . . . . . . . . . . . . 213
6.4 The graph of 1(x) = x2 and the partial sum Sl (x). . . . . . . . . . . 214
6.5 The graph of 1(x) = x2 and the partial sum S2 (x). . . . . . . . . . . 214
6.6 The graph of 1(x) = x2 and the partial sum S5 (x) . . . . . . . . . . . 214
6.7 The graph of 1(x) = x2 and the partial sum Slo(x). . . . . . . . . . 214
6.8 The graph of 1(x) = x2 and the partial sum S15 (x) . . . . . . . . . . 215
6.9 The graph of 1(x) = x2 and the partial sum Sioo (x). . . . . . . . . . 215
6.10 The odd periodic extension of 1(x) = 5 on [-L, L]. . . . . . . . . . . 216
6.11 The graph of the odd periodic extension of 1(x) = 5 and the first
partial sum S1 x . 216
6.12 The graph of the odd periodic extension of 1(x) = 5 and the first
partial sum S3 (x) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
xiv List of Figures
6.13 The graph of the odd periodic extension of 1(x) = 5 and the first
partial sum S7 (x) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
6.14 The graph of the odd periodic extension of 1(x) = 5 and the first
partial sum S15 (x) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
6.15 The graph of the odd periodic extension of 1(x) = 5 and the first
partial sum S99 (x) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
6.16 The graph of the odd periodic extension of 1(x) = 5 and the first
partial sum S2oi x . 217
6.17 The graph of 1(x) = x2 + x - 1. . . . . . . . . . . . . . . . 219 . . . . .
2 Ho
8.1 o fl(z) = tonhzand f2(z) =
Graph of L 30 4
z H0-HL)
..............305
(
z (HL
8.2 Graph of fl(z) = tan z and f2(z) = Ho )
z2 HHL + L 0
B.1 u(x, t) in the rectangle 0 < x< L and 0 < t < T . . . . . . . . . . . . 467
xvii
xviii
Course Outline
A possible outline for a one semester course is the following:
Chapters 1 through 8, which is the core material. This provides for the develop-
ment of the three classes of linear second-order partial differential equations, ellip-
tic, parabolic and hyperbolic and the three types of boundary conditions, Dirichlet,
Neumann, and Robin. Additionally, Chapters 1 through 8 gives a thorough discus-
sion of the separation of variables technique, coverage of the relevant theorems of
Fourier series and an introduction to the Sturm-Liouville boundary value problem.
Once Chapters 1 through 8 are covered, there are several options. For a complete
development of classical solution methods of second-order linear partial differen-
tial equations, I would suggest including Chapter 11, which develops the Fourier
and Laplace transforms. For a wider set of applications, I would suggest including
Chapters 9 and 10. It is also possible to chose selected topics from Chapters 9, 10,
and 11 for a broad discussion of applications and technique.
Although the text is not directly tied to a mathematical software package, such
as Mathematica, many of the exercises require the student to find partial sums of
Fourier series. Also, students are required in the exercises to graph both the Fourier
series representation of a function and a three-dimensional view of the solution of
a partial differential equation for various partial sums. Thus, students should be
familiar with some type of mathematical software package.
You may contact the author directly for Mathematica files and other comple-
mentary material related to the text by email at
[email protected]
Acknowledgments
In writing this text, I have drawn from my classroom experience, and I have
been influenced by many sources. I benefited from discussions with colleagues and
students. I thank everyone who has made comments and suggestions. Their interest
in improving the book encouraged me.
It is with great pleasure that I thank George Lobell, my Prentice Hall Editor, for
his support, valuable suggestions, and encouragement. In addition, I would like to
thank Bob Walters, my Production Editor for his assistance and advice and Adam
Lewenberg for his assistance and work on the graphics.
I want to take this opportunity to express my gratitude to the following math-
ematicians who reviewed earlier versions of this text and offered many valuable
XiX
suggestions and insights, which certainly improved the content, readability, and ac-
curacy of the text:
Dr. Daniel C. Biles, Western Kentucky University;
Dr. Christine M. Guenther, Pacific University;
Dr. Jose Barrionuevo, University of South Alabama;
Dr. Mark Kon, Boston University;
Dr. William Margulies, California State University;
Dr. Mikhail Shvartsman, University of Saint Thomas;
Dr. Johnny Henderson, Auburn University.
I would like to express my thanks to the following individuals who gave valuable
support, advice, and suggestions during the development of the text:
Dr. Keith Bergeron, United States Air Force Academy;
Dr. Patricia Egleston, United States Air Force Academy;
Dr. Ali Haghighat, Nuclear Engineering Program University Park;
Colonel and Dr. Daniel W Litwhiler, United States Air Force Academy;
Dr. Glenn E. Sjoden United States Air Force Academy;
Dr. Dawn L. Stewart United State Air Force Academy.
Finally, I want to thank Dr. Michael Round, United States Air Force Academy,
for his time in reviewing several drafts of this text for grammatical accuracy.
I am particularly grateful to my wife, Jean, and to a friend of mine, Sonya, for
their support during this long project. It is with great pleasure that I dedicate this
book to them.
Michael K. Keane
Chapter 1
Introduction
The theory of ordinary differential equations (ODEs) was well established in the
early part of the 18th Century. However, the theory of partial differential equations
(PDEs) was still in its infancy and only studied by a few pioneers. Then, in 1747,
the famous mathematician Jean Le Rond D'Alembert, while studying the problem
of vibrating strings, developed the following form of the wave equation:
a2u(x, t) a2u(x, t)
ate - axe
He also published, in the Memoirs of the Berlin Academy, the solution to Equation
(1.1), i.e.,
u(x,t) = f(x+t)+g(x-t),
where f and g are arbitrary functions.
The field was advanced further by another famous mathematician, Leonhard
Euler. He developed the solution
However, it wasn't until 1782 that P. S. Laplace formalized Equation (1.3) while
working in celestial mechanics. Equation (1.3) is now known as Laplace's equation.
1
2 Chapter 1: Introduction
was studied extensively by J. B. Fourier. In 1822, his paper on the solutions of the
heat equation, Theorie analytique de la chaleur set forth his idea that any function
y = 1(x) could be represented by a trigonometric series, now known as a Fourier
series. The Fourier series is one of the main topics of this text because it provides
a solution of second-order PDEs that are linear and homogeneous with linear and
homogeneous boundary conditions.
Careful inspection of Equations (1.2, 1.3, and 1.4) reveals something that they
have in common: they are all second-order PDEs.
Although the study of second order PDEs is two and a half centuries old, it is
far from obsolete. In fact, you can develop a second order PDE for almost anything
that occurs in nature or is constructed by man. They have been used traditionally
in physics, celestial mechanics, and meteorology. However, second order PDEs are
being developed and applied to problems in economics, mathematical physiology,
nuclear transport theory, aerospace industry, geophysics (particularly in the areas of
lava flow and plate tectonics), car design, electrical engineering, forestry, industrial
and community pollution, oceanography, and a host of other areas. With these
diverse fields researching second-order PDEs, one might get the impression that
there is no common thread to all the different fields. However, this impression is
quite wrong, second-order PDEs can be broken down into three major classes, known
as elliptic, hyperbolic, and parabolic. These names may seem strange; however, they
are rooted in the study of quadratic equations.
A form of the quadratic equation is axe + 2bx + c = 0, which may be solved by
using the quadratic formula
r= -b f a
b2 -ac
the principal part of Equation (1.6). From the principal part of the equation, you
can determine if a linear second-order PDE is elliptic, hyperbolic, or parabolic. For
instance, Laplace's equation, Equation (1.3), has a discriminant equal to -1. Thus,
it is representative of the elliptic class of PDEs. We first encounter this class of
PDE in Chapter 5. The wave equation, Equation (1.1), has a discriminant equal
to 1. Hence, it is representative of the hyperbolic class of PDEs. We encounter
this class of PDE in Chapter 3. Finally, the heat equation, Equation (1.4), has
a discriminant equal to 0. Therefore, it is representative of the parabolic class of
PDEs, which is introduced in Chapter 2.
Laplace's equation and the heat and wave equations are the primary represen-
tatives of their particular classes. However, not all linear second order PDEs are
readily classified. For example, Tricomi's equation,
a2u(x, y) a2u(x, y)
axe + a2
y
is elliptic for y> 0, parabolic for y = 0, and hyperbolic for y < 0. Thus, one must
be very careful when classifying PDEs.
As we study second-order PDEs, it is very important to remember the three
classes, since everything that we learn about the representative of each class of PDE
applies to the entire class. Now, we move on to the study of PDEs by investigating
the heat equation.
EXERCISES 1
1.1. For each of the following linear second order PDEs, identify in what regions
of the two-dimensional plane the equation is elliptic, hyperbolic, or parabolic.
(b)
a2u(x,y) + 2 a2u(x,y) + a2u(x,y) + u ( x' ) 0.
axe y y
a2u(x , y> a2u(x , y) a2u(x , y) au(x , y> _
(c) Sin(xy)
ax 2 - 6 ayax + ay 2 + a y
- °'
82u(x,y) 82u(x,y) 82u(x,y) 8u(x,y) au(x,y)
(d) axe -cos(x)
ay8x + 8y2 + 5y 8x
+5u(x,y)= 0.
Chapter 2
5
6 Chapter 2: One-Dimensional Heat Equation
Let's start by defining heat transfer. Simply put, heat transfer is the way heat is
distributed throughout a material (sometimes steady-state, sometimes time depen-
dent) when either an internal or external heat source is applied. Next, we discuss
how it occurs in a one-dimensional rod.
Physically, every one-dimensional rod has a cross-sectional area, A. Thus, the
one-dimensional rod is actually three-dimensional. See Figure (2.1). A rod is con-
sidered one-dimensional if the heat transfer occurs in a one directional manner,
down the length of the rod. This means the heat transfer across the cross-sectional
area is uniform. Thus, the heat moves in one direction through the length of the
rod. For example, consider a lit match and two different rods. The first rod has the
cross-sectional area equivalent to the cross-sectional area of one human hair; the
second rod has the cross-sectional area of a large steel I-beam. When the end of
the first rod is placed near the flame of the lit match, the end is uniformly heated,
and the heat moves uniformly down the length of the rod. When the end of the
second rod is placed near the flame of the lit match the flame doesn't even cover
the entire cross-sectional area of the rod. Therefore, the cross-sectional area of the
second rod is not uniformly heated, and the flow of heat through the rod is not in
one direction. Now, we'll investigate the three types of heat transfer, starting with
convection.
Convection of heat is the transfer of heat by the actual motion of the material
being heated. For example, if you hold your hand above the surface of boiling
water your hand feels the heat. Some of this heat is from the movement of the air
molecules, initially against the surface of the boiling water. As the air heats above
the boiling water, the air molecules rise, taking heat with them. As these molecules
rise, other cooler air molecules replace them. The rising molecules can even be seen
as steam. If you think of a rod as a thin piece of copper wire, do the molecules of
the wire move away from the wire as it is heated? No. Thus, heat transfer in a
one-dimensional rod is not convection. Having eliminated convection, we proceed
and consider radiation as a possible reason for heat transfer in a rod.
Radiation refers to the continual emission of energy from the surface of all bodies.
For example, hold your hand near the side of a radiator. (Note: do not put your
hand near the top of the radiator, because then heat convection also occurs). You
feel the heat. This is due to heat coming out of the radiator in a wave motion, in
the infrared electromagnetic spectrum similar to electromagnetic waves. In a one-
dimensional rod, heat transfer doesn't move in a wave-like motion down the length
of the rod. Therefore, just as in convection, it seems radiation of heat doesn't
account for heat transfer in the rod.
The last type of heat transfer is conduction. Conduction comes from collisions
of neighboring molecules, transferring heat by kinetic energy. Here, the molecules
move slightly, but do not actually move out of position like they do in convection;
nor do they move like an electromagnetic wave as in radiation. Now their movement
is more like that of tennis balls in a cylindrical container that is a little wider than a
tennis ball, and long enough to hold several tennis balls. Imagine the tennis balls are
molecules. When heat is applied, they can move slightly back and forth, bouncing
against each other and against the side of the can. But they remain essentially in
the same position. This is what happens when heat is applied to one end of a solid
rod. The molecules get excited, but they do not move away from the heat, as in
convection, nor do they move in a wave similar to an electromagnetic wave, as in
radiation; instead, they collide with one another and retain their relative position.
Using these brief explanations of heat transfer, we can see that heat transfer in
a one-dimensional rod is best described by conduction. Why is this so important?
Once we understand that heat transfer in a one-dimensional rod is conduction, we
may use three very important basic properties of conduction, called laws, which are
attributed to Joseph Fourier. These heat conduction laws are as follows:
Law 1. Heat flow is from points of higher temperature to points of lower tempera-
ture.
Law 2. Conduction can only take place in a body when different parts of the body,
including the ends, are at different temperatures.
Law 3. Heat flow changes depending on the material. Different materials, even at
the same temperature have different heat flows.
These three laws form the basis for the mathematical model of heat conduction
in a one-dimensional rod that we'll now develop.
c(x)p(x)
Du(x,t) =
(K0(X)
5u(xt))
at ax ax
-/3(x) [u(x, t) - v(x, t)} + Q(x, t). (2.1)
This is quite an intimidating equation, and its solutions aren't easy to find. Even-
tually, you will be able to solve Equation (2.1) for particular cases. However, we
will simplify Equation (2.1) to develop a careful approach to its solution. First,
we must identify the terms, then make reasonable assumptions about them. This
allows us to derive a new equation. Mathematicians use this method all the time.
Thus, viewing Equation (2.1), we see that there are three double-variable functions
u(x, t), Q(x, t), and v(x, t) and four single variable functions c(x), p(x), Ko(x), and
,3(x).
The first double-variable function, dependent on distance and time, is u (x, t) :
a continuous function that describes the temperature distribution in the rod. This
function depends on distance because the temperature, u(x, t), usually is not the
same at every point x along the length of the rod. It depends on time because
the temperature, u(x, t), is expected to change with time due to some external or
internal source.
The second double-variable function, dependent on distance and time, is Q (x, t) :
a function representing internal heat source, called "source" for short. It describes
the heat energy generated inside the rod. This is an interesting function because
not all materials that can be used to make a rod generate heat; some materials
will only generate heat after an external heat source is applied. For example, a rod
made of steel does not generate heat even when an external heat source is applied,
but a rod composed of a radioactive material generates its own heat, regardless
of external source. Generally, the source function is either known or determined
experimentally.
The last double-variable function is v(x, t), the temperature of the surrounding
medium. It depends on position and time, since the rod could connect different
temperature areas that may change over time. This function plays a big part in
Equation (2.1) if there is no lateral insulation.
Moving into the single-variable functions, we have c(x), the specific heat of the
rod. Specific heat is the heat energy required to raise the temperature of one unit
of material mass one degree. Specific heat is a spatial function, since the material
composition of the rod depends on position. If the rod is composed of only one
material, or the mixture of materials is uniform, the specific heat, c(x), becomes
the constant, c.
The next single-variable function is p(x). It is the mass density of the rod. It
also depends on position in the rod because the rod could be composed of several
materials, each having a different mass density. Mass density is usually measured
as mass per unit volume. As in specific heat, if the rod is composed of one material,
or the mixture of materials is uniform, the mass density, p(x), becomes a constant,
p
Section 2.2: Derivation of Heat Conduction 9
Function Description
u(x, t) temperature distribution of the rod
cp
8u(x, t) =KO 8zu(x, t) + Q(x t).
(2.2)
at axe
Equation (2.2) is the mathematical model for heat conduction in a one-dimensional
uniform rod with perfect lateral insulation.
10 Chapter 2: One-Dimensional Heat Equation
Now, we are going to derive Equation (2.2). Consider the physical model of a
one-dimensional uniform rod with perfect lateral insulation. Apply the conservation
of thermal energy law, which basically states that the rate of change with respect
to time of thermal energy in the rod must equal the rate of flow of thermal energy
across the boundaries, plus the thermal energy produced within the rod. However,
when conservation of thermal energy is applied, the problem must be approached
from a calculus point of view. In other words, consider only a thin slice first and
then expand it mathematically to the full rod.
Shown in Figure (2.2) is a one-dimensional rod oriented in the positive x direc-
tion. That is, the x-axis runs down the center of the rod and the rod has length L,
sothat 0<x<L.
Now, applying the conservation of thermal energy to any arbitrary slice [a, b] of
our rod (see Figure (2.3)) we can claim that
Rate of change with respect to time of total heat inside the slice
= Rate of flow of heat across the boundaries of the slice
+ Total heat generated inside the slice.
By using the previously defined functions, variables, axioms, and some calculus,
the foregoing equation can be expressed mathematically. (Note: In the following
equations, A is the cross-sectional area of the rod).
First,
b
d [fbA(t)d]
We will keep the constants c, p, and A inside the integral for the moment. This
makes the calculations easier.
The second term in Equation (2.3), rate of flow of heat across the boundaries of
the slice [a, b], is somewhat more complicated. Since there is a rate, we know from
differential calculus that a derivative is somehow involved. This derivative must
relate temperature distribution at the boundaries of the slice to heat flow at these
same boundaries. The correct solution lies in understanding heat flow as described
by the three heat Conduction laws:
Law one says heat flow is from points of higher temperature to points of lower
temperature. This law describes how heat flows.
Law two says conduction can only take place in a body when different parts of
the body, including the ends, are at different temperatures. Therefore, Law two
describes when heat flow takes place.
Law three says heat flow changes depending on the material. Different materials,
even at the same temperature, have different heat flows. This law indicates heat
flow depends on material. Hence, we must introduce a function describing the
thermal conductivity of the material. At present our rod is uniform, so the function
describing thermal conductivity becomes a constant.
Now, suppose the temperature in the rod is hotter to the left of the slice [a, b]
than it is to the right of the slice. See Figure (2.4).
By Law 1, this means that heat flow is in the positive x direction. The equation
for heat flow across a cross-sectional area A is
heat flow = _K o
u(x, t) A
(2.5)
ax
12 Chapter 2: One-Dimensional Heat Equation
a Heat Flow b
The constant Ko is the thermal conductivity of the material. The partial derivative,
au(x, t)
, is negative since it indicates the slope of the derivative in the x direction;
ax
therefore, the negative sign in front of Ko must be introduced to show that heat
flow is to the right, the positive x direction. The relationship between the three
heat conduction laws to heat flow, expressed in Equation (2.5), is Fourier's2 law
of heat conduction.
Using Equation (2.5), the slice [a, b] gains thermal energy at the boundary a and
loses thermal energy at the boundary b. Therefore, the second term in Equation
(2.3), rate of flow of heat across the boundaries of the slice [a, b], is mathematically
modeled as
_xoa [au(a, t) _
8x
au(b, t)(2.s)
8x
Again using calculus, we can model the third and final term in Equation (2.3),
total heat generated inside the slice [a, b] per unit time t as
fb
Therefore, utilizing Equations (2.4, 2.6, and 2.7), we can rewrite Equation (2.3) as
t F 6
cpAu(x,t)dx = -KoA raal st) _ aat)l
Lx
J
fb
+ AQ(x, t) dx. (2.8)
Ja
2J. B. Fourier (1768-1830), a very influential mathematician, first postulated the heat conduc-
tion laws, and developed what is known as Fourier's law of heat conduction.
Section 2.2: Derivation of Heat Conduction 13
b p uxt
(s ) dx =- x0 a(x,t)
_aax,t)]
Ia
b
+f a
Q(x,t)dx.
t d
Ia
6
c p u (xt) dx
and
rau(a,t) _ au(b,t)l
-K0
L 8x 8x J
we get
[fbf(x,t) t)
dx.
dt dx Ia b of at,
3 G. W. Leibniz (1646-1716), a contemporary of Sir Isaac Newton, developed the modern no-
tation for calculus in 1676.
14 Chapter 2: One-Dimensional Heat Equation
The first term of Equation (2.9) satisfies Leibniz's formula because u(x, t) is
au (x, t)
continuous and is expected to be continuous. Therefore,
at
dF
dt if
La
b
cp u(x, t)dx =
]
f a
6
cp
au(x,t)
dx.
/'6 p 8u(x, t)
c dx = b Ko a2 a( 2' t) dx + J b Q(x, t) dx. (2.10)
J
f
For equation (2.10) to be true, we must have
b au(x, t) a2u(x, t)
cp - K° - Q(x, t) dx = 0. (2.11)
a at axe
Since the choice of a and b was arbitrary, Equation (2.11) must be true for all choices
of a and b within the length of the rod. Using proof by contradiction, we can show
that Equation (2.11) is true only if the integrand is zero. Thus, we can omit the
integration and consider only the integrand that is,
au(x, t)
at
- K0 a2u(x,
axe
t)
- Q(x, t) = 0
or
au(x, t) a2u(x, t)
p at =K0
K° axe
+Q(x,t). (2.12)
have the initial temperature distribution of the rod, known as the initial condition
(IC) of the rod. Since we are talking about temperature distribution, the IC is a
function of x. It is the condition that must exist before the experiment starts (at
the start of time for the experiment). Therefore, at t = 0, we have u(x, 0) = f (x),
0 < x < L. Remember, even if the IC is a constant c throughout the rod, the IC
may still be thought of as a function of x.
The following equation, Equation (2.14), describes heat conduction in a one-
dimensional rod with the constraints of perfect thermal lateral insulation, constant
mass density, constant thermal conductivity, constant specific heat, no heat source,
and an IC:
au(x, t) _ t)
at axe
(2.14)
EXERCISES 2.2
2.2.1. Suppose you are given a thin slice of a uniform one-dimensional rod with
perfect lateral insulation. State the word equation for conservation of thermal
energy.
2.2.2. Given specific heat, c(x); mass density, p(x); thermal conductivity, Ko (x);
and temperature, u(x, t), state Fourier's law of heat conduction.
2.2.3. Briefly explain the basic idea of Fourier's law of heat conduction.
2.2.4. Suppose we have heat conduction in a rod with perfect lateral insulation, no
internal heat sources, and specific heat, mass density, and thermal conduc-
tivity as functions of x, that is, c(x), p(x), and Ko(x). Starting with the
conservation of thermal energy law, derive a new form of the heat conduction
equation.
2.2.5. Suppose we have heat conduction in a uniform rod with an internal source of
heat energy, but there is no lateral insulation. Thus, the heat flows freely in
and out across the lateral boundary at a rate proportional to the difference
between the temperature, u(x, t), in the rod and the surrounding medium,
/3(x, t). Starting with the conservation of thermal energy law, derive a new
form of the heat conduction equation. Hint: This problem requires a new
formulation of the conservation of thermal energy law.
2.2.6. Suppose you have a perfect laterally insulated uniform rod, but instead of a
heat source it has a heat sink (a sink is where heat is absorbed). Starting with
the conservation of thermal energy law, derive the heat conduction equation.
16 Chapter 2: One-Dimensional Heat Equation
2.2.7. Suppose you have a perfect laterally-insulated uniform rod, but the cross-
sectional area is a function of x (that is, A(x)). Starting with the conservation
of thermal energy law, derive a new form of the heat conduction equation.
au(x,t) a2u(x,t)
at axe
subject to
u(x, 0) = sin x.
au(x, t) 3 a2u(x, t)
at 4 ax 2
subject to
u(x,O) = cos2x.
5e-3x)
2.2.10. Show that u(x t) = 1 + e2t(4e3x + is a solution of
au(x, t) 2 a2u(x, t)
at -9 ax 2
subject to
4e6x + 5
u(x, 0) = 1 +
e3x
2.2.12. Solve u'(t) = atu(t) where a E R. Graph several members of the family of
solution curves.
2.2.13. Solve u' (t) = atu(t) + sin t where a E R. Graph several members of the family
of solution curves.
Section 2.3: Boundary Conditions 17
Also, suppose that the rod's boundary at x = L is held in a different bath of hot
water where the temperature changes in time, but differently than at x = 0. Then
that boundary may be mathematically modeled as
u(L,t) = g2(t).
(Note: gl (t) and g2 (t) describe the temperature of the different baths as they change
with time. Some texts use ubl (t) and ub2 (t) to describe the temperature in different
baths).
Sometimes, the temperatures of baths do not depend on time and are there-
fore constants, Tl and T2. If Tl and T2 equal zero, then we have u(0, t) = 0 and
4 Peter Gustav Lejune Dirichlet (1805-1859) was a Prussian born mathematician who was highly
influenced by Fourier in the early 19th century
18 Chapter 2: One-Dimensional Heat Equation
u(L, t) = 0. In this case, the boundaries being described are in contact with zero-
degree baths, and they are known as homogeneous boundary conditions of the first
kind. Homogeneous boundary conditions are required to develop the separation of
variables solution technique. This technique is one of the major topics covered
in this book.
heat flow is from left to right in the positive x direction. This means the rod is
hotter than the surrounding medium. See Figure (2.5).
From Fourier's law of heat conduction, the equation that models this example
is
au(L, t) _
-KoL
f ) (t), (2.16)
where fi(t) is given. (Note: Ko(L) is the thermal conductivity of the insulation.
Also note, Equation (2.16) describes the value of the derivative at the point L. Thus,
we cannot simply integrate to make it a specified temperature. A very special case
of Equation (2.16) occurs when /(t) = 0. Then Equation (2.16) can be written as
au(L, t) _ 0
ax '
meaning we have the case of perfect insulation. (This may seem impossible, but it
is a standard case to model mathematically).
5 F ranz Neumann, (1798-1895) was a mathematical physicist who worked in Konigsberg.
Section 2.3: Boundary Conditions 19
in Sections (2.2.1 and 2.3), we may now mathematically model a physical situation
involving heat conduction. This is done in the following example.
EXAMPLE 2.1. Mathematically model heat conduction in a one-dimensional
uniform rod of length L with no internal heat source, thermal diffusivity of k, perfect
lateral insulation, and initial condition as a function of x. Also, the left boundary,
at x = 0, is in direct contact with a zero-degree bath, and the left boundary, at
x = L, is perfectly insulated.
8u(L,t) _ o
8x
Notice, both the initial condition and the boundary conditions must be given for
a complete physical description. When both are given, we can generate an accurate
mathematical model of the physical description.
Knowing how to mathematically model a physical situation involving heat con-
duction is important. However, just as important is the capability of giving a
physical description from a mathematical model. This is done in the next example.
EXAMPLE 2.2. Given that
au(x, t) _ a2u(x, t)
cQ(x, t),
at - ax2
subject to the BCs
8u(0, t)
- 6 watts,
8x
(L)
8u(L,tt)
=0
8x
ax
and IC
u(x,0) = f(x),
Solution: A possible physical model for the equations in this example is heat
conduction in a perfect laterally insulated uniform one-dimensional rod of length L,
with thermal diffusivity of k. Also, the source term is a sink, and it is being modi-
fied by a proportionality constant, a. This means heat energy is being withdrawn
from the rod. The left boundary, at x = 0, has a rate of heat flow of a constant
six watts. This may be due to imperfect insulation. The right boundary, at x = L,
is perfectly insulated. Finally, the initial condition is a function of x. Please note,
the initial condition does not have to match the boundary conditions at x = 0 and
x = L.
EXERCISES 2.3
2.3.1. Write a short paragraph (five sentences or less) that describes the physical
problem modeled by the equations
au(x, t) a2u(x, t)
subject to IC
u(x, 0) = -x,
and the BCs
8u(0, t)
_ 0 and u(L, t) = a(t).
8x
2.3.2. Write a short paragraph (five sentences or less) that describes the physical
problem modeled by the equations
au(x, t) _ 1 a2u(x, t) _t
at 4 ax2 - 15[ u(x, t -
) (8x - 2e + 2 )],
subject to IC
u(x, 0) = 8x,
2.3.3. Suppose a nonuniform metal rod of length it, with perfect lateral insulation,
has an initial temperature distribution of sin x. Initially, one end of the rod
is fixed at a temperature of 0°C, while the rest of the rod is placed in liquid
nitrogen. What would be a possible mathematical model that describes this
problem? Explain your answer.
22 Chapter 2: One-Dimensional Heat Equation
2.3.4. Consider a one-dimensional rod of length L. Assume that heat energy is flow-
ing into the rod at x = 0 proportional to the temperature difference between
the end temperature of the rod and the known external temperature. Develop
the mathematical model for this condition. Briefly justify your answer.
2.3.5. Consider a one-dimensional rod of length L. Assume heat energy is flowing
into the rod at x = L proportional to the temperature difference between the
end temperature of the rod and the known external temperature. Develop the
mathematical model for this condition. Briefly justify your answer.
2.3.6. Given Equation (2.17), show that if h - * 0, then we get Neumann's condition
of perfect insulation. If h - * oo, then we get Dirichlet's condition.
2.3.7. Consider a one-dimensional rod of length H. Assume the rod has no lateral
insulation, is nonuniform, and an internal heat source doesn't exist. Also,
assume at the boundary x = 0, the rod is held at a constant temperature of
15°C, and at the boundary x = H, the rod is imperfectly insulated, allowing a
heat energy flow of a constant -8 watts. Develop the mathematical model that
includes a possible initial temperature-distribution equation. Briefly explain
your choice of initial temperature-distribution equation.
2.3.8. Show that u(x, t) = e-2t sin x is a solution of
au(x, t) a2u(x, t)
at = 2 axe '
subject to the BCs
u(0, t) = 0 and u(27r, t) = 0,
and IC
u(x, 0) = sin x.
u(x,0) = cos3x.
a2u(x, y) _
2.3.10. Solve
axay xy.
Section 2.4: The Maximum Principle 23
Exercises (2.3.11 through 2.3.15) involve first-order ODES. A review of this material
may be found in Appendix C.
2.3.11. Carbon 14 obeys the law of radioactive decay. Determine k if the half-life of
carbon 14 is 5568 years. Next, suppose the initial amount of carbon 14 is 9
gm. Determine the amount of carbon 14 left after 256 years.,
2.3.13. Suppose in a population of yeast cells, growing exponentially, the initial pop-
ulation of cells is 1200. Fifteen minutes later it is 1700. Find the growth rate
for the population.
2.3.14. A cup of coffee is initially at boiling point, 100°C. The temperature of the
room is 20°C. Find the temperature of the coffee as a function of time. (Hint:
use Newton's law of cooling.)
2.3.15. In a furnace, the temperature of the inner wall of an area 2m2 is 450°C. The
temperature of the outer wall is 80° C. There is a 0.5m of brick insulation
(thermal conductivity of brick is 0.38) between the walls. How much heat
escapes in three minutes? (Hint: Assume steady state has been reached
across the walls and remember Fourier's law of heat conduction:
du(x) A
heat flow = -K
° dx )
au(x, t) _ ka2u(x, t)
(2.1s)
at axe
24 Chapter 2: One-Dimensional Heat Equation
and to IC
u(x, 0) = 1(x). (2.20)
We want to determine if a unique solution exists. Our intuition tells us that there
is a unique solution. However, proving it is a slightly different matter. We start by
stating a maximum-minimum theorem for diffusion of heat in a one-dimensional
rod. Then we state and prove an immediate result of the maximum-minimum
theorem, which shows that the problem given in Equations (2.18, 2.19, and 2.20)
has an unique solution. Finally, we state another result of the maximum-minimum
theorem, that of continuous dependence of the solution on the initial data. Thus,
the problem is well-posed.
0 L
Figure 2.7: The domain of u(x, t) in the rectangle 0 < x < L and 0 < t < T.
au(L,t)
(L
=ht
ax ()'
and to IC
u(x,0) = f(x),
has a unique solution. However, that method is beyond the scope of this text.
Another result of the maximum-minimum theorem is continuous dependence of
the solution on the data (both the boundary conditions and the initial conditions).
Corollary 5. (Continuous dependence of the solution on the initial data) The so-
lution of the problem given in Equations (2.18, 2.19, and 2.20) depends continu-
ously on the initial and boundary conditions in the following way: let u1 (x, t) and
26 Chapter 2: One-Dimensional Heat Equation
u2 (x, t) be solutions to the problem with initial data gl (t), hl (t), and Ii (x) and g2 (t),
h2 (t), and f 2 (x), respectively. Let T and s be any positive real numbers. If
Om I- .f2(x)I C E+
om I91 (t) - 9a (t) I e, and
The heat equation in polar, cylindrical, and spherical coordinate systems are given
by Equations (2.23, 2.24, and 2.25) respectively.
au (r,O,t) _ k1 1 8 rr8u(r, B, t) 1 D2u(r, e, t)
ae2 (2.23)
at r 8r L 8r +r2
a2u(r, e, z, t)
(2.24)
uz2
1 a2u
(2.25)
r2 sin2 0
Section 2.4: The Maximum Principle 27
The right side of Equations (2.21, 2.22, 2.23, 2.24, and 2.25) is called the Laplacian,
and it is given in many equations as V2u where the arguments of u are understood
from the nature of the problem.
We will state and use Equations (2.21, 2.22, 2.23, 2.24, and 2.25) in later chapters
in the text. However, you should know how to derive each of the equations from
the Cartesian counterpart. For instance, the heat equation in polar coordinates,
Equation (2.23), may be derived from the Cartesian two spatial dimension heat
equation, Equation (2.21). A demonstration of the derivation is given in Example
(2.3
of the heat equation from the two spatial dimension Cartesian form,
ar _ x _ a2r _ y2 _ sine 8
ax - cos 8, which implies ax2
/x2 + y2 (x2 + y2) 3 r
Similarly,
Dr a2 r cost 6
ay
= sin 8, which implies
a y2 = r
Also,
and
We are now ready to find the polar coordinate form of the heat equation. Using
the chain rule, we have
au Du Dr Du DO
ax = Dr ax + ae ax'
28 Chapter 2: One-Dimensional Heat Equation
which means
D2 u _ 3 Du Dr Du DO (D2u Dr D2 u DO Dr Du D2 r
Dx2 Dx Dr ax + 093x Dr2 d x+ DBr d x d x+ Dr Dx2
D2 u DO D2 u Dr DO Du D20
+ Dal dx + are dx dx + de Dx2
D2 u 2 D2 u 2 sin 8 cos 8 D2 u sing 8 Du sing 8 Du 2 cos 8 sin 8
cos 0- r
Dr2 d r9 + 382 r2 + dr r + 38 r2 .
Similarly,
32 u D2 u 2 D2 u 2 sin 8 cos 8 D2 u cost 8 Du cost 8 Du 2 cos 8 sin 8
32 sin 0+
Dr2 are r + 382 r2 + dr r De r2
Thus,
D2 u D2 u D2 u D2 u 2 sin 8 cos 8 D2 u sing 8 Du sing 8
cos2 0-
3x2 + 5 2 5r2 are r + 382 r2 + 3r r
Du 2 cos 8 sin 8 D2 u 2 D2 u 2 sin 8 cos 8
+ r2 +D-2 sin e +
38 are r
+- 32 u cost 8
502 r2
+ Du cost 8 - Du 2 cos 8 sin 8 _ D2 u
dr r de r2 d r2
.+' _
1 Du
rdr
.+'
1 D2 u
--
r2 382
_ 13 du 1 32u
r
r Dr Sr + r2 392
.
Hence, we have
Du(r, 8, t) _ 1D Du,r(rdr8, t) 1 D2u(r, 8, t)
St \r 3r + r2 382 .
In the exercises, you must derive the cylindrical and spherical form of the Lapla-
cian.
EXERCISES 2.4
72vi(r yi
'
- - - -L
D2u
Dx2
D2u
Dy2
D2u
Dz2 .
Section 2.4: The Maximum Principle 29
(2) If x = r sin 8 cos , y = r sin 8 sin , and z = r cos 8, show that the
Laplacian can be written in the spherical coordinate system as
1a au a2u
V2u('r 8 ,q5) =
1
r2 sin 8
r2 ar ar + r2 sin 8 a82
1
a2 u
+
r2 sin2 a
(a) x = cos b.
(b) py = sin .
8¢ _ -sin q5
8x p
dJ 8¢ _ cosh.
ay p
(2) Given r = p cos i + p sin b7 + zk. Show that the square of the element
of arc length is
(ds)2 = drdr = (dp)2 + (dz)2.
y = p sin b, and
z=z;
p2 = x2 + y2 and q5 = tan-1(a
), prove that the cylindrical coordinate system
x
is orthogonal.
30 Chapter 2: One-Dimensional Heat Equation
x = p sin 8 cos b,
z = p cos 8;
p2 = x2 -+
y2 + z2 and b = tan-1
(v), show that
x
ap
(a) ax = sin 8
a
(b) p = sin 8 sin b.
Dy
ap
(c) = cos 8.
az
sin b
(d) Dx p sin 8
(e)-=
Dy
Dqcos b
psin 8
(2) Given r = p sin 0 cos p sin 0 sin bj + p cos 8k, show that the square of
the element of arc length is
y = aQ, and
z=z;
show that the square of the element of arc length is
Now, suppose g(t) and h(t) are the fixed constants a and b, respectively. Then,
the solution to this model will simplify to a steady-state temperature distribution,
independent of time t. This solution is known as the equilibrium solution.
To understand what we mean by steady-state, think of a uniform rod with perfect
lateral insulation and no internal sources. Now, apply a constant temperature
forever to both ends. Since neither end is insulated, eventually the temperature in
the rod will adjust to the temperature distribution specified by the heat sources at
both ends.
Definition 6. A steady-state temperature distribution is a temperature distribu-
tion that does not depend on time.
t)
From Definition (6), u(x, t) from Equation (2.26) becomes u(x). Thus,
au(x,
at
Du(t) 32
= 0 and k
8(xz t) k d dx ax) - 0 or
d2u(x)
= 0, (2.29)
dx2
with BCs
u(x) =
b-a (2.32)
0 L
Du(x, t) __ D2u(x, t)
Dt +
Q(x).
Dx2
Would the solution, Equation (2.32), change? How would this effect Figure (2.8)?
These questions are important to answer for a fuller understanding of this steady-
state problem; they'll be asked again in the exercises at the end of the section.
Next, suppose we want to find the steady-state temperature distribution in a
uniform one-dimensional rod with no internal source, perfect lateral insulation, and
perfect insulation on the boundaries. We again start with Equation (2.29)
d2u(x)
dx2 = 0.
However, the B Cs change to
du(0) _ du(L)
dx dx
This implies u(x) = C2. It appears that steady-state temperature in a uniform rod,
with perfect insulation everywhere, is an arbitrary constant. But is it arbitrary? No.
Section 2.5: Steady-State 33
In this case, it turns out that the initial condition of the rod, u(x, 0) = 1(x), plays
an important part. The energy associated with the initial temperature distribution
cannot escape because of the perfect insulation. Since temperature distribution in
the rod must follow the three heat conduction axioms, the temperature distribution
levels out to a constant. But what constant? This question can only be answered
by going back to Equation (2.8), which is the mathematical representation of con-
servation of thermal energy. It states that for any thin slice of rod [a, b],
d
1b rau(a, t) _ au(b, t) 1
cpAu(x, t)dx = -K°A
dt L 8x 8x J
This equation is valid for the entire rod, not just the thin slice [a, b]; therefore, it
can be written as
L
+ AQ(x, t)dx. (2.33)
0
Because there is no internal source, Q (x, t) = 0. Also, we have a uniform rod, which
means c p, and Ko are constant and can be written as k = K° Thus, Equation .
cp
(2.33) may be written as
d 8u(0, t) 8u(L, t)
u(xt)dx] = - (2.34)
LLL 8x 8x
We can further reduce Equation (2.34) by remembering that both ends are perfectly
au(0, t) = 0 and au(L, t) = 0.
insulated, which means Thus Equation (2.34)
' ax ax
becomes
L
This tells us that the total initial heat energy inside the rod must equal the total final
heat energy inside the rod. We did not say that the initial temperature distribution
is the same as the final temperature distribution. Since u(x, 0) = 1(x), the total
initial heat energy is
Also, since u(x) = C2, the total steady-state heat energy in the rod (the total final
heat energy) is
L
f C2 dx = C2 L.
Setting total initial heat energy equal to total final heat energy and solving for C2,
we arrive at
fL
C2 = L f (x) dx,
J0
which is the average of the initial temperature distribution.
As a final example, consider the following problem:
EXAMPLE 2.4. Consider a one-dimensional uniform rod with perfect lateral in-
sulation, internal heat source of 25 cos x, thermal diffusivity of 5, and initial temper-
ature distribution of 25 cos x. At the boundary x = 0, the rod is in direct contact
with a bath held at the constant temperature of 25°C, and at the boundary x = it,
the rod has perfect insulation. Determine and graph the steady-state temperature
distribution.
Solution: First, describe the above physical problem as a mathematical model.
We have
Du(x, t) a2u(x, t)
at - 5
axe
+ 25 cos x,
subject to the IC
u(x, 0) = 25 cos x
and BCs
u(O,t) = 25
8x
Second, state and solve the steady-state problem. The steady-state problem is
d2u(x)
5 + 25 cos x = 0,
dx2
subject to BCs
u(0) = 25
du('ir)
l = 0.
dx
Section 2.5: Steady-State 35
u(x)
25
15
x
0
it
EXERCISES 2.5
2.5.1. Given aone-dimensional uniform rod with perfect lateral insulation, determine
and graph the steady-state heat distribution given the following boundary
conditions, source, and thermal diffusivity:
(1) u(0) = 0, u(L) = a, Q(x) = 0, k = 2.
u(O,t) = 1 and -1
and IC
u(x, 0) = cos x - x.
(1) Write a short paragraph (five sentences or less) describing a possible
physical model.
(2) Determine the steady-state temperature distribution.
2.5.3. Determine the steady-state solution for a uniform rod with perfect lateral in-
sulation if the boundary at x = 0 is kept at a constant temperature of -10°C,
the boundary at x = 100 cm is kept at a constant temperature of 15°C, and
there exists a time-independent heat source, which is a linear function based
on the position in the rod.
2.5.4. Find the steady-state solution for a uniform one-dimensional rod with no inter-
nal source and perfect lateral insulation that satisfies the radiation condition
Du(0t)
- u(0, t) = 0
Dx
at the end x = 0 and is kept at a constant temperature T2 at the end x = L.
2.5.5. Find the steady-state solution for a uniform one-dimensional rod with an
internal source of Q(x, t) = x, perfect lateral insulation, that is kept at a
constant temperature Tl at the end x = 0, and satisfies the radiation condition
Du(L, t)
Dx
-u(L,t)=0
at theendx=L.
2.5.6. Consider the mathematical model
Du(x, t) D2u(x, t) tee_x
(cos ?Cx) =e + t2
Dt 3x2 1'
subject to the BCs
u(o,t)= o and a i
and IC
u(x,0) = x2.
Section 2.5: Steady-State 37
2.5.7. Consider
au(x, t) _ a2u(x, t)
8t 8x2
+x-i3,
subject to the BCs
as°'t)=0andaax't) =o
and IC
u(x, 0) = cos
x
(1) Find the equilibrium temperature distribution.
(2) For what values of 3 does the equilibrium temperature distribution exist?
Explain physically.
Du(x, t) __ a2u(x, t)
9 +9x,
Dt Dx2
and IC
and IC
u(x, 0) = 7rsinx.
2.5.11. Consider a uniform one-dimensional rod with no internal source and perfect
lateral insulation. Assume that u(x, 0) = 1(x), and suppose the following
boundary conditions are given:
Du(s t) __ Du(-7r,t)
u(7r, t) = u(-7r, t) and
Dx Dx
(1) State a possible physical explanation. Briefly justify your answer.
(2) Does a steady-state solution exist? If so, what is the steady state so-
lution? Explain your answer. If a steady-state solution does not exist,
explain why.
2.5.12. Given the Cauchy-Riemann equations
Du - Dv _0
Dx ay
Du Dv °
Dy+Dx= '
show that both u and v satisfy Laplace's equation
D2z D2z
Dx2 + Dy 2 = °'
which may be considered as the multi-dimensional steady-state equation.
Laplace's equation is discussed in Chapter 5.
2.5.13. Fick's law of diffusion for chemical species in mathematical physiology is given
by
Q = -DVc,
where Q is the flow of a chemical across a membrane, D is the diffusion
coefficient and is dependent on the solute and the fluid in which the chemical is
dissolved, and c is the amount of chemical. Fick's law can be used to derive an
analogue of Ohm's law for a membrane of thickness, L, with different chemical
concentrations on each side of the membrane. If the medium is isotropic
(diffusion occurs the same regardless of the direction of the measurement),
then we get
Dc D2 c
at = D Dx2 '
subject to
c(0, t) = Cl, the chemical concentration on the left of the membrane,
and
c(L, t) = Cr, the chemical concentration on the right of the membrane.
Find the steady-state solution.?
7James Keener and James Sneyd, Mathematical Physiology, ©1998 by Springer-Verlag, New
York, pp. 36-38. Reprinted by permission.
Chapter 3
/a atD2u at\
+b
axe + x ax - 2 - °'
where u(x, t) is the displacement of a diaphragm in a capacitor microphone. Other
physical examples are the electromagnetic waves in a transmission line, wave motion
in an ocean, and vibrations in a beam. In later chapters, we will develop equations
and solutions for some of these physical examples. However in this chapter, we
consider the most basic wave equation: vibration of a one-dimensional string.
We derive the wave equation for vibration in a one-dimensional string in Sec-
tion (3.2). Boundary conditions (BCs) for the one-dimensional wave equation are
discussed in Section (3.3). Section (3.4) covers conservation of energy for the wave
equation and uniqueness of solution when using Neumann boundary conditions. We
conclude this chapter with the method of characteristics for first-order PDEs and
d'Alembert'sl solution for the one-dimensional wave equation.
41
42 Chapter 3: One-Dimensional Wave Equation
a string that is not tightly stretched. Gravity could play a completely different
role in a string that is vertical versus a string that is horizontal. In this section,
we restrict ourselves to a tightly stretched horizontal string. Also, we make the
following assumptions:
Vibrations of the string are small. This means that as the string vibrates, we
have a very small change in the slope of the string from the "at rest" position,
which is horizontal.
Only vertical vibrations will be considered horizontal vibrations can be ne-
glected because of the small slope.
Vertical displacement then depends on the position on the string, along with
time, and can be modeled as y = u(x, t). Thus, the slope is represented by
dy _ au
tan[B(x, t)].
dx ax =
The string is perfectly flexible. It offers no resistance to bending.
Newton's law of motion, F = ma (which we will write as ma = F), is applied
to a small section of the string (x to x + Ox).
Consider Figure (3.1), which is out of proportion for labeling purposes:
'L(x + &, t)
8(x+&, t)
x+&
Here T, a tangential force, represents the tension in the string, and B is the angle
of the displacement of the string from the horizontal at the ends x and x + Ox.
It is unknown whether the string is uniform. Therefore, we represent the mass of
the string as a function. However, we only expect the mass of the string to change
with position, x, not with time, t. Thus, we chose p(x) to model mass, m, per unit
length and assume it is a known quantity.
In addition to mass, we must consider other forces on the string. One example
is the restoring force. This is the force that tries to return the string to its at-
rest position. This force models the added tension on the string caused by vertical
Section 3.2: Derivation of the One Dimensional Wave Equation 43
D2 u
Mass times acceleration is represented by the term (/2x)p(x) . The right side
Dt2
of Equation (3.1) is the sum of forces. We have two terms for tension: one for the
left end at x and one for the right end at x + Ox. Since only vertical vibrations are
considered, the terms r(x + Ox, t) sin [9(x + Ox, t)] and r(x, t) sin [9(x, t)] are the
vertical components of the tensile force. Assuming that a > 0, the term (L2x)c-
models the resistance force. (Ox)/3u with j3> 0 models the restoring force. Other
possible external forces, such as gravity, are modeled as (x)Q(x, t).
Dividing Equation (3.1) by Ox and taking the limit as Ox - 0, we get
D2u Du
p(x)
Dt2
D
= x ((x, t) sin [9(x, t)]) - a
a a
t - u + Q(x, t). (3.2)
dy
dx
- au
8x
sin [e(x, t)].
a Du
D2 u
p(x) Dt2 = Dx T(x, t)
Dx
- a Du
Dt
- /% + Q(x, t). (3.4)
Equation (3.4) is the mathematical model for small vibrations in a small piece of
perfectly flexible string that is horizontal and tightly stretched.
If we make the assumption that the string is perfectly elastic (a valid assumption
for most strings), then the tensile force r(x, t) may be approximated by the constant
44 Chapter 3: One-Dimensional Wave Equation
T, which is the initial tension on the unperturbed string. Also, if we assume the
string is made of a uniform material, then the mass density p(x) becomes the
constant p and Equation (3.4) becomes
82u 82u 8u Nqu
p 8t2 - T 8x2 - a 8t - + Q(x, t).
Another assumption we can make is that the term Q (x, t) only models gravity.
If we know the tension force is high compared to the force of gravity, we may neglect
gravity. Doing so yields
012 012
u u au
pate = T cat - ,3u'
axe -
which is a form of the well-known Telegrapher's equation.
As its name implies, the Telegrapher's equation models electromagnetic wave
transmission in a wire. Also, it is an important equation in analyzing the time-
dependent Boltzmann's equation in the theory of neutron transport. We discuss
solutions to the Telegrapher's equation in a later chapter.
To simplify Equation (3.5) we may, for the moment, neglect the forces of friction
and restoration. Mathematically, this means we assume that a and j3 are zero. Thus,
the mathematical model for vibrations in a one-dimensional, uniform, perfectly
flexible, highly stretched string is
012 012
p=u TT axeu
or in its more usual form
012
u 2 Du
u
ate
T= c ax2'
where c2 = T . Because T is tension and p is mass per unit length, c has the
p
dimension of length/time, known as velocity. Actually, c is the specific velocity of
wave propagation along the string.
The one-dimensional wave equation applies to many different physical systems.
For example,
012 012
u u
at2 01x2
and
Dv L Di
Ri = 0
ax + at +
where the variables have the following meaning:
R is the resistance.
L is the self-inductance.
Using Equations (3.8 and 3.9), we can derive a form of the wave equation by
differentiating Equation (3.8) with respect to x and differentiating Equation (3.9)
with respect to t. This yields
D2 i D2v Dv
(3.10)
axe + cDxat + GDx = °
and
D2v D2 i Di
+ L at2 + R =0. (3.11)
Dtax Dt
Dv
cR Di =0.
D2 i D2 i
+G - CL 5t2 - (3.13)
axe Dx Dt
46 Chapter 3: One-Dimensional Wave Equation
av
Now, using Equation (3.9) in the form = -L a2 - Ri, we can rewrite Equation
Dx Dt
(3.13) as
D2i Di D2i Di
axe - GL Dt - cL Dt2 - cRDt -GRi = o
or in the more familiar wave form,
D2 i D2 i Di
CL = - (GL + CR) - GRi. (3.14)
a t2 a x2 at
In a similar fashion, we can derive a wave equation for the potential:
D2CLv D2v Dv
at 2
=
ax
2 - (GL + CR) t - GRv.
a
(3.15)
8u(x, 0)
l 8t - 9(x)
EXERCISES 3.2
/3u, and
Q(x, t).
3.2.2. Show that for all positive integers m, each of the following functions satisfies
82u
ate
=c 2 82u
ax2
(1) ul(x, t) = sin[mirx] sin[mirct].
(2) u2(x, t) = sin[mirx] cos[mirct].
ai + C av
Gv = 0
ax at +
av - L a2
+ Ri = 0.
ax at
CL t2
2 2
x2 - (GL + CR) t - GRv.
48 Chapter 3: One-Dimensional Wave Equation
au au a3 u
at
+6u-
Dx + ax3
.
where u = u(x, t) is the velocity of the gas and p = p(x, t) is the density of
the gas. Show that u and p satisfy the wave equation.
3.2.10. Solve
subject to ICs
u(0)=0
u'(O) = 0.
Figure 3.2: String attached to elastic ends with their ends fixed.
The spring constant is assumed to be positive. Also, we will assume the spring
constant is the same at both ends, and we will denote it as k.
The mathematical model for this physical condition is
au(0t)
T = ku(0 t
ax )
Note: The spring constant may not be the same at both ends. This condition would
indicate different springs attached to either end of the string.
Another example of Robin's conditions is shown in Figure (3.3). Here, the end
Figure 3.3: String attached to elastic ends with their ends displaced.
x = 0 is attached to a spring where the end of the spring can move in a vertical
52 Chapter 3: One-Dimensional Wave Equation
direction. The displacement of the end of the string is described by u (O, t). The
displacement of the left end of the spring can be described as d(t). The tension on
au(0, t)
the end of the spring at its attachment to the string is described by T By
ax
setting the vertical tension of the spring equal to the difference of the displacements,
we arrive at the mathematical model
Du(O, t)
______ = k[u(O,t) - d(t)]. (3.18)
Assuming the spring obeys Hooke's3 law, k is the spring constant. The signs in
Equation (3.18) are identical to Newton's law of cooling, which was developed in
Chapter 1. The end x = L has a similar mathematical model, and it is left as an
exercise.
EXERCISES 3.3
3.3.1. Explain in your own words what happens to the Robin's condition
5u(O,t)
= h[u(O,t) - 'yi(t)]
when
(1) h -f oc.
(2) h -f 0.
3.3.2. State the mathematical model for the given information:
(a) Small vertical vibrations in a uniform tightly-stretched string.
(b) Fixed left end at 4.
(c) Free right end.
(d) String length of 37r.
(e) Tension of 4.
(f) Mass density of 3.
(g) Initial displacement of x2 + 2x + 4.
(h) Initial velocity of 0.
3.3.3. State the mathematical model for the given information:
(a) Small vertical vibrations in a nonuniform tightly-stretched string.
(b) Free left end.
3R. Hooke (1638-1703) was professor of geometry at Gresham College and secretary of the
Royal Society.
Section 3.3: Boundary Conditions 53
au(x, 0)
_ In (x + 1) - 2.60759.
at
3.3.6. Given the following equation, describe the physical situation:
ate = a
a2u au
(x + 1) 2
ax
(3x+5)-)
ax
+ sin xt,
subject to BCs:
u(0, t) = o
8u(L, t)
_ 3 cost
8x
and ICS:
u(x,0) =ln(x+1)
Su(x,0)
at
-0
54 Chapter 3: One-Dimensional Wave Equation
3.3.7. Develop the mathematical model for Robin's conditions at the boundary x =
L described in Figure (3.3).
3.3.8. A uniform string with fixed ends has an initial displacement of 2x for 0 < x <
r and 37r - x for 7r < x < 37r. It is known that the initial velocity is 0 and
the string is vibrating in a medium that resists the vibrations. (The medium
produces a resistance proportional to the velocity.) Suppose the resistance
constant of proportionality is 0.01. State the mathematical model.
3.3.9. A uniform string with a fixed end at 0 and free end at 27r has an initial
displacement of -x for 0 < x < and 3x - 67r for x < 27r. It is
- 2 2 - -
known that the initial velocity is 0 and the string is vibrating in a medium
that resists the vibrations. (The medium produces a resistance proportional
to the velocity.) Suppose the resistance constant of proportionality is 0.03.
State the mathematical model.
,0<x<1
u(x,0)= x-,1x<3
-2x+8,3<x<4.
It is known that the initial velocity is 0 and the string is vibrating in a medium
that resists the vibrations. (The medium produces a resistance proportional
to the velocity.) Suppose the resistance constant of proportionality is 0.13.
State the mathematical model.
of the mass point at any time t. Also, potential energy is given by E = lkx2 ,
where k is a force constant of proportionality and x is the coordinate of the body.
Consider the equations for a uniform tightly stretched vibrating string with free
ends and initial conditions,
52 u 252u
c (3.19)
ate = ax2 '
Section 3.4: Conservation of Energy for a Vibrating String 55
subject to BCs
(3.20)
and ICs
(3.21)
8u(x, 0)
8t -
Here, the kinetic energy is
EP = 2 fL fa"(x't)Jl2dx.
(3.23)
Equations (3.22 and 3.23) are different than the single-point mass equations because,
for a string, we must sum over the entire length. Thus, we have
E=Ek+E
= 2 L (5u(x,t)2d x + 2 J L (5u(x,t)2dx. (3.24)
J0 0
1
We need the sum of the instantaneous kinetic and potential energy, which means
taking the derivative with respect to time, t. Hence, the equation
dE_ d [1 t)1 2
+ c2
ty12
dx (3.25)
dt 2 ,f L l
Using Leibniz's formula on Equation (3.25) yields
2
dE La au(x, t)2 2a 1' 't)
dt - 2, at at
+
at ax ) dx
56 Chapter 3: One-Dimensional Wave Equation
which reduces to
dE 2 La au(x, t)au(x, t)
dt CL 8x [ at ax ] dx
which means the sum of the instantaneous kinetic and potential energy is equal to
a constant.
We may now proceed and prove that if Equations (3.19, 3.20, and 3.21) have a
solution, then the solution is unique.
Proof. Let ul (x, t) and u2 (x, t) be solutions to the given equations. Then, v(x, t) _
av(0, t) _ aul (0, t) au2 (0, t) av(L, t) _
ul (x, t) - u2 (x, t) . Thus, = 0 and
ax ax ax ax
_
Dui(L,t) au2(L, t) _ av(x, 0) _
ax ax -
0. Also, v(x, 0) = ul (x, 0) - u2 (x, 0) = 0 and
at
aul (x, o) _au2 (x, o) _
at at -
0. Thus, the sum of the instantaneous kinetic and potential
energies of v(x, t) must equal a constant. We form the energy equation
L
1 c2o L (av(xt)\2
dx dx. (3.28)
E- 2o at
(av(x,t)\2 + 2 ax
Taking the derivative with respect to t yields
dE
2 dt f oL
t)12
J
dx + 2 L (av(xt)
J o
a 2 dx
'
(3.29)
Section 3.5: Method of Characteristics 57
which becomes
dE _ rav(x, t)av(x, t) 1 L
dt L at ax 0
This means that E(t) = 0, which in turn implies that the integrands of Equation
Dv (x, t) (x, t)
(3.28) are zero. Thus and must be identicallyY zero which implies
cat Dx
v(x, t) - 0. Thus, ul (x, t) = u2 (x, t). D
EXERCISES 3.4
3.4.1. Given a uniform tightly stretched vibrating string with a fixed end at x = 0,
a free end at x = L, and initial conditions, determine what happens to the
total energy E.
3.4.2. at the end x = L, and the springs other end fixed, and initial conditions,
determine what happens to the total energy E.
Note: The spring constant is assumed to be a positive.
3.4.3. Given the equations for a uniform tightly stretched vibrating string with ho-
mogeneous fixed ends and initial conditions,
a2 u 232u
=c axe' (3.30)
ate
subject to BCs:
u(0, t) = o
(3.31)
{
u(L, t) = 0
and ICs:
u(x,0)=f(x)
(3.32)
Du(x,0)
a:
at 9(x).
Show that if this problem has a solution, then the solution is unique.
where -oc <x < oc and 0 <t < oc. Unlike the second-order wave equation, which
has two initial conditions, a first-order PDE will only have one initial condition.
This initial condition describes the initial position at time t = 0. Remember, initial
conditions depend on the number of partial derivatives with respect to time that
are in the equation. A first-order PDE only has one partial derivative with respect
to time. Also, there are no boundary conditions because the variable, x, varies from
negative infinity to positive infinity.
One way to solve this problem is to consider the rate of change of z(x(t), t) as
measured by a moving observer, x = x(t). The chain rule implies
tz(x(t),t) - 5zdx 5z
ax dt + at'
(3.35)
Comparing these quantities with equation (3.33), we see that if the observer moves
dx
along at velocity -c then = -c and z(x(t)' t) = 0. That is, z is a constant.
dt dt
Dz
The term in Equation (3.35) represents the change in z with respect to time
Dt
Dz dx
at the fixed position x. The first term in Equation (3.35), represents the
Dx dt '
change that the observer sees as the observer moves into different regions of the
domain of z.
Using Equation (3.35), we have reduced the PDE in Equation (3.33) to two
dx dz
first-order ordinary differential equations: = -c and = 0. If we consider
dt dt
that
dx _
dt
-c '
we see that the solution is
where a represents the initial point on the characteristic curve when t = 0. Using
the IC (3.34), we have z(x(t), t) = z(a, 0) = f (a).
Section 3.5: Method of Characteristics 59
z(x, 0) = sin x,
which implies
x(t)= -3t+a.
Also,
dz _ 0
dt '
which implies
z(x(t),t) = c.
Thus,
z(x(0), 0) = sin(x(0)).
z(x, t)
it
Dz Dz
Figure 3.4: The solution sketch for -3 =0
Dt Dx
Figure (3.4) shows a sketch of the solution for values 0 < t < 1 and -?C <x <7r.
Two of the parallel characteristics x = a - 3t are clearly evident in the solution
sketch. They are represented by a solid line for a = and a dashed line for a = it.
2
We will now consider the more general first-order constant coefficient PDE
Du Du
a u=0 (3.38)
with IC
where-oc<x<oo,0<t<oo, and8>0.
As introduced in the beginning of this section, one way to solve this problem
is to consider the rate of change of u(x(t), t) as measured by a moving observer,
x = x(t). Applying the chain rule yields
du _ Du dx Du
(3.40)
dt Dx dt + Dt
dx
If we assume that = a the left side of Equation (3.40) can be substituted into
dt
Equation (3.38). Then, we have
du du
+/3u=0 or = -/3u,
dt
Section 3.5: Method of Characteristics 61
Because of the term e-13t, we know as time goes to infinity, u(x, t) tends to zero. This
seems quite reasonable if we assume Equation (3.38) is a first-order wave equation.
In this case, the term 3u is a damping function that tends to flatten out the wave
as time passes.
5
1
u(x, t)
with IC
u(x, 0) = sin x, (3.43)
where -oo <x < oo and 0 < t < oo. Since c = 0.25 and ,3 = 0.5, we know from
Equation (3.41),
u(x, t) = f(x - cEt)e_t,
that the specific solution is
u(x, t) = sin(x - 0.25t)e-o.5t.
Figure (3.5) shows a sketch for -2ir < x < 2ir and 0 < t < 5. Two parallel
characteristics are shown on the graph. However, remember these characteristics
take into consideration the exponential e-o.5t
EXERCISES 3.5
3.5.1. Solve and graph the following using the method of characteristics.
az az
(1) 5 = 0 with IC z x 0= 2x.
az az
(2) -3 = 0 with IC z x 0 = x2 - 2x 1.
aw aw
(3) ax +6w = 0 with IC w(x, 0) = cos x.
at
au au
2u = 0 with IC u(x, 0) = sinx.
(4) at + 4ax -
3.5.2. Consider the simple concentration problem
au au
= o, 0 < t < oo, - oo < x < oo,
at + ax
with IC
u(x, 0) = cosx.
Solve and graph the solution using your choice of mathematical software. Does
the solution satisfy the PDE and the IC?
3.5.3. Solve
av av0<t<oo
-+x-=0, -oo<x<oo
at ax
with IC
v(x, 0) = 0.5x2.
Graph the solution using your choice of mathematical software. Does the
solution satisfy the PDE and the IC?
Section 3.5: Method of Characteristics 63
3.5.4. Solve
au au
at
+ax+tu=0, 0<t<oo, -oo<x<oo,
with IC
u(x, 0) = x2.
Graph the solution using your choice of mathematical software. Does the
solution satisfy the PDE and the IC?
3.5.5. Solve the surface wave problem
Du +a Du + au + u=0 0<t<oo -oo<x<oo -oo< <oo
at ax a "
with IC
u(x,y,0) =sin(x+y),
where c, ,3, and 'y are constants.
3.5.6. Solve the surface wave problem
au
at
au
+cEax
au
+/3-
ay
+u = 0 0< t <00, -00 <x <00, -00 <yy <00,
o0 ,
u(x,y,0) =sin(x+y),
where c = 2, ,3 = -1, and 'y = 1.
3.5.7. Given the first-order semilinear PDE
au(x,y) au(x,y)
a (x,y, u(x, u)) + b(x ,y, u(x ,y)) ax = c (x,y, u(x,y)),
ax
show that the method of characteristics yields
dx dy du(x, y)
a(x, y, u(x, y)) b(x, y, u(x, y)) c(x, y, u(x, y))
au au
Equation, let v = u w = and z = . Show that v w, and z must
ax ' at >
Du
Letting z = c ields the first-order PDE
yields
Dt + Dx
az az
at-cax=°,
which we know from the previous section has the general solution
8u 8u
z(x, t) = at + sax = Q(x + ct), (3.47)
av av
at
+c ax =0,
which has the solution
au au
v(x, t) = at -sax =Pox - Ct) (3.48)
To apply the second IC, we use the chain rule. Consider the term G(x + ct) in
Equation (3.49). It is actually a composition of the functions G(X) and X (x, t) _
aG _ dG aX ax aG _ aG _ dG
x + ct. Thus, The fact that = c implies c c
at dX at ' at at aX dX '
G
since G is a function of X only. Therefore, cddX = cG' (X ) = cG' (x+ ct). Similarly,
aF _
at - -cF' (x - ct)
Using the method of the previous paragraph and applying the second IC, we
obtain
8u(x, 0)
= c [G'(x) - F'(x)] = g(x). (3.51)
To find a solution for G(x), we take the derivative of Equation (3.50) with respect
to x, multiply by c, and add it to Equation (3.51), which yields
2
[+i'x)]
c
( )
In Equations (3.52 and 3.53), k1 and k2 are constants of integration. However, since
G(x) + F(x) = 1(x), we must have k1 + k2 = 0. Therefore, the general solution of
Equation (3.44) is the sum of Equations (3.52 and 3.53), each shifted a distance of
ct; u(x, t) can be written in the form
f9(S)dsJl
u(x, t) = 2 [f(x + ct) + f (x - ct)] + g(s)ds -
2c [,I o +ct
subject to
h, if x <a
u(x, 0) _
0, if x > a
8u(x,0) _ p
at
Determine the solution using d'Alembert's method and graph the solution for sev-
eral values of time t.
u(x,t) = {f(x+ct)+f(x-ct)]
-a a
h
if x <a
F(x)=G(x)=
I 0, if x > a.
In Figure (3.6), we have t = 0 and u(x, t) = 2 [f(x) + 1(x)] = f (x) = h. Since
x <a, we have the endpoints of the graph at -a and a. Remember, the full pulse
h
is made up of two smaller pulses. Each smaller pulse has a height of 2 and as soon
68 Chapter 3: One-Dimensional Wave Equation
a
Figure 3.7: Graph of d'Alembert's solution for 0 < t < .
c
h/2
H H
-2a 2a
a
Figure 3.8: Graph of d'Alembert's solution for t = .
c
endpoints for the left pulse, which are -2a and 0. The right endpoints are obtained
similarly by letting x = x - ct = x - a. Thus, the right endpoints become 0 and 2a.
a
Finally, for Figure (3.9), t> , and we have two separate pulses, each of height
c
h/2
-a - ct a-ct -a +ct a + ct
a
Figure 3.9: Graph of d'Alembert's solution for t> .
c
h
2. One pulse is traveling to the left and the other pulse is traveling to the right.
The pulse traveling to the left has endpoints of -a - ct and a - ct. These endpoints
are determined by letting x = x + ct while remembering x < a. Thus, we have
x + ct <a or -a < x + ct <a, which becomes -a - ct <x <a - ct. Similarly,
using x = x - ct, the right endpoints are found to be -a + ct and a + ct.
EXERCISES 3.6
u(x, 0) _
au(x, 0) _ 0?
at
Sketch the solution for several values of time, t.
70 Chapter 3: One-Dimensional Wave Equation
u(x, 0) =0
and
au(x, 0)
at 00, , 1?
u(x, 0) _
0,
0, x>1
and
8u(x, 0)
0?
at
u(x, 0) =0
and
8u(x, 0)
at
subject to
- sin x, x < 7r
u(x, 0) _
0, >r
and
au(x, 0)
0?
at
8u(x, 0)
0?
8t
u(x, 0) = 0
and
8u(x, 0)
sin x, x < r
at 0, > it?
Sketch the solution for several values of time, t.
72 Chapter 3: One-Dimensional Wave Equation
u(x, 0) = 0
and
cosx, <
8u(x, 0)
at o xx> 2-.
Sketch the solution for several values of time, t.
3.6.9. Solve for propagation of electrical vibrations in an infinite conductor for the
condition
GL = CR,
where G, L, C, and R are the leakage conductance, self-inductance, capac-
itance, and resistance, respectively, per unit length of the conductor. At
the start of the experiment, the current is i (x, 0) = 1(x) and the voltage is
v(x, 0) = g(x).
3.6.10. Given
a2u 25uu
=c axe' (3.55)
ate
subject to
u(x,0) = f(x)
73
74 Chapter 4: The Essentials of Fourier Series
formally introduced to linear algebra. But, all students who have taken the calculus
sequence, ODEs, and some type of engineering mathematics course have worked
with linear algebra through most of their college career. Therefore, this chapter
builds on this material to develop the ideas from linear algebra that are needed
to understand the connection with Fourier series in the section titled "Elements of
Linear Algebra." The chapter then develops the mechanics of Fourier series. Finally,
even and odd functions and their Fourier series representation are discussed.
The linear algebra portion of this chapter should not be construed as a replace-
ment for a linear algebra course. We cover only those theorems and ideas relevant
to developing Fourier series. Also, we do not present the proofs nor much of the
mechanical manipulations that are necessary for a better understanding of linear
algebra. If you enjoy the material covered in the linear algebra portion, I would
suggest you take a linear algebra course specifically designed for applications. Also,
the notation of this chapter may be new to you. If this is the case, I suggest you
read Appendix D on mathematical notation.
This is a very concise definition, but what does it mean? We will start with
vector addition.
Vector addition includes several ideas. First, when adding two vectors in a vector
space V, the sum is also in V. This concept is known as closure of the space. Second,
vector addition is commutative: If x and y are vectors in V, then x + y = y + x.
Third, vector addition is associative. This means that if x, y, and z are vectors in
V, then x + (y + z) = (x + y) + z. Last, there exists a unique vector 0 in V such
that if x is any vector in V, then x + 0 = x. Also, there exists the vector -x in V,
such that x + (-x) = 0. Next, we consider scalar multiplication.
Scalar multiplication also includes several ideas. First, a scalar is nothing more
than a number, and sometimes you will read or hear the phrase "a real vector
space." This means that the scalars are all real numbers. If you are reading a math
text and you see the phrase "complex vector space," then the scalars come from
the complex plane and have the form a + bi. In this text, we will work exclusively
with real vector spaces. Also, a vector space V is closed under scalar multiplication.
Second, scalar multiplication distributes over vector addition. This means if x and
y are vectors in V, and a is any real number, then a(x + y) = ax + ay. Third, scalar
multiplication is associative. That is, if x is any vector in V and a and b are real
numbers, then a(bx) = (ab)x. Finally, there exists an identity element for scalar
multiplication. Since we are working exclusively with real vector spaces, you can
clearly see that 1 is our identity element.
In many texts, Definition (7) is written in the following form:
Definition 8. A real vector space V is a non-empty set of elements x and y (called
vectors), such that:
1. For any pair of vectors x and y in V, there exists a unique vector in V that
is the sum of x and y.
5. For every vector x in V, there exists the vector - x in V, such that x+ (-x) = 0.
6. For all vectors x in V and all real numbers, a, the unique vector ax is in V.
7. Scalar multiplication distributes over vector addition. For any pair of vectors
x and yin V and real number a, we have a(x + y) = ay + ax.
8. Scalar multiplication is associative. For all vectors x in V and the real num-
bers a and b, a(bx) = (ab)x.
Also, at first glance, the three vectors in the above example seem to be unrelated.
But, a careful inspection shows 3u - v = w or
-2
0
1
1
This means the vector w can be written as a linear combination of the other two
vectors. Thus, the vectors u, v, and w are linearly dependent. The formal definition
of linear dependence in a general vector space follows:
Definition 10. Let V be a vector space. Suppose the vectors u1, u2, u3, ..., un
are elements of V. If al, a2, a3, ..., an are scalars and a linear combination of the
vectors u1, u2, u3, ..., un ,
a1 u1 + a2 u2 + a3 u3 + ... + an un = 0,
(where ai, 1 < i < n) are not all equal to zero, then the vectors u1, u2, u3, ... , un
are linearly dependent.
To explain this definition, consider the three vectors u, v, and w given in Ex-
ample (4.1). Since w = 3u - v, we have w - 3u + v = 0. Here a1 = 1, a2 = -3,
and a3 = 1. Now that we know what linearly dependent vectors are, we need to
define linearly independent vectors.
Definition 11. Let V be a vector space. Suppose the vectors v1, v2, v3, ..., vn are
elements of V and a1, a2, a3, ..., an scalars. Then, the linear combination of the
vectors v1, V2, V3, ..., vn is given by
a1v1+a2v2+a3v3+...+anvn = 0.
The vectors v1, v2, v3, ..., vn are said to be linearly independent only when ai =
0 f or all i such that 1< i< n.
To demonstrate this definition clearly, consider the following example.
EXAMPLE 4.2. Let
1 -1 0
v1= 1 ,v2= 1 ,and v3= 1
-1
-1 --11 1
We will show that these vectors are linearly independent. Suppose we form the
linear combination of the vectors v1, v2, and v3 as
1 - 1 0 0
al 1 +a2 1 + a3 1 0
-1 -1 1
1 0
where we assume at least one of the scalars a1, a2, or a3 does not equal 0. From the
first line across all three vectors, we find that a1 = a2 because a1 - a2 = 0. Also,
78 Chapter 4: The Essentials of Fourier Series
from the second line across all three vectors, a3 = -2a1 since a1 + a2 + a3 = 0.
Using substitution we have 2a1 + a3 = 0. However, from the third line across all
three vectors, -a1 - a2 + a3 = 0, implying a3 = 2a1. Thus, 2a1 = a3 = -2a1. This
in turn implies a contradiction to our assumption, since 2a1 = a3 = -2a1 implies
al = 0, which means a2 = 0 and a3 = 0.
It is easy to see that the vectors
L
] and [o]
in R2 and
1 0 0
0 1 and 0
0 0 1
in 113 are linearly independent. These vectors are referred to as the standard basis
vectors for the vector spaces IR2 and IR3, respectively. Considering the format of the
standard basis in R2 and J3, we could guess that the standard basis in Rn would
have n vectors, each vector with n components, one in the nth position and all other
positions with a zero. This would be a good guess and quite correct. In fact, the
usual way to state the standard basis in 1Rn is
1 0 0
0 0 0
0 1 0
0 0
0
L0i 0 r
0 r r
1
where it is understood that there are n vectors with n components in each vector.
There are other bases for IR2, 113, and IRn. These will be discussed in a later
subsection. Next, we'll discuss orthogonality and inner product. These topics may
sound new to you, but they really are not.
We use the term inner product because the dot product only works on the vector
spaces Jn, for n e N. In other vector spaces, the dot product would make no sense.
For instance, the matrix space M2,2 cannot use the dot product as its inner product.
This naturally leads us to the general definition of an inner product.
Definition 14. Let V be any real vector space. An inner product on V is a func-
tion that assigns a real number to each pair of vectors u and v of V, written
(u, v) = a e III, satisfying:
It is left as an exercise for you to show that the inner product (dot product) in
the vector spaces IIS2 and ii is, in fact, an inner product. Also, it should be noted,
any vector space where an inner product has been defined is known as an inner
product space.
An example of an inner product space, that is not one of the W' 's is the inner
product space composed of all 2 x 2 matrices with real-number components. If A and
B are elements of M2,2, then the inner product is defined as (A, B) = tr(BT A).
In the next example, we prove the space of all 2 x 2 matrices with real-number
components is an inner product space.
EXAMPLE 4.3. We must show that the space of all 2 x 2 matrices with real-
number components, M2,2, is an inner product space. We are using the inner
product, (A, B) = tr(BT A). We start with the vector A, where
a b
A c d
Thus, we have
(A, A) = tr(AT A) = tr (f a c1 ra
\lb dJ L
a2 + c2 ab + cd
= tr ba + do b2 + d2
If
A- 00
0 0 '
then
Also, if
Section 4.2: Elements of Linear Algebra 81
and we have
a2 + c2 ab + cd 11
= tr
ba + do b2 + d2
= a2 + c2 + b2 + d2 = 0
A=La d andB=Lg e
Then,
(A,B)=tr(BTA)=tr(F e g1 [a b 1Lf
h] Lc d])
=tr rI ea + gc eb + gd
fa+hc fb+hd ])
L
= ea+gc+ fb+hd,
and
c1 re f 11
(B,A)=tr(ATB)=tr (f
\lb
a
dJ[9 h]J
=tr ae+cg of +ch
be+dg bf +dh ])
=ae+cg+bf +dh.
Therefore,
(A,B) _ (B,A) for any A and B in M2,2.
Next, we show that
(A+B,C) = (A,C)+(B,C) for any A, B, and C in M2,2.
82 Chapter 4: The Essentials of Fourier Series
Let
Then,
=tr(f m(b+f)+o(d+h) 11
n(a + e) + p(c + g) n(b +1) + p(d + h) J J
and
Lra
tr pJ
pJ Lg hJ/
ma+oc mb+od r me+og mf +oh
tr
na+pc nb+pd ) +tr I( ne+pg of +ph I
= ma+oc+nb+pd+me+og+nf +ph
Therefore,
Let
A=La dJ andB=[e
9
h]'
Section 4.2: Elements of Linear Algebra 83
Then,
(f e g 1 ,<3a 3b 11
(,<3A, B) = tr (BTA) = tr
\Lf hi L dj)
e/3a + g/3c e(33 + g/3d
= tr
f /3a + h/3c f /3b + h/3d )
e g 1 [a b n"
h][c d])
=,6tr I L fa+hc fb+hd ])
= /3(ea+gc+fb+hd).
Thus,
Therefore, we have shown that M2,2 is an inner product space with inner product
defined by (A, B) = tr(BT A).
A - aI = C5 3 1 Ol(5-\
)_ 3 1
3 5 0 1 3 5-,\
Next, find the determinant of A - ,\I, and set A - aI = 0. That is,
5-a 3
0= 35-a = (5-a )(5-a -9=a2-10a+25-9
)
=a2-10\+16.
The equation ,\2 - 10\ + 16 = 0 is called the characteristic polynomial, and it
can be solved for roots by factoring. Once factored, the roots are = 2 and \ = 8.
These are the eigenvalues. Generally, they are labeled as = 2 and '\2 = 8.
To determine the associated eigenvector for each of the eigenvalues, form the
equation (A - \I) x = 0 for each eigenvalue. Thus, for the eigenvector correspond-
ing to the first eigenvalue, we solve
0= (A-)1I)x= 5 3- 1 5-\1
3 x1
x2
This equation can be solved using the Gauss-Jordan reduction method and the
eigenvector x is given by
(xi(
x
We now have the corresponding eigenvector to the eigenvalue )1 = 2. In a similar
fashion, the corresponding eigenvector x to the eigenvalue )2 = 8 is
Section 4.2: Elements of Linear Algebra 85
-1) and I 1
4.2.5 Significance
In the previous subsections, vector spaces, linear combinations, linear independence,
inner product, orthogonality, and eigenvalues and eigenvectors were discussed, but
they were not related. Here, hopefully, we will fit the pieces of the puzzle together.
All of the examples will be centered on the real vector spaces, and R3, since you
are most familiar with them.
Everything that occurs in this section occurs in a vector space. This is an impor-
tant concept. It means mathematics and reality are tied together in some fashion
and that fashion is the development of mathematical objects that act the same as
three space,
In a vector space, a basis is extremely important and necessary. A basis is
composed of linearly independent vectors. Another important point in the real
vector spaces W, n e N, is that the n in the superscript position tells us how many
vectors are in the basis. For instance, IIhas three vectors in the standard basis,
and they are
Also, the n in the superscript position identifies the vector space to be finite di-
mensional. This does not mean the space does not go on to infinity, it just means
there are only a finite number of copies of the real axis being used to make up
the space. Are there infinite dimensional spaces? The answer is yes, and in the
next section you will be introduced to one, the function space of piecewise smooth
functions. But, before we discuss infinite dimensional space, let's tie together some
of the tools we need from the real vector spaces.
Basis vectors are linearly independent. This means any vector in the space can
be expressed as a linear combination of the basis vectors.
EXAMPLE 4.5. Express the vector
-3
18
7
86 Chapter 4: The Essentials of Fourier Series
Although Example (4.5) is easy to complete and, in fact, can be completed al-
most unconsciously, it really should bring up two very important questions. First, is
the standard basis the only basis? Second, if another basis exists, how do we deter-
mine the multiplicative constants of the basis to form a linear combination as in the
standard basis? We will answer each of these questions because an understanding
of the underlying concepts is necessary.
The standard basis is not the only basis for R3, or for that matter any vector
space. The standard basis is the easiest one to work with because in most vector
spaces the standard basis is known to be orthonormal, meaning the vectors that
make up the basis have unit length and are orthogonal to each other.
Another orthonormal basis for IIis
1 1
r -1 1
1 L/J
If you want to express
as a linear combination of the orthonormal basis in Equation (4.1), you must deter-
mine the constant multipliers. It does not seem to be an easy task. However, the
inner product is the tool that makes it happen. If you compute the inner product
of the vector with each vector in the basis in Equation (4.1), then the answers are
the respective constants.
EXAMPLE 4.6. Express the vector
LVJ LJ
Section 4.2: Elements of Linear Algebra 87
Solution: First, compute the inner products of the vector and the orthonormal
basis. That is, find
r-1-1
2
1
LJ
and
r -1 1
10
L/J
Note: Here the inner product yields the contribution of the vector in the direction
of each unit vector; this is because the basis vectors are orthonormal. Next, form
the linear combination, which is
r-1-1 r -1 1
22 32 10
0
1 L/J
1
LJ
Example (4.6) shows us how to find the constants of the linear combination
when we have an orthonormal basis.
Since we know other bases exist for the real vector spaces, how do we find them
and how do we make them orthonormal? The answer lies in two theorems stated
here, with proofs in Appendix B and a process, and it involves topics discussed in
the previous subsections. The topics are eigenvalues and eigenvectors.
Theorem 16. Let , '\2, /\3, ... ) be distinct eigenvalues of an n x n matrix.
Then, the corresponding eigenvectors x1, x2, x3,. . xn form a linearly independent
.
1 <i <n.
Theorem 17. If an n x n matrix A has n distinct eigenvalues, then the correspond-
ing eigenvectors form a basis for W.
88 Chapter 4: The Essentials of Fourier Series
EXERCISES 4.2
4.2.1. Which of the following sets of vectors in IIare linearly dependent, and which
are linearly independent:
0 1 3
(1) 2 2 6 ,
LaJ 3 3 6 J
( 1 -1 0
(2) 2 2 1 ,or
( 3 -5 7
J
1
2
3
(1)
2 ii
[
2 -2 3
(2) 0 3 -2 and
0 -1
-1 2
2
2 23
(3) 1 2 1
2 -2 1
Section 4.2: Elements of Linear Algebra 89
find the eigenvalues and eigenvectors. Then, perform the following process:
(1) Name the eigenvectors a1, a2, and a3 (order doesn't matter).
(2) Let vector /1 = a1.
(3) Find the vector /32 by using the formula
(a3,,3i) (a3,132)
/33 = a3 - (/3/3)/31 - (/32,/32)/32
(s) The vectors i1, /32, and /33 are an orthogonal basis for Make them
an orthonormal basis.
(6) This process is known as the Gram-Schmidt process. Prove you have
produced an orthonormal basis for R3, which can be used to describe
every vector in II.
4.2.4. Show that the dot product in the vector spaces IIg2 and IIis an inner product.
4.2.5. Show that the space of all 2 x 3 matrices with real-number components, M2,3,
is an inner product space. Note: Use the inner product (A, B) = tr (BTA),
where A and B are elements of M2,3.
4.2.6. Show that the space of all 3 x 2 matrices with real number components, M3,2,
is an inner product space. Note: Use the inner product (A, B) = tr (BTA),
where A and B are elements of M3, 2 .
4.2.7. If w1 and w2 are positive real numbers, show that the definition
1.5
0.5
a b
x
-2 -1 1 2
- 0.5
0 x
Definition 19. A function f (x) on the interval [a, b] is piecewise smooth if the
interval [a, b] can be broken up into a finite number of subintervals such that the
df
function, 1(x), and its derivative (x) are continuous on each of the subintervals.
dx
This definition of a piecewise smooth function given implies that the function
1(x) may not be continuous on the interval [a, b], but the discontinuities are only
jump discontinuities. Thus, the function f (x) is bounded on the interval [a, b]. Both
Figures (4.1 and 4.2) are good examples of piecewise smooth functions. Another
example is f (x) = ( x, graphed in Figure (4.4) on the interval [-2,2]. Note: The
The functions and their first derivatives are continuous except at possibly a
finite number of jump discontinuities.
Understanding the setting for the function space of piecewise smooth functions
is only part of the story. The other pieces of the story are discovering reasonable
definitions for inner product, orthogonality, and basis.
Definition 20. Let V be any real vector space. An inner product on V is a func-
tion that assigns to each pair of vectors u and v of V a real number, written
(u, v) = c e Ilk, satisfying the following:
Remember, the words "real vector space" mean the associated scalar field to the
vector space is the real numbers.
Let's look at the different operations we can perform with functions. First, there
is addition. In general, when two functions are added together, we get another func-
tion not a real number. Next, we could consider multiplying two functions together.
But, multiplication of two functions usually yields another function. Moving into
the realm of calculus, differentiation could be considered. However, differentiation is
really defined at a point and does not apply to functions that are piecewise smooth
because of the jump discontinuity. Therefore, differentiation would not be a good
candidate for the inner product. Finally, we can consider integration. A definite
integral can be applied to a function on the entire interval and produces an answer
that is a real number. Also, a function that has jump discontinuities, as defined
previously, can be integrated, and we can easily integrate the product of two func-
tions. Thus, integration is a good candidate for the inner product. Therefore, we
define a possible new inner product as
e
(u, v) = f u(x)v(x) dx.
Section 4.3: A New Space: Function Space of Piecewise Smooth Functions 93
Note: u and v are no longer boldfaced since they are functions. Thus, not vectors
in the traditional sense.
Now, it must be determined if this definition meets all the criteria of the defin-
ition of an inner product.
First, we must ask, "is
(u,u)=f v.(x)u(r)
e dx = 16 [u(T)]2dz > 0
J
on the interval [a, b]?" We know, from calculus, the square of a function is always
nonnegative. The integral of a nonnegative function is also nonnegative, since the
integral is always nonnegative when the curve of the entire function is above the
x-axis. Therefore, (u, u) > 0 for all functions u(x) in the function space of PWS
functions. Also, if u = 0, then
f6 fb
(u,u) = (0,0) = Odx=J Odx=O.
Ja
If
b b
(u+v,w) /eu(a)w(x)+v(x)w(x)dz
= f [v.(x)+v(x)]w(x)dx= J
6
for any u and v in the function space of PWS functions and c a scalar in IIt.
We have shown the definition
b
( u, v) = u(x)v(x) dx
f
is an inner product on the function space of PWS functions. In fact, it is the
standard inner product for the function space of PWS functions, which is analogous
to the dot product in the vector spaces IIYn, for any n e N.
It is convenient to change the notation sightly as in
b
x
-L L
Figure 4.5: The graph of 1(x) = -x and g(x) = 2 on the interval [-L, L].
0 0
o"(x) = -A(x),
subject to the BCs
dcp(L) dcp(-L
dx dx
on the interval [-L, L]. We must solve Equation (4.2) for A < 0, A = 0, and A > 0.
For A <0, we let -s = A, s > 0. Then, Equation (4.2) becomes
p"(x) = sp(x).
From your previous courses you should recognize this ODE and know that the
solution is dependent on A. First, for A <0, we have the solution
Since the hyperbolic cosine function is an even function and the hyperbolic sine
function is an odd function, we rewrite Equation (4.5) as
(x)=cicosh(s/x). (4.7)
o"(x) = 0.
Integrating twice yields
2d1 L = 0,
The second BC from Equation (4.3) indicates that we must find the derivative of
cp(x) = d2. Since d2 is a constant, the second BC is trivially true. Thus, for A = 0
we have (x) equal to the constant d2.
Finally, we assume that A > 0. The solution, in this case, is
(x) = a cos Vx + b sin Vx. (4.11)
Since the cosine function is an even function and the sine function is an odd function,
we rewrite Equation (4.12) as
Equations (4.13 and 4.15) are very similar. Dividing out the 2a and 2b yields
sin (v5L) = 0
2
for both equations. Solving for A, we find A = (-i--) n = 1,2,3,... Remember,
the sine function equals zero when the argument of the sine function is an integer
multiple of 71. Thus, when A > 0, the solution for Equation (4.2) is
n7rx
cp=
(x) a cos + b sin
L
. (4.16)
L
98 Chapter 4: The Essentials of Fourier Series
We have all the values of A, called eigenvalues, which allow for a nontrivial solu-
tion to Equation (4.2) subject to the BCs given in Equation (4.3). These eigenvalues
are analogous to the eigenvalues in Theorem (16) and Theorem (17). Also, Equa-
tions (4.10 and 4.16) make up the general solution of Equation (4.2). The solution
is made up of three different types of functions. The three different functions are a
constant, cos and sin called eigenfunctions. They are analogous to the
L L
corresponding eigenvectors of Theorems (16 and 17). The eigenfunctions are often
referred to as orthogonal sequences. The orthogonal sequences are cos and
L
n7rx n7rx n7rx
sin . Remember, when n = 0, sin = 0 and cos = 1. Thus, the con-
stant may be included in the cos sequence. However, in your previous course
L
work, the orthogonal sequences were referred to as the fundamental solution set.
We know the functions of a fundamental solution set are linearly independent, and
linear independence is a basic property of any basis, and since they are also known
as orthogonal sequences, we know that they are orthogonal. However, we will not
take for granted that they are orthogonal. This will be proved.
As the previous paragraph suggests, solving the ODE, Equation (4.2), is analo-
gous to solving an n x n matrix for the distinct eigenvalues and the corresponding
eigenvectors. Remember, eigenvalues are found when you solve the matrix equation
Ax = Ax or A - AI x = 0. Also, when eigenvalues for an n x n matrix are found,
the eigenvectors may be determined. By Theorem (17), Eigenvectors form a basis
for the n-dimensional space that is represented by the n x n matrix. The space
we are considering is a function space. Therefore, instead of eigenvectors, we have
eigenfunctions, for the function space formed on the interval [-L, L]. It remains for
us to show they are orthogonal.
To show that the eigenfunctions found are orthogonal on the interval [-L, L], we
must show several things. You must remember there is one constant and an infinite
number of both sine and cosine eigenfunctions, all of which must be orthogonal to
each other. We start by showing that
L
L cosmirx
sinn7rx L dx = 0, n, m E N.
J-L
First, the trigonometry identities
si n ( a + b) = s i n a cos b + cos a s i n b
and
sin (a - b) = sin a cos b - cos a sin b
must be used. We have
n7rx m7rx
s in
(-i- + L si n L cos m7rx
n7rx
L + cosn7rx
L sin .
L ( 4 . 17)
and
nirx nirx
si n L - mirx
L = si n L - cos nirx
L cosmirx L sin mirx
L
,
. ( 4 . 18 )
Section 4.3: A New Space: Function Space of Piecewise Smooth Functions 99
L L -z
Dividing both sides by two and integrating from -L to L, we find
mirx nirx nirx
1
2 J-L
L
{sin
n'irx
L
+
L
+ sin
L
- mirx
L
dx =
J-L
L
sin
L
mirx
cos L dx.
Integrating yields
2-L(LL)
1
+ sin
nirx mirx
L . nirx L- mirx
L dx
2
fLsin/ (n
L )xdx
L L
(n+m)n / -L (n-m)n
z
L )(_L)+(n_m)4CO5( L )(_L)]
=0.
Remember, the cosine of the angle equals the cosine of minus the angle.
The rest of the orthogonality integrals are
L nirx mirx
sin-----sin------- dx =0, n, m E N, n m;
-L L L
L
cosnirx
L cosmirx
L dx = 0, n, m E N, n
I L
m;
L n7rx
d2 cos L dx = 0, d2 a constant, n e N, n 0;
-L
and
L
f d2 sin n7rx
L
dx = 0, d2 a constant, n e N, n 0.
L
They are left as exercises.
100 Chapter 4: The Essentials of Fourier Series
We have shown that the eigenfunctions d2, a constant, sin and cos
L L
are orthogonal eigenfunctions. In fact, they form the standard basis of the function
space of PWS functions. Therefore, we are now ready to define a linear combination
of our basis. Note: A linear combination of our basis must represent every function
in the function space of PWS functions.
where a0, an, and bn are the scalar multipliers of our eigenfunctions and means
approximately. Note: A scalar times a constant is just another constant. Thus, a0
is the product of an arbitrary scalar and the constant d2.
The right side of Equation (4.19) is known as a trigonometric Fourier series. It
is called trigonometric, since sine and cosine functions are used in the sum. Also,
Equation (4.19) is sometimes called an eigenfunction expansion, where sine and
cosine functions are the eigenfunctions. There are other types of Fourier series.
They also form a standard basis but for different intervals/coordinate systems.
However, at this time, there is no need to go into detail about them. They will be
discussed in Chapter 8, when generalized Fourier series are introduced, and used
extensively in Chapter 10. For purposes of this text, the trigonometric Fourier
series above will be called a Fourier series for short. When we introduce generalized
Fourier series, appropriate names will be applied.
Equation (4.19) identifies the linear combination of our proposed standard ba-
sis. The constants a0, an, and bn, are known as Fourier coefficients, and we must
determine them.
For each different function in the space of PWS function, the scalars a0, an,
and bn are found in the same way the scalar multipliers were found in Example
(4.6). In Example (4.6), the scalar multipliers were found by using the standard
inner product for IR, called the dot product. The vector to be represented was
used in the dot product with each eigenvector, yielding the scalar contribution in
the direction of each eigenvector. When the eigenvectors and corresponding scalars
were put in a linear combination, they formed another representation of the vector in
question. Thus, it would seem natural to find the scalars a0, an, and bn in Equation
Section 4.3: A New Space: Function Space of Piecewise Smooth Functions 101
(4.19) for each function in the space of PWS functions by using the standard inner
product for the space of PWS functions, integrating the product of the function to
be represented and the eigenfunctions from [-L, L]. The obvious solution is not as
easy as the question. How do we do it?
First, for convenience, the ti in Equation (4.19) will be changed to =. Thus,
Equation (4.19) becomes
00
The inner product for the function space requires the integration over the in-
terval [-L, L] of the product of two functions. This means we have to multiply
Equation (4.20) by some function, and the object of the multiplication and subse-
quent integration is the determination of a Fourier Coefficient. This implies using
the orthogonality of the sine and cosine functions to our advantage. Thus, to find
an, for any n e N, we would multiply Equation (4.20) by cos
L
for some (f)
particular m e N, m 0. This yields
mirx 1
f (x) cos ( L x) = ao cos ( L
l
00
00
-I- cos Lan cos (nLx) -I- bn sin (nLx) ] . (4.22)
n=1
\L x1
We integrate Equation (4.22) from -L to L. Thus, we have
L
f (x) cos ( L x) dx = a0 f L cos ( L x) dx
L
+
f L n-1
cos ( L x) [ate, cos (nLx) + sin (nLx) ] dx. (4.23)
'Lmirx
LLcL) dx,
102 Chapter 4: The Essentials of Fourier Series
in Equation (4.23) is in its final form unless we know the function, 1(x). The second
integral,
L mix
cos ( L ) dx,
Jf L
ao
an
f cos (nZ)(\)
xcos L xdx,
will, by orthogonality, equal zero whenever m n. However, when m = n, we have
(nirx) 2 /
dx = an LL 2 -I- 2 cos 12 L x ) dx = an L.
an LL cos
Thus, the entire infinite sum in Equation (4.24) becomes one term, and that term
is an L . Therefore,
L nix
f (x) cos L dx = anL.
fL
Solving for the Fourier Coefficient an yields
1 L nirx
an = zf f (x) cos L dx.
L
Section 4.3: A New Space: Function Space of Piecewise Smooth Functions 103
and
and the integral equations for determining the Fourier coefficients, which are
L
(4.26)
J
dx,
L f(x)
/'L
an = L f (x) cos (nLx) dx, (4.27)
J L
and
L
(nirx)
bn = L
I L
1(x) sin dx. (4.28)
Thus, for any function, f (x), in the function space of PWS functions on [-L, L],
we can define a Fourier series and determine the Fourier Coefficients. It still must
be determined if a Fourier series actually converges to the function, f (x), in the
function space of PWS functions. This important point is discussed in the next
subsection.
f(x)=x3-4-7x2-4x-1O.
104 Chapter 4: The Essentials of Fourier Series
f2
a° 2L f
,
L
L
f fix) dx = 4
J 2
x3 + 7x2 - 4x - 10 dx = 3
and
11
L
f(x
f () cos os-
nirx
dx
p2
1
(x3 -I- 7x2 - 4x - 10) cos (nLx) dx
2
/2
1 (224 cos (nom) 224 sin (nom) 72 sin (nor)
2 n
n
since cos(nir) _ (-1)n and sin(ner) = 0 for all n. Thus, the Fourier series represen-
tation of the function 1(x) is
00
-2
x3 + 7x2 - 4x - 10 =
3
+
n=1 n
112(-1)n
2 2
cos
nirx
L
+
96(-1)n
n3 3 sinnirx
L()j
Figures (4.7,4.8,4.9 and 4.10) are four graphs of the Fourier series of the function
1(x) for the partial sums Si(x), S5 (x), S12 (x), and S25 (x), respectively. Note: The
Fourier series converges to the entire function, 1(x), rather quickly.
_ -2 96(-1)n
Si(x) 3 +
1
n=1 n
112(-1)n
22 cos
nirx
(----)
L + n3 3
sin
nirx
(-h--)
5
(fl1X) (fl7IX)J
cos + 96(313)n sin .
L
n=1
x
x
Figure 4.7: The Fourier series repre- Figure 4.8: The Fourier series repre-
sentation of 1(x) = x3 + 7x2 - 4x -10 sentation of f (x) = x3 + 7x2 - 4x - 10
for the partial sum 81(x). for the partial sum Ss (x) .
-2
12
r112(-1)
cos
/nlrxl + 96( )n sin
Siz(x) -
3
n=1
L n2 l J
-2 25 r112 -1)n
n2 nixl 96(-1) (fl7IX
Szs (x) = 3 + cos (-i l + sin .
n=1
I
L
-z )j
106 Chapter 4: The Essentials of Fourier Series
Figure 4.9: The Fourier series repre- Figure 4.10: The Fourier series rep-
sentation of 1(x) = x3 + 7x2 - 4x -10 resentation of 1(x) = x3 -1-7x2 -4x- 10
for the partial sum 512(x). for the partial sum S25(x).
In the next two examples, the functions chosen demonstrate the action of the
Fourier series at a jump discontinuity. An explanation is provided, then formalized,
in Fourier's convergence theorem.
EXAMPLE 4.9. Consider the function h(x) = x2 -I- 2x - 1 on the interval [-5, 5]
shown in Figure (4.11) . The Fourier series representation of h (x) is
00
nix nix
x2 + 2x - 1 = a0 -E-
n=1
[an cos
5n-
-I- bn sin
---j
5
where
L f5
a° 2L h(x) dx = 10
J 5
x2 + 2x - 1 dx = 3 ,
Section 4.3: A New Space: Function Space of Piecewise Smooth Functions 107
an = L J L h(x) cos
(fl7X)
dx = 5 5(x2 + 2x - 1) cos
(nx) dx
J
100 100 48 100(-1 n
= n sin(ner)
and
(fl7IX) nrx
sin dx = 5 J (x2 -I- 2x - 1) sin (-) dx
5
20 20 20(-1)n
= n27t2 sin(ner) - n cos(nir) _ -
nor
n7r
Thus,
x2+2x-1= 2+ 100(-1)nirx
00
n2 cos - - 20(-1)n
n sin
nirxl
5 5
n=1
Figures (4.12, 4.13, 4.14, 4.15, 4.16, and 4.17) are six graphs of the Fourier series
representation of h(x) for the partial sums Si (x), S5 (x), S25 (x), S5o (x), S75 (x), and
S1oo (x), respectively. Note: The partial sums converge to the function h(x) much
more slowly than the partial sums represented by Figures (4.7,4.8,4.9 and 4.10) of
Example (4.8). The reason for the slower convergence of the partial sums are the x-
values, x = -5 and x = 5. At these x-values h(-5) h(5). Thus, the Fourier series
representation of the function does not seem to converge to the values of h(-5) or
h(5). This example demonstrates the action of the Fourier series representation of
a function with a jump discontinuity at the endpoints.
x
x
Figure 4.12: Fourier Series represen- Figure 4.13: Fourier Series represen-
tation of h(x) = x2 -I- 2x - 1 for the tation of h(x) = x2 + 2x - 1 for the
partial sum 81(x). partial sum S5 (x).
1.
22 r100(-1)n nirx 20(-1)
n
Sl (x) = 3 + L 5J
L
108 Chapter 4: The Essentials of Fourier Series
22 . r100(-1)n
SS(x) - 3 + I,
n=1 L
nirx
n
20(-1)n
5
nirx
25
Szs(x) = 3 +
22 r100(-1)n
L
nix n
20(-1)n nirxl
5J
n=1
X100(- 1)1
cos -
nirx
5
20(-1)n
nor
nirxl
sin-I
30
25
20
15
10
x x
-4 2 4
Figure 4.14: Fourier Series represen- Figure 4.15: Fourier Series represen-
tation of h(x) = x2 + 2x - 1 for the tation of h(x) = x2 + 2x - 1 for the
partial sum S25 (x) . partial sum S5o (x) .
35
30
25
20
15
10
x
4 2 4
Figure 4.16: Fourier Series represen- Figure 4.17: Fourier Series represen-
tation of h(x) = x2 + 2x - 1 for the tation of h(x) = x2 + 2x - 1 for the
partial sum S75 (x) . partial sum Sioo (x)
Sioo(x) = 3 +
22
100
r100(-1)n
L
nix 20(-1)
n 5
nirxl
J
In Figure (4.18), the graph of the Fourier series representation of f (x) over the
interval [-15,15] is shown. We see the Fourier series duplicates itself three times
on this new interval. This is known as the periodic extension of the Fourier series.
Also, we see that the Fourier series representation of 1(x) does converge at f(-5)
and f(5). Although, what it actually converges to is not yet known.
x
15 -10 -5 5 0 15
` 100
Figure 4.18: h (x) = 22
+ Li n=1
100(-1)n
cos - 20(-1)n
n sin 9-] graphed on
the interval [-15,15] .
2.5
1.5
0.5
x
-0.5 0.5
Figure 4.19: The graph of the function g(x) on the interval -1 < x < 1.
Next, we will consider a function that has a jump discontinuity in the interior
of the interval.
EXAMPLE 4.10. Consider the function
1, -1<x<0
g(x) =
2,0<x<1,
shown in Figure (4.19).
110 Chapter 4: The Essentials of Fourier Series
where
/'L 1
ao = 2L
L
j(x) dx = 2 f 9(x) dx = 2
1
3
= sin(nir) = 0,
nir
and
bn L
f Lg(x)sin
(nirx)
dx = 1 J 1 g(x) sin (nlx) dx
1 1 [i_(_i)n]
_ - cos(nir) _
nir nor nor
Thus,
The Fourier series representation of the function g(x) are graphed below for the par-
tial sums Si (x), S1o (x), S25 (x), S5o (x), S1oo (x), and S1ooo (x), respectively. Also,
note that as n increases from 1 to 1000, the graphs of the Fourier series represen-
tation of the function g(x) become increasingly more accurate. Again, please note
that at the x values of -1 and 1, the Fourier series representation of g(x) does not
seem to converge to the points g(- 1) and g(1). Also, at x = 0 the Fourier series
representation of g(x) seems to equal both 1 and 2. In Figure (4.26), the Fourier
series representation of g(x) is graphed over the interval [-3, 3]. Please note that
the Fourier series representation of g(x) repeats itself. Again, this is known as the
periodic extension of the Fourier series.
2.5
1.
0.5
x x
-1 -0.5 0.5 -0.5 0.5 1
1 1
0.5 0.5
x x
-1 -0.5 0.5 1 -1 -0.5 0.5 1
3 3
2.5 2.5
2 2
1.5
0.5
x x
-1 -0.5 0.5 -1 -0.5 0.5
2.5
1.5
0.5
x
-3 -2 -1 1 2
on the interval [-1,1 ] has a jump discontinuity at x = 0 and g(- 1) g(1). Again,
if we consider the graph of the Fourier series representation of g(x) in Figure (4.26),
we see that at the points (-1, g(-1)), (0, g(0)), and (1, g(1)), the Fourier series does
converge to something.
The convergent point of the Fourier series representation of a function 1(x) on
the interval [-L, L] with jump discontinuities within the interval is the average of
the left and right limits of the function at the jump discontinuity. For instance, this
means in Example (4.10) where
1, -1<x<0
9(x)
2,0<x<1,
at x = 0, the left limit of the g(0) = 1 and the right limit of g(0) = 2. Therefore,
the Fourier series representation of g (x) at x = 0 converges to the value
g(0) +2 g(0) =1.5.
The Fourier series representation of a function 1(x) on the interval [-L, L],
where f (- L) f (L) , converges to the average of the left and right limits of the
endpoints on the periodic extension of the function. What is the periodic extension
of a function? The easiest way to answer this question is to think of the cosine
function on the interval [-7r, 7r] graphed in Figure (4.27). A periodic extension of
the cos x on the interval [-7r, 7r] is the cos x graphed on the interval [-37r, 37r] shown
in Figure (4.28).
Figure (4.28) repeats the original curve shown in Figure (4.27), once on the
left and once on the right. This is an example of a periodic extension. Ba-
sically, the curve is repeated on the left of the original curve as many times as
Section 4.3: A New Space: Function Space of Piecewise Smooth Functions 113
x
-2. 2.5 5 7.
0.5
needed to reach negative infinity, and on the right of the original curve as many
times as needed to reach positive infinity. Therefore, in Example (4.9), where
f (-5) f(5), the Fourier series representation of 1(x) converges at x = -5 to
f(-5) + f(-5) _ f(5) + f(-5) _ 34+ 14 = 24 at x = 5 the Fourier series
2 _ 2 _ 2
converges to f (5 ) + f(5) = f (5 ) + f(-5) = 34 + 14 = 24.
2 2 2
Theorem (21) is Fourier's convergence theorem, which was proved by Peter Gus-
tav Lejeune Dirichletl in 1828 in the journal Journal fur die refine and angewandte
Mat hematik.
2. the average of the left and right limits of the function f (x) at any jump dis-
continuity. That is, the Fourier series converges at a jump discontinuity x = xo to
f (xo) + f (x)
2
3. if the endpoints, f (-L) and 1(L) are not equal, then the Fourier Series ton-
verges at -L and at L, to f (L
) + f (-L)
2
The proof of Theorem (21) may be found in many different texts on Fourier
series. In particular, Fourier series, by Georgi P. Tolstov, provides a complete
derivation of this theorem and many of the other theorems governing convergence
of Fourier series.
Since the Fourier series of a piecewise smooth function is guaranteed to converge
to the function whenever the function is continuous, or to the average of the left-
and right-hand limits wherever the function has a jump discontinuity, we do not
have to graph the function or the periodic extension of the function before writing
the Fourier series down and determining the Fourier coefficients. This concept is
very important. We now develop the notion of uniform convergence of a Fourier
series to a function.
Understanding how a Fourier series converges requires some knowledge of series
convergence. Hence, we start with two definitions; absolute and uniform conver-
gence. If you have not had a real analysis course, this definition and several other
concepts may be difficult to understand. However, it gives you an idea of conver-
gence and how it must be shown. So let's get started.
Definition 22. An infinite series is said to be absolutely convergent if the series
formed from it by replacing each term by its absolute value is convergent.
Definition 23. Let (fn) be a sequence of functions defined on a subset S E R.
Then, (fn) converges uniformly on S to a function F defined on S, if for each
> 0 there exists a number N such that
I- f(x)I <E
for allxES and alln>N.
This definition gives us a precise way of determining uniform convergence. The
following theorem, called the Weierstrass M-test,2 provides a way to determine if a
sequence of functions converges uniformly. The proof of this theorem may be found
in Analysis with an Introduction to Proof by Steven R. Lay as well as many other
texts on real analysis.
Theorem 24. (Weierstrass M-test) Suppose that (fn) is a sequence of functions
defined on a subset S E R, and (Ma) is a sequence of nonnegative numbers such
that
fn(x)I <M
2Karl Weierstrass (1815-1897) was one of the greatest mathematics teacher of the mid-
nineteenth century. His greatest contributions to mathematics were in the field of power series
representation of a function.
Section 4.3: A New Space: Function Space of Piecewise Smooth Functions 115
00 00
converges, the Fourier series, Equation (4.29), converges absolutely and uniformly.
Proof. Since
n7rx n7rx n7rx n7rx
an cos L + bn sin L < an cos L + bn sin L
Similar theorems are stated for the Fourier sine and cosine series in the next
section.
EXERCISES 4.3
-L
sin dx = 0, n E N, n m
by direct computation.
4.3.2. Show that
nirx
J(L)
L
-L
cos dx = 0, n E N
by direct computation.
4.3.3. Show that
nirx
J(L)
L
sin dx = 0, n E N
-L
by direct computation.
4.3.4. Show that
nirx mirx
J(L)(L)
L
-L
cos dx = 0, n E N, n m
by direct computation.
4.3.5. Plot the functions 1(x) = x and g (x) = 1 on the interval [-L, L]. Are the
functions 1(x) and g(x) orthogonal? Show by direct computation
pL
f(x)g(x) dx = 0.
J_L
u(x)v(x) dx = 0.
11
Section 4.3: A New Space: Function Space of Piecewise Smooth Functions 117
and
1
2
I [v(x)J2 dx =
u(x)v(x) dx = 0.
and
2
[v(x)] dx=-.
z
11
4.3.8. Some functions are orthogonal with a common weight function. For example,
the Hermite polynomials are orthogonal with weight function ex In Chapter
.
Hl (x) = 2x,
4.3.9. Show, by direct computation, that the Fourier coefficients a0 and bn have the
formulas
L
1
ao =
2L -L
and
1 L n7rx
bn = L
f L
1(x) sin L dx.
4.3.10. Determine the Fourier series for the following functions on the given bounds:
(1) 5for-2<x<2.
(2) 2x2 - 3x for -7r < x < 7r.
(3) 4 cos 3x for - <x < .
2 2
(4)
1 + cos 2x + sin 5x for
3 3 - x <-
4.3.11. Determine the Fourier series representation for the function 1(x) = 2x - 1
on the interval [-5,5]. Using your favorite mathematical software, plot the
function 1(x) and the Fourier series representation of 1(x) for n = 1 to 5,
n = 1 to 10, n = 1 to 50, and n = 1 to 200. In your own words, explain how
the Fourier series representation of 1(x) converges at the points -5, 0, and 5.
4.3.12. Determine the Fourier series representation for the function g (x) = 3 sin x
on the interval [-7r, 7r]. State in your own words any conclusions you may
determine about the Fourier series representation for the function g(x).
4.3.13. Let
1
'
- 2 <- x <
37r
2
Determine the Fourier series representation for the function h(x). Using
your favorite mathematical software, plot the function h(x) on the inter-
- 37r 37r
val 2 , 2 and the Fourier series representation of h(x) on the interval
[-37r, 37r] for n = 1 to 5, n = 1 to 10, n = 1 to 50, and n = 1 to 200. In your
own words, explain how the Fourier series representation of h(x) converges at
3-r
the points - and 0.
2 ' 2
4.3.14. Do you need to solve for the Fourier coefficients before graphing the Fourier
series representation of a function?
Section 4.4: Even and Odd Functions and Fourier series 119
4.3.15. Suppose 1(x) are piecewise smooth functions on [-L, L]. Prove Bessel's in-
equality:
0o L
a2 + a2 + b2 < 1 f[f(x)]2dx.
n= 1
Bessel's inequality implies the sum of the squares of the Fourier coefficients
of any square integrable function always converge.
4.3.16. Show that the Fourier series
00
n7rx n7rx
a0 + an cos + bn sin
n=1
L L
this chapter, Fourier series were described as useful for solving PDEs like the heat
and wave equations. However, Fourier series were developed on the interval [-L, L].
Therefore, there must be some kind of mathematical translation. This translation
is based on even and odd functions.
The definition for an even function is as follows:
Definition 28. The function 1(x) is an even function on the interval [-L, L] if
f (-x) = f (x)
for all x E [-L, L].
x
-L
Definitions (28 and 29) explain how to show a function as even or odd. However,
another way of describing a function as even or odd is graphically. The graph of an
Section 4.4: Even and Odd Functions and Fourier series 121
even function on the interval [-L, L] is symmetric about the y axis. Figure (4.29)
demonstrates this feature.
The graph of an odd function on the interval [-L, L] is not symmetric about
any axis. However, if you perform a 180° rotation of the first quadrant of the graph
of a function, and it is identical to the third quadrant of the graph of the function,
then the function is odd. Figure (4.30) demonstrates this feature.Please note, the
graph of the function in the third quadrant is identical to a 180° rotation of the
graph in the first quadrant.
When two even or two odd functions are multiplied together, the result is an
even function. When an even and an odd function are multiplied together, the
result is an odd function. Also, integration benefits from even and odd functions.
If we integrate an even function, 1(x), from -L to L, then we can simplify the
integral by multiplying it by 2 and integrating from 0 to L, that is,
EXAMPLE 4.11. Show how the function 1(x) = x on the interval [0,2] can be
expanded in a Fourier series.
Solution: There are two choices and both are correct. First, Figure (4.31) shows
the graph of f (x) = x on the interval [0,2]. Next, the function 1(x) is extended
as an even function, which we will call 1(x) _ Ix on the interval [-2, 2]. Figure
(4.32) graphically illustrates the even periodic extension of 1(x) _ x on the interval
122 Chapter 4: The Essentials of Fourier Series
[-2, 2]. The Fourier series representation of 7(x) = x) on the interval [-2, 2] is then
developed. The Fourier series representation of the function 1(x) = x I is
00
Lnirx nrx
J(x) = x) = a0 + an cos + bn sin
n=1
L
00
nirx nirx
= a0 + an cos + bn sin
2 2
n=1
where
2
ao = 12L-fL (x) dx = -1 xi dx.
L 4 -2
However, 7(x) = x l on the interval [-2, 2] is an even function. Therefore,
2 2
a0=
1
4
fIxl dx= fx dx= 1.
2
1
Also,
an = L
J LL f (x) cos
nLx dx = 2J2 x cos n2x dx.
2
nrx lx) cos nrx on the interval [-2, 2] is an even function; therefore,
7(x) cos -i=
2
2 2
1 nrx 4 ((-1)n -1 )
an = 2 lxi cos 2 dx = x cos 2 dx =
-2 0 (nor)
nrx nix
Since 1(x) sin = lx sin on the interval [-2, 2] is an odd function, the
2 2
integral is zero. Thus,
00 4((_1)n_1)
n=1 (nit)
2 cos 2xn
Section 4.4: Even and Odd Functions and Fourier series 123
Figures (4.33,_4.34, 4.35, and 4.36) are four graphs of the Fourier cosine series of
the function 1(x) _ x for the partial sums Si (x), Sio (x), S25 (x), and S50 (x),
respectively. Figure (4.37) is the Fourier cosine series representation of 1(x) _ x
for the partial sum S50 (x) on the interval [-8, 8].
x x
-2
Figure 4.33: Fourier cosine series Figure 4.34: Fourier cosine series
representation of 1(x) _ x for the representation of 1(x) _ x for the
partial sum Si (x). partial sum Sio (x) .
2 2
1.5 1.5
0.5 0.5
+- x x
-2 -1 2 -2 -1
Figure 4.35: Fourier cosine series Figure 4.36: Fourier cosine series
representation of 1(x) _ x for the representation of 1(x) _ x for the
partial sum S25 (x) . partial sum S50 (x) .
S25 (x) = 1 +
25
4((-1)-1) coS nix
2
n=1 (nit) 2
124 Chapter 4: The Essentials of Fourier Series
and
S50 (x) = 1 +
50
4((-1)-1) 2 COS
n7rx
(nor) 2
n=1
The interval [-8, 8] was chosen because it shows that the convergence is exact on
x
-2 -1 1
-2
00
n7rx
f (x) = x = ao + :i: an cos + bn sin
n=1
L L
00
n7rx
= ao + [an cos + bn sin
2 2
n=1
Section 4.4: Even and Odd Functions and Fourier series 125
a0
1
2L -L
L
1(x) dx =
1
4
fx 2
2
dx = 0,
and
1 L n7rx 1 2
an = L 1(x) cos L dx = 2 x cos 2 dx = 0.
-L -2
Whereas,
L 2
1 nrx 1 2 nrx nrx
bn = L
f L
f (x) sin L dx = 2
-2
x sin
2
dx =
0
x sin
2
dx
nor
Thus,
°° -4 (-1)n n7rx
f (x) = x = > sin ,
nor 2
n=1
Figures (4.39, 4.40, 4.41, and 4.42) are four graphs of the Fourier sine series of
Figure 4.39: Fourier sine series rep- Figure 4.40: Fourier sine series rep-
resentation of f (x) = x for the partial resentation of f (x) = x for the partial
sum Si(x). sum Sio (x) .
the function f (x) = x for the partial sums Si (x), Sio (x), S25 (x), and S50 (x),
respectively. Finally, Figure (4.43) is the Fourier series representation of f (x) = x
for the partial sum S50 (x) on the interval [-8, 8].
Si (x) _
1
-4 (-1)n
sin
nrx
and S10 (x) _
10 -4(-l)' sin nrx
n7r 2 n=1 n7r 2
n-1
126 Chapter 4: The Essentials of Fourier Series
2 2
1 1
x
1
-2 -2
Figure 4.41: Fourier sine series rep- Figure 4.42: Fourier sine series rep-
resentation of 1(x) = x for the partial resentation of 1(x) = x for the partial
sum S25(x). sum S50(x).
x
-7.5 -2. 5 2.5
Figure 4.43: Fourier sine series representation of f (x) = x on the interval [-8,8].
25
_4 (-1)n 50 -4 (-1)n nirx
S25 (x) _ sin 2and S50 (x) = sin
n7r n=1 nor 2
n-1
The interval [-8, 8] was chosen because it shows the convergence is not exact
on the interval [-2, 2]. At the endpoints, x = -2 and x = 2, the Fourier series
representation of the function, 1(x) = x, converges to 0.
where
11L
ao = L f (x) dx
2 L n' rx
an = L 1(x) cos L dx.
0
The odd extension is called the Fourier sine series representation of the function,
and it is given by
00
n7rx
1(x) _ bn sin
n=1
L
where
bn =
2 JL f (x) dx.
L 0 L
However, which of the two different extensions is the right one to use? This is
generally answered by the format of the problem. In most cases, either the Fourier
sine series representation will be the correct one to use or the Fourier cosine series
representation will be the correct one to use, but not both.
We conclude this section with two theorems on when the Fourier sine and cosine
series representation of a function are continuous. As in Theorem (27), they follow
directly from Theorem (26) stated at the end of the last section. These theorems are
important and will be continually used throughout the rest of the text particularly
in Chapter 6 and 8.
Theorem 30. If 1(x) is a piecewise smooth continuous function on the interval
[0, L], then the Fourier cosine series representation of the function 1(x) is contin-
uous.
EXERCISES 4.4
4.4.1. Determine the Fourier cosine series for the following functions on the given
interval:
(1) 7for0<x<2ir.
(2) sin x for 0 < x < 0.5.
128 Chapter 4: The Essentials of Fourier Series
subject to
4.4.7. Find the Fourier cosine series representation for the function 1(x) = x2 on the
interval [0, 1]. Using your favorite mathematical software, plot the function
1(x), the even extension of 1(x), and the Fourier cosine series representation
of the function 1(x). Determine the necessary n, so that the Fourier cosine
series representation of the function 1(x) is accurate.
4.4.8. Find the Fourier sine series representation for the function 1(x) = x2 on the
interval [0, 1]. Using your favorite mathematical software, plot the function
1(x), the odd extension of 1(x), and the Fourier sine series representation of
the function 1(x). Determine the necessary n, so that the Fourier sine series
representation of the function 1(x) is accurate. Compare your results with
those of the previous problem.
4.4.9. Is the Fourier cosine series representation of the function g(x) = x3 on the
interval [0, 3], or the Fourier sine series representation of the function g(x) _
x3 on the interval [0, 3], equivalent to the Fourier series representation of the
function h(x) = x3 on the interval [-3, 3]?
4.4.10. Given the fact that 1(x) is a continuous function on the interval [-L, L],
(1) state the conditions when the Fourier series representation of the function
1(x) is equal to the function 1(x) for all x in the interval,
(2) state the conditions when the Fourier cosine series representation of the
function 1(x) on the interval [0, L] is equal to the function 1(x) for all x
in the interval [0, L], and
(3) state the conditions when the Fourier sine series representation of the
function 1(x) on the interval [0, L] is equal to the function 1(x) for all x
in the interval [0,L].
Chapter 5
Separation of Variables:
The Homogeneous Problem
5.1 INTRODUCTION
So far, we have discussed both the steady-state temperature solution for the dis-
tribution of heat in a rod and d'Alembert's solution for the one-dimensional wave
equation. In this chapter, we introduce a third method for solving PDEs: separation
of variables.
Separation of variables is an important technique to master when studying solu-
tion methods of PDEs, because this technique leads to an infinite series solution, the
Fourier Series solution. However, separation of variables does have its drawbacks.
First, unlike the steady-state temperature solution, which makes no requirement on
the homogeneity of the PDE or boundary conditions (BCs), separation of variables
requires both the PDE and the BCs be homogeneous. Second, in general the spa-
tial variable must have finite boundaries. If the spatial variable has semi-infinite
or infinite boundaries, separation usually does not work. We tackle the problem of
semi-infinite and infinite boundaries in Chapter 11. Finally, separation of variables
requires the PDE and the BCs to be linear. Therefore, this chapter starts with
a brief discussion of linear and homogeneous equations and BCs. To discuss this
properly, we introduce the notion of an operator (don't get nervous, you've already
seen lots of them). Operators are very important to mathematics and are used
extensively by physicists and engineers.
Once we complete our discussion on operators and linear equations, we proceed
with separation of variables. First, we cover separation of variables and the heat
equation, followed immediately by separation of variables and the wave equation.
The similarities between the two sections is quite evident and should be no surprise.
Then, we apply separation of variables to multidimensional spatial problems. We
complete the chapter with another application of separation of variables, Laplace's
equation in Cartesian coordinates.
131
132 Chapter 5: Separation of Variables: The Homogeneous Problem
/' b
(x + 2x -
a
In the study of ODEs, you learned that a differential operator can look like
d2- d_
a b,
dx2 + dx +
which is the sum of other operators. Again, III is the vector space that is involved
with the two calculus operators and the differential operator.
Judging from these three operators, it would seem reasonable to assume a partial
differential operator could be
a_ a2_
at - ax2
In fact, Equation (5.1) is a partial differential operator known as the heat operator.
Actually, an operator is like a function. However, it is more versatile. An opera-
tor is a mapping of vector spaces or function spaces. This is important to remember
because the objects that operators operate on are, in this course, functions.
Now that we have looked at several different examples of operators, let's discuss
a very special type of operator: the linear operator.
Note: Vector space in the previous definition may be replaced by function space.
b
For example, we can show that L = Ja _ dx is a linear operator. This means
that for two different functions 1(x) and g(x) and two constants c and d, we have
b
By the property of definite integrals, we know the integral of the sum of two func-
tions is equal to the sum of the integrals of each function. Thus, we have
b b b
f [cf(x) + dg(x)] dx = c f (x)dx + dg(x)dx.
a a a
Applying the property of definite integrals again, we find that the integral of the
product of a constant and a function is equal to the product of the constant and
the integral of the function, resulting in
02(ui)
=c1 -cig(x,t)(ui)+
Ox2
134 Chapter 5: Separation of Variables: The Homogeneous Problem
2
-c2g(x,t)(u2).
a2
Let's start with the left side of this equation and work to the right side. We have
the equation
l a(u1)
at
a(u2) _
at
lka2(ui) _ 2ka2(u2) _ c,g(x, t)ul- c2g(x,
ax2 ax2
t)u2.
Cl
a("') +CZ a(u2) - a2(u') -cZk a2(u2)
- cig(x, t)ui - c2s(x, t)u2
at at ax2 ax2
2
a(ui)
C1
ax2
-cig(x,t)(ui)+
at
2
a2(u2)
C2 aat ) 2) - cZg(x, t)(u2),
L=
at
-k-x2 - 2
9(x, t)-
is a linear operator.
Understanding a new concept is often easier if a counterexample is given. Con-
sider the following example, also based on the heat operator, a nonlinear heat
operator.
EXAMPLE 5.2. Consider the operator
- () 2
Oax2 + ax
3
- at
Section 5.2: Operators: Linear and Homogeneous Equations 135
For the operator J, the parentheses, (), indicate a holding place for the function
being operated on. To prove that J is a nonlinear operator, we show that
a(C1u1 + C2u2)
- (c1u1 + c2 U2) a2(C1u1 + C2u2) a3(C1u1 + C2u2)
at ax2 ax3
3(u1)
____
at
- 11 2(u1)
Dx2 ax3
+Cla3(ui)
ax2 ax2 ax3 ax3
No matter what algebraic method we apply, we can not rearrange this to be
Cl a(u1) _ Clul a2(u1) + a3(u1)
Cl
at ax2 ax3
a(u2)
+CZ
at ax2 ax3
Thus, J is a nonlinear operator.
Now, having the skill to identify a linear operator only comes in handy if you
understand its placement in a linear equation.
136 Chapter 5: Separation of Variables: The Homogeneous Problem
au(x,t) - Q(x,t).
Here, the linear operator is a partial differential operator giving us a linear PDE.
Another example is the wave equation with a damping function, Q(x, t),
2u(x,t) a2u(x,t)
ate
-c axe
=Q(x,t).
) y) + d(x, y) y)
a(x,y)a ux'y)
z + b(x, y) a ayax + c(x, y) a uay ,
Ou(x, y)
-- a(.r.. ail + f(x, y)u(x, y) = g(x, y).
ay
A nonlinear PDE is one where one or more of the functions a(x, y), b(x, y),
c(x, y), d(x, y), e(x, y), or f (x, y) in Equation (5.3) is a function of u(x, y). For
example, consider the equation
y)
8x 8y8x 8y 8x
Ou(x, y)
+ e(x, ii) + f(x, y)u(x, y) = g(x, y).
ay
It is a nonlinear equation because a(x, y) in Equation (5.3) has been changed to
u(x, y). For another example, consider
02u(x,
u(x, t) t) - CZ a2u(x, t)
= Q(x, t).
ate axe
02u(x, t)
This equation is nonlinear, since we have the product u(x, t)
t2 Note: The
study of nonlinear PDEs is beyond the scope of this text. However, you still need
.
Boundary conditions can also be defined as linear and homogeneous. The special
au(0, t) _ 0
BCs studied in Chapters 2 and 3 where u(0' t) = 0 and u(L' t) = 0 or
ax
au(L, t)
and = 0 are obviously linear and homogeneous. It is more difficult to
ax
au (0, t) au (L, t )
determine whether Robin's conditions (where = - hu(0 t) and
ax ax
hu(L, t), h is a positive constant) are linear and homogeneous. To determine if BCs
are homogeneous, set u(x, t) - 0 in the BCs. If you get 0 = 0, then the BCs are
homogeneous. To determine if BCs are linear, consider whether there is a product
of the function u(x, t) with the operator on the function u(x, t). For instance,
au(o,t) __ a2u(0,t)
ax -u(0' t) axe
is a nonlinear BC, whereas
au(o, t) _ _hu(o, t),
8x
138 Chapter 5: Separation of Variables: The Homogeneous Problem
EXERCISES 5.2
ate axe +
(a) Identify the operator.
(b) Show that the operator is linear.
(c) If ,3(x, t) = 0, what can you say about the equation?
5.2.3. Let
8u 8u
L(u)=u--k--.
8x
au
Note: Simply stating that there is a product between u and is not enough.
at
5.2.4. Show that
iaral ia2
L r 8r Lr ar + r2 ae2
is a linear operator.
5.2.5. Consider the equation
a2u 1 a / aul 1 a2u a2u
8t2 r ar r 8r + r2 ae2 + 8z2 .
(a) Identify the operator.
(b) Show that the operator is linear.
5.2.6. Consider the equation
au _ 1 a 2 au 1 a2u 1 a2 u
sin 0 ae2 + r2 .
at r2 ar r ar + r2 sin o
sin2
0 acb2
Section 5.3: Separation of Variables: Heat Equation 139
We want to have a solution for Equation (5.5) subject to the BCs, Equation (5.6),
for any time t and 0 < x < L. This is called solving the initial value problem.
To do this, we first assure ourselves that Equations (5.5 and 5.6) are linear and
homogeneous, using the tests described in the previous section. If Equations (5.5
and 5.6) are linear and homogeneous, we can apply the separation of variables
technique, and we assume the function u(x, t) is a product of two functions, one of
time G(t) and one of space p(x). In other words
u(x, t) =
Then we find the appropriate derivatives. We have
au
= G' (t)'p(x)
and
Separating Equation (5.8) so that one side of the equation is in terms of t and the
other side is in terms of x is called separation of variables. After doing so, we have
G'(t) _ p"(x)
5.9
kG(t) p(x)
This may be very difficult to satisfy, since we are saying that a function of time is
equal to a function of space. Or, more precisely, the left side of Equation (5.9) is a
function of time, t, and does not vary with the variable x. However, it is equal to a
function of x, which does not vary with time, t. Thus, both sides of Equation (5.9)
must be equal to the same constant. The constant is called the separation constant,
and we will denote the constant as -A. Thus, we have the equation
G'(t) - o"(x) - (5.10)
kG(t) p(x)
The choice of -A as the separation constant is for convenience. We could have just
as easily have chosen A.
(tt)
Having separated Equation (5.5) into time _ _A) and space
G
G/
" (x) _ -A equations, we now separate the BCs, Equation (5.6). Using our
\o(x)
assumption for u(x, t) = G(t)p(x), we find that
u(0, t) = G(t)p(0) = 0. (5.11)
If G(t) = 0 for all time t, then the solution to Equation (5.11) is trivial. That is,
u(x, t) = 0, which is not a very interesting problem. But since a BC usually refers
to a spatial variable, it makes sense to assume that
'p(0) = 0.
Similarly
u(L, t) = G(t)cp(L) = 0
implies that
p(L) =0.
Thus, our BCs, Equation (5.6), have become
(,p(0) =0
J (5.12)
(L) = 0.
subject to the BCs, Equations (5.12). Thus, we have completely separated Equation
(5.5) subject to the BCs, Equation (5.6). Note: The initial condition for the time
problem will be applied at a later stage of the solution process. Therefore, it does
not need to be separated.
We will solve the spatial equation first, though, in general, either way will suffice.
The second, more time-consuming, method would be to solve the time equation first,
but then you'd have to fill in the variable A after solving the spatial problem. Also,
it is important that you develop a standard methodology that always works when
solving PDEs using separation of variables. This helps when we start solving much
more complicated problems using this method.
Case 1: A <0.
In this case it is convenient to assume that A _ -s, where s > 0. Thus, Equation
(5.14) becomes
gyp" (x)
= sp(x),
subject to
p(0) =0
(L) = 0.
We now have a second-order constant coefficient ODE where the coefficient is s > 0.
The solution to this problem is
o(x) = c1e"x + c2e-`fix.
and
eax + e-ax
cosh ax = 2
Following are the graphs of cosh x, in Figure (5.1), and sinh x, in Figure (5.2).
Since sinh (0) = 0 and cosh (0) = 1, we must have c3 = 0. Applying o (L) = 0 yields
o (L) = 0 = c4 sinh L;
since the sinh (x) is 0 only when x = 0, we must have c4 = 0. Therefore, when A <0
there is only the trivial solution to Equation (5.14) subject to the BCs, Equation
(5.12).
Case 2:A=0.
Here Equation (5.14) becomes
o" (x) = 0.
Case 3: A> 0.
Equation (5.14) is a second-order ODE with constant coefficients and a solution of
co(x) = hoe-i'x
and
eiax + e-iax
cos ax = 2
Since sin 0 = 0 and the cos 0 = 1, we have co(0) = 0 = h5. Applying the second BC,
o (L) = 0, we have
o (L) = 0 = h6 sin V L .
We could always claim that h6 = 0, but this would only give us the unusable trivial
solution again, co(x) = 0. Hence, we must look for other possible solutions. The real
clue comes from considering the function y = sin x shown in Figure (5.3). We see
x
0
the family
u(x, t) = G(t)cp(x),
u(x, t) = sin n = 1, 2, 3, .. .
L
Since cn and hn are arbitrary constants being multiplied together, we can replace
them by B. Therefore,
t 27rx
u(x, t) = Bee_
sin L ,
Recall that, in Chapter 4, we learned that any piecewise smooth function can be
represented by a Fourier series for 0 < x < L. Therefore, we can solve for the
coefficients Bn by using the orthogonality of the sine function. They are
L
B=
n
fp f (x) sin (nLx) dx
= f(x)sin-_) dx.
L
fsin2()dx L JoL nLxC
This completes the solution of Equation (5.5) subject to the BCs, Equation
(5.6), and IC Equation (5.7). It describes temperature distribution in the one-
dimensional rod for all time t. In fact, if we take the limit as time goes to infinity
in Equation (5.16), we get the steady-state solution for Equation (5.5). We state
this mathematically as
roo
_ 'n t nurx
lim u(x, t) lim Bne sin =0.
t-+00 t-+00 L
n=1
A fully worked example of the separation of variables method for the heat equa-
tion is provided in Example (5.3).
EXAMPLE 5.3. Find the time-dependent solution for u(x, t) when
au _ a2 u
at -
1 .14 axe (5.17)
subject to
u(o, t)= o
8u(1, t)
8x
=0
and
The PDE, Equation (5.17), and BCs, Equation (5.18), are linear and homogeneous.
Therefore, separation of variables technique is valid. Letting u(x, t) = G(t)(p(x),
yields the time equation
G'(t) = -1.14AG(t), (5.20)
When solving the spatial equation, we must determine the valid values of A. Thus,
we first assume A <0. In this case, let A = -s and solve
p "(x) = scp(x),
The first BC, cp(0) = 0, applied to Equation (5.25) yields c5 = 0. The second BC,
p' (l) = 0, indicates either c6 = 0, in which case we only have the trivial solution,
or the eigenvalues
(2n-1)ir 2
An= 2 ,n=1,2,3, .. (5.26)
L
Substituting the eigenvalues from Equation (5.26) into Equation (5.20), we deter-
mine the solution for the time equation, which is
Gn (t)
= dne-1.14Ant
2 1 (2n - 1) rx 32
bn = - (-x 2+ 2x) sin dx =
1 0 2 r
(2n -1)33
Replacing the values for bn in the Fourier sine series yields the specific solution for
u (x, t) , which is
°° 32 (2n-1)7rx
u(x, t) _ I e-1'14ant sin
n-1 (2n - 2
Section 5.3: Separation of Variables: Heat Equation 149
A graph of the solution is shown in Figure (5.4) for the partial sum of n = 25 with
0 < t <2. Also, the initial temperature distribution is outlined on the graph by a
boldface line. It is interesting to note from Figure (5.4) how fast e-1'14ant reduces
the initial temperature distribution to the approximate steady-state solution.
EXERCISES 5.3
0
(4) Bcs: and IC: u(x 0)= x3 - - x2
2
=0
5.3.3. Consider the PDE
au(x, t) a2u(x, t)
at axe
150 Chapter 5: Separation of Variables: The Homogeneous Problem
(c) Use the first 25 terms in the series for u(x, t) to find the approxi-
mate temperature at x = 7 cm when t = 12 secs for each material.
Compare this answer with the previous two answers.
(3) Use the first five terms of the series to determine the amount of time,
t, it takes for each material to reach a temperature of 0°C at the center
of the rod. Compare the amount of time it takes for each material, and
write a short essay on the reasons why the time is different or the same.
5.3.6. Consider a uniform one-dimensional rod of length L, without an internal heat
source, which is not laterally insulated. (Heat can flow in and out across
the lateral boundary.) By experimentation, you discover that heat is flowing
across the lateral boundary at a rate proportional to the difference between
the temperature u(x, t) and the surrounding medium that is kept at 0°C.
(1) Given that the convection constant of proportionality is greater than 0
and the ends of the rod are not insulated and held at 0°C, set up the
mathematical model.
(2) Suppose the initial temperature of the rod is x when 0 -< x < L
2 and
(c) the heat energy flowing out of the rod per unit time at each end,
and
(d) the relationship between parts (b) and (c).
(3) Suppose that after Part (5.3.3) has completed, the heat source is turned
off.
(a) Find the equations that describe the mathematical model at this
time.
(b) Using the equations that you just found, solve the heat flow problem
for any time t.
(c) Using your favorite mathematical software, graph the solution for
0<t<5.
5.3.13. Diffusion through a Membrane: Fick's Law can be used to derive an analogue
of Ohm's Law for a membrane of thickness, L, with different chemical concen-
trations on each side of the membrane. If the medium is isotropic (diffusion
occurs the same regardless of the direction of the measurement), then we get
ac a2 c
at - D axe'
subject to
c(0, t) = C1, the chemical concentration on the left of the membrane
and
c(L, t) = Cr, the chemical concentration on the right of the membrane.
(1) Find the time-dependent solution.
(2) Show that as we let t -- oc, we get the same steady-state solution as in
Chapter 2 Section 2.5.1
As in the previous section, we want a solution for Equation (5.28) subject to the
BCs, Equation (5.29), for any time t and 0 < x < L. To obtain the solution, we
must assure ourselves that Equations (5.28 and 5.29) are linear and homogeneous
using the techniques developed in Section 2. If Equations (5.28 and 5.29) are linear
and homogeneous, we again assume the solution function u(x, t) can be written as
the product of two functions, one of time G(t) and one of space p(x). In other
words,
u(x,t) = G(t)(x).
We then find the appropriate second partial derivatives,
a2u
= U" (t)cp(x)
ate
and
a2 u
.
ax 2
Now substituting into Equation (5.28), we obtain
G"(t)5p(x) = (5.31)
Equation (5.31) is now ready for separation of variables, where one side of the
equation is in t and the other side of the equation is in x. Again, we'll set the
equation equal to a separation constant, as we did in the previous section, resulting
in
G"(t) - cp"(x) - (5.32)
c2G(t) p(x)
As in the previous section, we have separated our PDE into two ODEs, one
ODE in time and one ODE in space. Now let's separate the BCs, Equation(5.29).
Section 5.4: Separation of Variables: Wave Equation 155
u(0, t) =
G(t) = 0 for any time t, then our solution is trivial. Also, a BC usually refers to
a spatial variable; therefore, it makes sense to assume that
(0) = 0.
Similarly,
u(L, t) = G(t)cp(L) = 0,
implying that
cp(L) =0.
0
(5.35)
(L) = 0.
We have now completely separated Equation (5.28) and BCs, Equation (5.29),
into the boundary value problem
(5.36)
G"(t) = -Ac2G(t).
Note: The initial conditions will be applied after a general solution to the problem
is found. Thus, they do not need to be separated.
CL)
n = 1,2,3,...
n'rx
Pn (x) = hn sin L
156 Chapter 5: Separation of Variables: The Homogeneous Problem
n t) = an cos n = 1,2,3,...
L L
Since an, bn, and hn are arbitrary constants and hnan and hnbn are the products,
we will let An = hnan and Bn = hn b. Thus, we have for n = 1, 2, 3, ... ,
nlrx
u(x, t) = LAn cos
L
+ Bn sinL ()] sin L
Again the question remains, which n gives the solution we really want? We know
that for n = 1, we have
rrct\
u(x, t) = I Al cos I Lt I + Bl sin sin (--)
L \ / \ /J
When n = 2, we obtain
/2irct\ 2irct
r
u(x, t) = I AZ cos I L I + BZ sin
L \
L sin2irx
L
which is different from the case n = 1. When n = m, we have
which is different from the previous two cases. Again, we have an infinite number of
independent solutions for u(x, t). By the principle of superposition, we know that
Section 5.4: Separation of Variables: Wave Equation 157
Thus 1(x) is set equal to a Fourier sine series. Now to find An, we use the orthog-
onality of the Sine function, resulting in
=
ff(x)sin()dx =f f (x)
L
L
L
sin n7tx
L dx.
fo sin2 ()dx 0
To satisfy the second IC in Equation (5.30), which is the initial velocity of the
string, we set
au(x, 0)
= >BTh (
Just like f (x), g(x) is set equal to a Fourier sine series. Now, to find Bn, we again
use the orthogonality of the sine function, resulting in
1L
Bn
CL/cL 0
9(x) sin (nLx) dx.
This completes the solution of Equation (5.28) subject to the BCs, Equation
(5.29), with ICs, Equation (5.30). We provide the following example for the wave
equation.
subject to
8u(0, t) _
8x (5.38)
u(1, t) = 0
and
u(x, 0) = x3 + 2x2 - 3
(5.39)
8u(x, 0) _ x-
2 1.
at -
Find a time-dependent solution for u(x, t). Since the PDE in Equation (5.37) and
BCs, Equation (5.38), are linear and homogeneous, separation of variables technique
can be applied. Assuming u(x, t) = we have the spatial equation
subject to
,d(0) = 0
(5.41)
Working with the spatial equation first, it can be determined that for A < 0 and
A = 0 there is only the trivial solution. For A> 0, we have
G( t) = -4A Gn (t),
Section 5.4: Separation of Variables: Wave Equation 159
where An = ancn and Bn = bncn. Applying the IC, u(x, 0) = x3 + 2x2 - 3, yields
00
(2n - 1) rx
u x 0 = x3 2x2 -3= A cos
2
n=1
1
(x3 (2n -
An = 2 2x2 - 3) cos dx
2
2 (2n - 1)irx
B= (2n - 1)
1
f (x2 - 1) cos 2 dx
(5.45)
(2n - 1)44 '
Therefore, the specific solution for u(x, t) is
°O (2n -
u(x, t) _ {An cos [(2n - Bn sin [(2n - cos 2
n=1
where An and Bn are given by Equations (5.44 and 5.45) respectively. Figure (5.5)
is the graph of u(x, t) for the partial sum of n = 15 with 0 < t < 5. Please look
carefully at Figure (5.5). The boundary at x = 1 is fixed at 0. Whereas, the
boundary at x = 0 is freely oscillating. This is exactly what the BCs indicated.
Also, the initial displacement of the string is indicated by a boldface line.
The one-dimensional problems are useful for developing Fourier Series Solutions
to a PDE. However, more interesting problems arise in the multidimensional spatial
variable PDEs, which are discussed in the next section.
EXERCISES 5.4
5.4.1. Determine if the following PDEs are separable. If so, separate the PDEs into
appropriate ODEs. If not, explain why.
C72u(x, t) a2u(x t)
(1) =C 2 v +2l(x t) , .
ate axe
a2u(x, t) 2 t
mau'x, t>
m ax +u(x , t)
(2)
at2 - c ax 2 - .
(3) c(x)p(x)
a2u(x, t) _ a2u(x, t)
ate - 9 axe
For each set of BCs and ICs, solve the initial value problem. Using your
favorite mathematical software, graph the solution for 0 < t < 2. Clearly
indicate the initial displacement on your graph.
u(0, t) = o u(x, 0) = rx -
(1) BCs: and ICs:
8u(x, 0) _ 0
u(rr, t) = 0
at
Section 5.4: Separation of Variables: Wave Equation
5.4.4. A uniform string with mass density 0.03 lbs/ft, and tension 300 lbs, is fixed
at the left end and has a freely moving right end. The string has length 20 ft
and is initially at rest, with linear displacement from 0 to 1.
(1) Model this problem mathematically. Note: You must model and explain
your choice of boundary conditions.
(2) Solve the mathematical model.
(3) Write the solution with all known quantities substituted into it.
(4) Check to see that this solution satisfies the equation, the boundary con-
ditions, and the initial conditions.
(5) Using your favorite mathematical software, graph the solution for 0 <
t <5.
5.4.5. Consider the following information:
(a) a perfectly flexible string of length 2ir ft,
(b) tension of 50 lbs/ft,
(c) mass density of 0.02 lbs/ft,
162 Chapter 5: Separation of Variables: The Homogeneous Problem
5.4.6. Consider a slightly damped vibrating string with a restoring force that satisfies
a2 u _ 252u au
ate c axe - a - /3u, where a and 3 are constants.
at
(1) Explain why a > 0.
(2) Explain the action of the restoring force.
(3) Find a series solution subject to
u(x,0) = f(x)
u(o> t)= o
BCs: J and ICs:
8u(x, 0)
u(L, t) = 0
l at
Explain your choice of the magnitude of a and 3 as you solve this prob-
lem.
5.4.7. A uniform string with fixed ends is excited by the impact of a rigid plane
hammer, which gives it the following initial distribution of velocities:
8u(x, 0)
at - 225 , 7r < x < 27r,
Find the vibrations of the string, if the initial displacement was zero.
Section 5.4: Separation of Variables: Wave Equation 163
5.4.8. A uniform string with fixed ends is excited by the impact of a rigid sharp
hammer, which gives it the following initial distribution of velocities:
0' 0 < x < 7r
au x 0
(
at
) = 225 cos(x - r )' it-2 <x < 3ir2
2
-3ir<x
0'2 - <2r.
-
Find the vibrations of the string, if the initial displacement was zero.
5.4.9. Consider longitudinal vibrations of a uniform flexible rod with free ends. De-
termine the following:
(1) The mathematical model if the initial displacement and velocity are ar-
bitrary functions of x in the longitudinal direction.
(2) Solve the initial value problem.
Note: Consider the possibility of uniform linear motion of the rod for
the entire problem.
5.4.10. A uniform string with fixed ends has an initial displacement of 2x for 0 <
x < it and 3ir - x for 7r < x < 3ir. The initial velocity is zero and the string
is vibrating in a medium that resists the vibrations. (The medium produces
a resistance proportional to the velocity.) Suppose the resistance constant of
proportionality is 0.01. Find the solution to the initial value problem.
5.4.11. A uniform string with a fixed end at 0 and free end at 2ir has an initial
3zr 3zr
displacement of -x for 0 < x < and 3x - 67r for < x < 27r. It is known
- 2 2 - -
that the initial velocity is zero and the string is vibrating in a medium that
resists the vibrations. (The medium produces a resistance proportional to the
velocity.) Suppose the resistance constant of proportionality is 0.03; find the
solution to the initial value problem.
5.4.12. A uniform string with free ends has an initial displacement of
-x,0-<x<-1
u(x,0)= x-,1x<3
-2x+8,3<x<4.
The initial velocity is zero, and the string is vibrating in a medium that
resists the vibrations. (The medium produces a resistance proportional to the
velocity.) Suppose the resistance constant of proportionality is 0.13; find the
solution to the initial value problem.
5.4.13. This problem develops the mathematical model for longitudinal vibrations of
a gas in a tube. Consider an ideal gas performing small longitudinal vibrations
164 Chapter 5: Separation of Variables: The Homogeneous Problem
8x
V=
a
L ay J
for a two-dimensional system. Thus, V2 = V V or
rai
ax
a
ax a2 a2
V2=V.V= =ax2 a2 ,
5115
ayJ LayJ
Section 5.5: The Multidimensional Spatial Problem 165
a2 a2
V2 = + ay2
ax2
and IC
Equation (5.49) and BCs, Equation (5.50), are linear and homogeneous. Hence,
the separation of variables technique can be applied as a solution method for this
problem. As in the one-dimensional problem, we assume that
1 dG ia2 a2
kG dt cp (c9x2 +
where -A is the separation constant. Also, the BCs are separated and become
p(0, y) =0
p(L, y) =0
8cp(x, 0) (5.52)
0
8y
8cp(x, H)
0 .
8y
dG
= - AkG
dt '
(5.53)
d2X d2Y
Y + X d 2 = - AX Y. (5.54)
dx2 y
l d2X _ l d2Y _
-T,
X dx2 - Y dy2
where -T is the new separation constant. Again, we must separate the boundary
Section 5.5: The Multidimensional Spatial Problem 167
H)
Ii dy
0.
(5.56)
X(L)
(L) = = o0;
the other in terms of Y(y),
d2Y
(5.57)
y
subject to the BCs
0
dy
(5.58)
dY(H
dY(H)
0.
d dy
and eigenfunctions
X (x) = bn sin n= 1,2,3,...
L
Thus, the complete solution for the spatial problem for X (x) is
2
Tn =
\L)
,n=1,2,3,...
nlrx
Xn (x) = bn sin
(-z)
168 Chapter 5: Separation of Variables: The Homogeneous Problem
d2Y
dy2 = -Y,
subject to the BCs, Equation (5.58); from previous work, we recognize that the
eigenvalues are
mir 2
m=
(-)
H
,m=0,1,2,3,... (5.62)
m=0,1,2,3,...
Ym,(2,/) = p,m, cos H
C may I
and
miry
Ym(y) = am cos , m=0,1,2,3,...
H
Therefore, we have for n = 1,2,3,... and m = 0,1,2,3,...,
nirx miry
(p(x, y) = Cnm sin L cos H , (5.65)
where
Cnm = bn am .
Also u(x, y, t) = Gnm(t)cpnm(x, y) Thus, using Equations (5.64 and 5.65) we get
for n = 1,2,3,... and m = 0,1,2,3,...,
nirx miry
u(x, y, t) = (5.67)
00
2irx may
- sin A2me -A 1mkt cos .
L
m=o
H
170 Chapter 5: Separation of Variables: The Homogeneous Problem
+ ... + sin
njrx Anme-Anmkt cos may -+... (5.71)
L
m=o
H
But since equation (5.71) is an infinite sum of an infinite sum, it could be easily
written as
00 00
n?rx may
u(x, y, t) = sin cos
n=1
L m=0 H
00
nirx
= Ano e -An0kt sin
L
n=1
00 00
miry nirx
+ Anme coS -fl- sin , (5.72)
L
n=1 m=1
This leaves us the task of determining the equations for the coefficients Anm and
Ano
Let's first determine the equation for the coefficient Anm. In other words, we
consider the case when n = 1, 2,3,. . . and m = 1, 2, 3, ..., which is
ux ,y )=f( ,y)=(5.73)
0 x
00
n=1 m=1
00
A cos
may
H
sin
L
Note: To do this we are actually applying the orthogonality of the sine and cosine
functions.
Section 5.5: The Multidimensional Spatial Problem 171
In Equation (5.73), the inner sum actually equals a function in y for each value
of n. That is,
00
miry
B(y)= = >Anmcos--.
H
(5.74)
m=1
Since f (x, y) is now set equal to a Fourier sine series, we can use the method that
we learned in Chapter 3 to solve for Bn (y). The solution is
nirx
B(y)=-L2 L
f (x, y) sin
L
dx.
0
Thus, we have Bn (y) equal to a Fourier cosine series. Next, solving for Anm yields
H
2 miry
Anm = H Bn (y) cos dy,
0 H
and on replacing Bn (y) with its integration equation, we get
2 H2 L nirx miry
Anm f (x, y) sin L dx cos dy
H 0 L 0 H
4 HL miry nirx
f (x, y) cos sin L dxdy. (5.75)
HL 0 0 H
Likewise we arrive at
H L
_ 2 nirx
f (x, y) sin L dxdy. (5.76)
Ano LH J 0 0
This completes the answer to the problem at the beginning of this section, giving
us
00
nhrx
u(x, y, t) = Anoe sin
n=1
L
00 00
miry nirx
+ Anm e - fin" cos sin
n=1 m=1
H L
where
2
HL nirx
Ano = f (x, y) sin L dxdy
LH 0 0
172 Chapter 5: Separation of Variables: The Homogeneous Problem
and
H
4 L miry nirx
Anm = HL f (x, y) cos sin L dxdy.
0 0 H
This completes the solution of Equation (5.49), subject to the BCs, Equation (5.50),
with IC, Equation (5.51). See below for a worked example for a multi-dimensional
problem.
subject to
u(0, y, t) = 0
u(2, y, t) = 0
(5.7s)
u (x, 0, t) = 0
u(x, 3, t) = 0,
with
0<x<1 0<y<1.5
0<x<1 1.5<y<3
u(x, y, 0) = f (x, y) (5.79)
1<x<2 0<yG1.5
1<x<2 1.5<y<3,
and
au(x, y, 0) _ 0
at
The PDE, Equation (5.77), and BCs, Equation (5.78), are linear and homogeneous.
Therefore, separation of variables technique is valid. Letting u(x, y, t) = G(t)Sp(x, y)
and substituting into Equation (5.77) and Equation (5.78) yields the time equation
G"(t) = -9AG(t),
and the spatial equation
y) y)
axe + aye ) _ -ap(x, y), (5.82)
Section 5.5: The Multidimensional Spatial Problem 173
The spatial equation, Equation (5.81), subject to the BCs in Equation (5.83), is a
linear and homogeneous system. Hence, separation of variables technique is again
valid. Letting Sp(x, y) = X (x)Y(y), and substituting into Equation (5.81) and
Equation (5.83), yields the spatial equation
X"(x) _ -TX (x), (5.84)
subject to the BCs
X(0) = o
J (5.85)
X(2) = o.
We know from Section 5.3, Equation (5.84), subject to the BCs, Equation (5.85),
has a nontrivial solution only when 'r> 0. For 'r> 0, the eigenvalues are
n?C
'rn= (---)
2
2
,n=1,2,3,... (5.86)
(5.89)
Y(3) =0 .
Again, we know from Section 5.3, Equation (5.90), subject to the BCs, Equation
(5.89), has a nontrivial solution only when > 0. For > 0, the eigenvalues are
m?t 2
gym= 3 ,m=1,2,3,..., (5.91)
174 Chapter 5: Separation of Variables: The Homogeneous Problem
2 +(-)
3 ,
(5.93)
We now can put together the total solution for u(x, y, t). Remember Sp(x, y) _
X (x)Y(y), so we have
u(x, y, t) = Spmn (x, y)Gmn (t) = Xn (x)Ym (y)Gmn (t) _
00
n=1 m=1
00
u(x,y,0) = f (x,y) =
00 00
n=1 m=1
A mn sin
miry
3
sin
-
nirx
2
(5.96)
96 mir nit
sin sin (5.97)
m2 n2 ir4 3 2
au(x, y, 0) = 0
Repeating the process with the second IC , , Y fields
at
00 00
au(x, y, 0) nirx
at
=0 = [3/Bmn] Sin miry
3
Sin
2
n=1 m=1
which implies Bmn = 0. Therefore, the complete solution is
00 00
miry nirx
u (x, y, t) _ Amn cos 3 Amn t sin sin
3 2
n=1 m=1
where Amn is given in Equation (5.97). Since we can't present four-dimensional
graphs, we graph the solution for the partial sums n = 25 and m = 25 and several
different times, t. Figure (5.6) depicts the surface u(x, y, t) at time t = 0. Here,
the four faces of the figure indicate the initial conditions. Figure (5.7) shows the
surface of u(x, y, t) at time t = 0.2. The third graph, Figure (5.8), portrays the
surface u(x, y, t) at time t = 0.4. Finally, Figure (5.9) depicts u(x, y, t) for time
t=0.8.
Section 5.5: The Multidimensional Spatial Problem 175
Figure 5.6: The graph of u (x, y, t) for Figure 5.7: The graph of u (x, y, t) for
time t = 0. time t = 0.2.
2 2
Figure 5.8: The graph of u(x, y, t) for Figure 5.9: The graph of u(x, y, t) for
time t = 0.4. time t = 0.8.
EXERCISES 5.5
5.5.1. For the following PDE, separate the PDE into its respective ODES. Also,
completely separate the BCs,
a2 u a2 u au a2 u
C7t2
2
= C ax2 b
t +aa
y
2,
subject to BCs
176 Chapter 5: Separation of Variables: The Homogeneous Problem
5.5.2. For the following PDE, separate the PDE into its respective ODES.
5.5.3. Consider a thin rectangular plate of length L = 27r meters and width W = 47r
meters with perfect lateral insulation. Find the temperature distribution in
the plate given the following:
(1) The rectangular plate is made of silver (he thermal diffusivity of silver
may be found in Appendix E) and is subject to the BCs
u(0, y, t) = 0
au(L, y, t) _ 0
ax
with the IC
cos(xy) - 1; 0<x<L
2
0<y<W
- 2
W
y cos(x - L);
2 <x<L
L
- - 0y
<- <
u(x, y, 0) _
xcos(y - W);
-0x<2 -- 2
y<W
2
W
cos((x - L)(y - W));
2 <x<L
L
- - 2<--<W.
y
(2) The rectangular plate is made of granite (the thermal diffusivity of gran-
ite may be found in Appendix E) and is subject to the BCs
u(L, y, t) = 0
u(x, 0, t) = 0
au(x, W, t)
,
ay
Section 5.5: The Multidimensional Spatial Problem 177
with the IC
W
0<x< L3 0<
_y < 4
with IC
u(x,y,0)=cos(3x) 0<x<L 0<y<W.
5.5.4. Consider a thin rectangular plate of length L = m and width W = 7r m,
2
which offers no resistance to bending. Find the time-dependent solution given
the following conditions:
(1) The plat is subject to the BCs
u(0, y, t) = 0
u(L,y,t) = 0
u(x, 0, t) = 0
u(x, W, t) = 0,
and
W
(L-x)(y-W); L2 <x<L
- - 2< --
y<W.
r au(0, y, t)
ax
au(L, y, t) _ 0
ax
au(x, 0, t) _ 0
ay
u (x, W, t) = 0,
and
au(x, y, 0) _ 0
at
Section 5.5: The Multidimensional Spatial Problem 179
y sin x; 0<x<L
- - 0<-y <W2
u(x, y, 0)
W
(y-W)sinx; 0<x<L
- - 2 <-y-
<W
and
5.5.8. Consider a perfect laterally insulated thin sheet of glass Pyrex with length
37r ft and width 27r ft. Suppose the sides x = 0 and x = 37r ft are perfectly
insulated and the other two sides are held at zero degrees. Find the solu-
tion if the initial temperature distribution in the sheet is given as f (x, y) _
(cos x) (sine y). Note: The thermal diffusivity of glass Pyrex may be found in
Appendix E.
(5.98)
VxE=0. (5.99)
Section 5.6: Laplace's equation 181
From Equation (5.99) we know that E is the gradient of a scalar function, known
as the scalar potential, I. Thus,
E=-V. (5.100)
V2u = 0. (5.103)
(Note: The maximum principle is easily extended to the space Rn). Proof of
Theorem 35 may be found in Appendix B. However, a common sense approach
gives us the idea behind the maximum principle.
If we consider u(x, y) in Laplace's equation as the steady-state temperature
distribution in a plate, then u(x, y) can't be greater at one point in the plate than
all the other points of the plate because Fourier's heat conduction axioms say that
heat diffuses from high to low. Thus, if one point in the plate is hotter than the
rest, then the heat must flow away from that point to the surrounding points, thus,
reducing the temperature at the hot point. However, this would mean that the
temperature would change with time, a contradiction to the steady-state nature of
Laplace's equation. The same reasoning applies to a minimum point.
Having stated the maximum principle, we can state and show uniqueness of
Laplace's Equation in I12 for the Dirichlet problem. The Dirichlet problem is
2
a2 u a2 u
u =axe -}- 2= 0 in SZ,
ay
subject to
u(x, y) = f (x, y) on aSZ,
where 1 is as given in the previous theorem. The strategy of the proof is to assume
there exist two different solutions. Then, show that the solutions must be equal to
each other.
Proof Suppose there exists two solutions to the Dirichlet problem stated earlier,
u1(x, y) and u2 (x, y). Then let v(x, y) = u1(x, y) - u2 (x, y). Thus, v(x, y) is also
a harmonic function in 1 and zero on 9, because u1 (x, y) = f (x, y) = u2 (x, y)
on Dft Also, v(x, y) is continuous on 1 n aSZ since both u1(x, y) and u2(x, y)
are continuous and the sum of continuous functions is continuous. Hence, by the
maximum principle, v(x, y) must attain its maximum and minimum values on 1K.
Thus, v(x, y) = 0. Therefore, v(x, y) = u1(x, y) - u2 (x, y) = 0, which implies
u1(x, y) = u2 (x, y).
Now that we have shown that the solution to Laplace's equation is unique, we
solve Laplace's equation in Cartesian coordinate system.
_ y) a2u(x, y)
V2u(x,
(x, y) _ u(x, y) - a2u(x,
axe
+ a2
y
__
0 (5.105)
Section 5.6: Laplace's equation 183
u (x, H) =f4(x)
u(x, 0) =f3(x)
(5.106)
H)= = f4(x ).
u(x, H)
The only solution technique you currently have to solve a PDE is based on sep-
aration of variables, which requires a linear homogeneous PDE and BCs. The BCs,
Equation (5.106), are linear, but not homogeneous. Thus, separation of variables
does not appear to help us. However, suppose we have four separate solutions to
Equation (5.105). Then, by the principle of superposition, the sum of the four
solutions would also be a solution. Thus, consider
We really need to know two things about the four solutions. First, what do the
four solutions look like? Second, we know the solutions satisfy Equation (5.105),
but how do they satisfy the BCs, Equation (5.106)? The answer to both of these
questions is tied together. We let each of the solutions, u( x, y), i = 1,... , 4,
have one nonhomogeneous boundary condition from Equation (5.106) and three
homogeneous boundary conditions. Therefore, we have the following solutions,
184 Chapter 5: Separation of Variables: The Homogeneous Problem
u = f(x) uI = 0 u2 = 0
u =f(y) u,=0
+
u = f(y)
u = f(y)
uI = 0 u2 =f2( )
u=J(x) u, = 0 u,=0
ua =.fa(x)
U, = 0
+ +
tt, = 0 u0 ua = 0
U, =f(x) U, = 0
and
u4 = f4(x).
the nonhomogeneous condition we require for u (x, y). Similarly, we can show
u(x, 0) = f3(x), u(L, y) = 12(y), and u(0, y) = fl(y). Therefore, the original
problem may be broken down into four separate problems, each with three homo-
geneous BCs and one nonhomogeneous BC. Can we use the separation of variables
technique on the four separate problems? If we consider each of the four separate
problems as a boundary value problem, then we cannot. However, if we consider
the variable with the nonhomogeneous condition as a "time-like" variable, then we
can use separation of variables.
Since the method of solution is nearly the same for any of the four, the solution
for u3 (x, y) is given here. u2 (x, y), and u4 (x, Y). Figure (5.12) depicts the rectan-
gular plate u3 (x, y) It demonstrates how a simple shift of a coordinate aids us in
.
the solution. The coordinate shift is possible because the solution of the resulting
ODE is invariant under a translation. That is, the translation does not change the
solution. The remaining three are left as an exercise.
u3(x, H) = 0
u3(0, y) = 0 u3(L, y) = 0
u3(x, 0) =f3(x)
Figure 5.12: Heat conduction in a rectangular plate, u3 (x, y), of length L and
height H.
subject to
u3(0, y) = 0
u3(L, y) = 0
(5.109)
u3 (x, 0) = f3 (x)
u3(x, H) =0.
We assume at x = 0 and x = L that we have boundary conditions and at y = 0
and y = H that we have "time-like" conditions. Letting u3 (x, y) = 3(y)G3(x) and
separating in the usual manner we arrive at the two ODEs
subject to
fX3(0)=0
Ps(L) _
and
subject to
G(0)=f(x)
G3(H)=0.
The spatial ODE, Equation (5.110), subject to its BCs, has the solution
(fl7 2
L)
n= 1,2,3,... (5.112)
n7rx
SP3n (x) =sin L
subject to
Gsn(0) = f3(x)
G3(H)
Gin (H) _ 0,
Section 5.6: Laplace's equation 187
As practice, you should show that Equation (5.115) solves Equation (5.113).
We may now apply the condition G3n (H) = 0, which yields
Gsn (H) _ 0 _ c.
Therefore, we have
n7r(y
Gsn (y) =din sinh - H) . (5.116)
L
Combining the solutions for 3n (y) and G371 (y), and using the principle of su-
perposition, we arrive at
00
nir(y H) nirx
u3(x, y) = d 3n sinh sin
n=1
L L
Using the "time-like" BCs, u3 (x, 0) = 13(x), we may determine the last unknown,
d3n , by using the orthogonality of the Fourier sine series. Thus, we have
00
nom( H) fx n3(
u x 0) = f3( x) = d3 sinh
L
sin , (5.117)
n=1
which implies
-2 L
f 3 (x) sin L dx. (5.118)
L sinh L o
The solutions for u1 (x, y), u2(x, y), and u4(x, y) are determined similarly, and
are left as an exercise. Note: For u1 (x, y), a coordinate shift must be determined.
You should prove to yourself that the translated solution of the ODE does not
change the differential equation. For u2 (x, y) and u4 (x, y), no shift is required.
Once all four solutions are found we use the principle of superposition to combine
the four solutions, which yields the solution for u (x, y). Also, please remember
that Laplace's equation is the primary representative of the elliptic class of linear
second-order PDEs.
188 Chapter 5: Separation of Variables: The Homogeneous Problem
In Chapter 10, we will again meet Laplace's equation. Only then we will work
in polar, cylindrical, and spherical coordinate systems.
EXERCISES 5.6
subject to
u(x,H) = f4(x).
Remember that u3 x, y) was solved in the text.
5.6.3. Find the complete solution of
52u(x, y) _ 0,
V2u(x,
(x, y) - D2u(x,
axe
y)
+ ay2
subject to
8u(0, y)
- .fi (y)
8x
au(x, H) _
f4(x).
ay
5.6.4. Consider Laplace's equation inside a rectangle. Suppose .fi(y) = f3(X)
f4(x) = 0 and 12(y) = y + 1. Determine the solution.
5.6.5. Given Laplace's equation in Cartesian coordinates
y) _ _
D2u(x, y) D2u(x, y)
2
v u(x, - oyy
(,y),
1 u0 =0 au(2'y) = ux 0 =0 andux 3 =x.
(2) u0 = 2 au(, y) =
u(x, 0)
- 0, and u(x, 27r) = sin x.
ax ' a
aua(0x, y) au(x, 0)
=0 ' u ('y) =cos y'
a
0, and u(x,1) = 0.
(3) 2
5.6.13. PROJECT:3 This project models the fluid in the cochlea surrounding the
basilar membrane, which is part of the inner human ear. It is assumed that the
fluid in the cochlea is incompressible and inviscid (not thick). If the equations
are nondimensionalize we arrive at
D2u(x, y) 52u(x, y) =0
axe + D y2
subject to
8u(0, y)
=1
ax
u(1, y) = 0
2u(x, 0)
' I
u(x, H) = 0.
where I is called the impedance in the damped harmonic oscillator.
(1) Find the general solution.
(2) Show that after truncating the series at N terms, multiplying by cos(mirx),
and integrating from 0 to 1, we obtain the system of linear equations
N
An nm fm,
n=0
where
f
1
cos(nirx) cos(mirx)
nm = 2 cosh(nirL) dx - 1 nor sink( nii H )nm
S
I 2
fm=LSmp-
f 1 x(2 - x) cos(mirx)
I
dx.
3Adapted from James Keener and James Sneyd, Mathematical Physiology, ©1998 by Springer-
Verlag, New York, pp. 707-711. Reprinted by permission.
Chapter 6
191
192 Chapter 6: The Calculus of Fourier series
First, the sets A and B in the definition are the set of real numbers, R, for
Fourier series. Note: Both sets A and B can be the same. Second, the subset
D of A in the definition refers to the interval [-L, L], a subset of Ilk for Fourier
series. Finally, the very nature of Fourier series is to assign to each element in the
interval [-L, L] a unique element in the range. Thus, a Fourier series is a function.
Although knowing a Fourier series is a function is important, the really important
concept is knowing if the Fourier series is a continuous function.
From Chapter 4, we know the Fourier series representation of a function on the
interval [-L, L] converges to the function at all points the function is continuous
and to the average value at any point of discontinuity. Consider the following two
examples.
f(x)
r
7r-x, 2 <x <?r.
The function 1(x) is a piecewise smooth continuous function. Its Fourier series
representation is
1(x)
- 16 + n=1 (urn \(-1)n - cos 2) cos nx
00,
-- ((6 sin
nor
- r -1 ))
n
1 sin nx.
n= 1
2 ---)
2n r
The solution is graphed in Figure (6.1) on the interval [- 37r, 3ir]. It shows the
Fourier series representation of 1(x) converges to - at x = -3ir and x = 37r.
4
Since Fourier series are periodic with period 2L (where L is the length), we know
the Fourier series representation of 1(x) at x = -?r or it also converges to .
4
Also, the Fourier series representation of 1(x) at x = -?r or it converges to ,
4
but the function equals
-it
and 0 respectively. This means = f
-it f
2 4 2 '
or the average value of the endpoints of the function. Thus, a jump discontinuity
exists in the Fourier series representation for 1(x). Therefore, the Fourier series
representation of 1(x) is not continuous and can not be differentiated term-by-
term.
Look closely at Figure (6.1). As the Fourier series Representation of 1(x) approaches
it, from the left or from the right, the graph seems to oscillate. Then the Fourier
Section 6.2: Fourier series as a Function 193
1.5
Figure 6.1: The graph of the Fourier series representation off (x) .
2, 2
g(x)= -x,---x<
-2,2<x<7r.
37r 7r
The function g(x) is a piecewise smooth continuous function. Its Fourier series
representation is
37r 2 n7r
g (x) = + ((_i) n - cos cos nx
16 7rn 2 2
n=1
+
00
-4 sin
n7r
sin nx.
7rn 2 2
n=1
The solution is graphed in Figure (6.2) on the interval [-37r, 37r], which shows the
Fourier series representation of g(x) converges to at x = -37r and x = 37r. As
2
in the last example, since the Fourier series is periodic with period 2L, we know
that the Fourier series converges to at x = -it or it. Thus, the Fourier series
2
representation for g(x) is continuous at the endpoints. And since there are no other
discontinuities in the function, the Fourier series representation of g(x) is continuous
on the interval [-7r, it].
Knowing when a particular Fourier series is a continuous function, without being
forced to graph the Fourier series, is very important. At the end of Chapter 4, three
theorems were given to clarify when a particular Fourier series is continuous. These
definitions are restated below:
Theorem 37. If f (x) is a piecewise smooth continuous function on the interval
[-L, L] and f (-L) = 1(L), then the Fourier series of the function 1(x) is contin-
uous on [-L,L].
Theorem 38. If f (x) is a piecewise smooth continuous function on the interval
[0, L], then the Fourier cosine series representation of the function f (x) is contin-
uous on [0, L].
Theorem 39. If f (x) is a piecewise smooth continuous function on the interval
[0, L] and f(0) = f (L) = 0, then the Fourier sine series representation of the
function f (x) is continuous on [0, L].
Notice in all three theorems the function f (x) is continuous and piecewise smooth
on the interval. Both characteristics are required. A function may be continuous
on the interval [-L, L] but not be piecewise smooth. A good example is the tan x
on the interval [- , ]. At - and the tan x goes to - oc and oc, respectively.
2 2 2 2
Likewise, if a function is piecewise smooth and not continuous, the Fourier series
representation will have jump discontinuities. Thus, the Fourier series as a function
will not be continuous. Also note, for two of the theorems, specific values for
the function at the end points are given. The endpoint values are very important
because they insure the Fourier series representation of the function in Theorems
(37 and 39) converges to the value at the endpoints of the function.
Why is continuity of a Fourier series so important? The answer is quite straight-
forward. Since you cannot differentiate a function on an interval if the function has
Section 6.3: Differentiation of Fourier series 195
and
u(x,O) = 1(x).
The solution is
n ix
00
1 L n7rx
an = L J 1(x) cos dx,
-L L
and
1 L
bn L (x) sin L dx.
- Lf
Note: The Fourier coefficients do not depend on the variable x.
Following the same procedure, which exists for second-order linear ODEs, and
noting the Fourier series solution in Equation (6.6) is continuous in the spatial
variable by Definition (37) and is assumed to be continuous for the time variable,
we must find
00
au = a [ane_(t)2t cos nrx + bne - (L) t sin nirx
2
a0 +
at at L n=1
L L
a a nirx
+bne - (L) sin
2t 2t n7rx
[ao]+ [ane (L) cos
at at Ln=1
L L
au a
a0 +
°°
[ane_()2t cos nrx + bne - ( L) 2 t sin nrx
ax ax n=1
L L
a a °° 2t nrx 2t
[ao] + [ane_ (L) cos +bne _ (L) sin ,
ax ax n-1 L L
and
a2 u a a °O nrx +bne n7rxi
-_L
= a0 + [ane_()2tcos - ( L ) 2 t sin
ax 2 ax ax n=1
L L
a
[aol
ax 125;;
aJa °° nrx
ax ax [anetcos
_ (L)
L
- bne _ (L)
2 2t
sin
-i-i] }.
(6.9)
n=1
What we need to perform on the left side of Equations (6.7, 6.8, and 6.9) is quite
straightforward. However, the right sides of Equations (6.7, 6.8, and 6.9) present
Section 6.3: Differentiation of Fourier series 197
interesting questions. How do you differentiate an infinite series, and what do you
get for a solution? Also, an interesting follow-on question is if we can differentiate a
Fourier series, which is an infinite series, can we differentiate all infinite series by the
same method? This is not a course in general infinite series, but the answer to the
last question depends on whether the infinite series is continuous and convergent, as
well as if the derivative of the infinite series is uniformly convergent. For Definitions
(37, 38, and 39), the Fourier series representation of the function is continuous,
converging uniformly to the function. This simple fact makes working with Fourier
series much simpler. It means differentiating a Fourier series is possible.
Since differentiation of a Fourier series is possible under certain conditions, what
is the end result of differentiation of a Fourier series, and how do we perform the
differentiation? We perform the differentiation term-by-term, which must be shown.
The end result is hopefully another Fourier series, which is also the Fourier series
representation of the derivative of the original function.
To prove these statements, we consider a general Fourier series representation
of a function, u(x, t), where all the Fourier coefficients depend on the variable t:
00
Also, we assume that the function u(x, t) is continuous for both time and space
variables and u(-L, t) = u (L, t). These conditions insure the Fourier series rep-
resentation of u(x, t) is a continuous function on the interval [-L, L]. Next, we
assume that u(x, t) is continuously differentiable with respect to both variables. Fi-
nally, we hope that the derivative with respect to either the spatial variable, x, or
the time variable, t, of a Fourier series is another Fourier series, which converges to
the derivative of the original function, u (x, t). Then, we assume that the derivative
with respect to either variable of u(x, t) is a piecewise smooth function. We find the
derivative with respect to time first, then we state the resulting theorem. Next we
find the derivative with respect to the spatial variable, and we state the resulting
theorem.
We have
where u(x, t) is continuously differentiable with respect to t, and the Fourier series
representation of u(x, t) is continuous. Also, we know that the derivative of u(x, t),
au(x, t) au(x, t)
is p iecewise smooth and thus may
Y be written as a Fourier series
at at
which is
au(x, t) °O n7rx n7rx
= Ao (t) + [An(t) cos + Bn (t) sin (6.12)
at
n=1
L L
Note: The coefficients of Equation (6.12) are different from the coefficients in Equa-
tion (6.11).
198 Chapter 6: The Calculus of Fourier series
L
an(t) = 1L-u(x,t) cos n7rx
L dx, (6.14)
L
and
L
1 n7rx
bn (t) = L-u(x,t) sin L dx. (6.15)
L
The equations for the Fourier coefficients corresponding to Equation (6.12) are
1 L 8u(x, t)
Aoff)
t = ZL L at dx (6.16)
_ 1 L au (x, t) n7rx
An cos dx, (6.17)
(t) L -L at L
and
1 L au(x, t) n7rx
Bn (t) _ L sin dx. (6.18)
-L at L
The form of the integrals in Equations (6.16, 6.17, and 6.18) should remind you of
Leibniz's theorem in Chapter 2, which we restate here.
af
Theorem 40. Suppose f (x, t) and the partial derivative (x' t) are continuous in
at
some region of the xt-plane where a < x < b, then
td J f r6(x, t) dx b of (at x, t)
dx.
Ia
au(x t)
Thus, if we further assume that at' is continuous in some region of the
xt-plane where -L < x < L, then Equations (6.16, 6.17, and 6.18) become
L
A0(t)
dt
[fu(xt)
2L
dx (6.19)
L
Section 6.3: Differentiation of Fourier series 199
L
,nLx dx (6.20)
`4n(t) dt L ,
and
rL
Bn(t) _ t L u(x, t) sin nLx dx (6.21)
J L
Using Equations (6.13, 6.14, 6.15), and substituting equivalent terms, we find that
Equations (6.19, 6.20, and 6.21), respectively, become
`4°(t)
dt
[u(x,t)
,f
2L
L
dx = dt
[ao(t)] = ao(t),
L
f u(x, nLx dx
L
A(t) _ [Lt)cos - at [an(t)] - an(t),
and
L
Bn (t)
dt L fL u(x, t) sin nLx dx = dt
[b(t)] = b( t).
Thus, the Fourier coefficients of Equation (6.12) are derived from the Fourier coef-
ficients of Equation (6.11). Therefore, term-by-term differentiation with respect to
a parameter t is valid, and we have proved the following theorem:
au(x t)
Theorem 41. If u(x, t) and are continuous functions, then the Fourier
series representation of u(x, t) on the interval [-L, L],
00
n=1
may be differentiated term-by-term with respect to the parameter t, and the result is
the Fourier series representation of the derivative of u(x, t) with respect to t,
00
au(x, t) n7rx , n7rx l
at
= a( t) + [a(t) cos L
+ bn (t) sin
L
n=1
u(x, t) = ao(t) -}- [an(t) cos nLx + b(t) sin nLx] , (6.22)
n=1
200 Chapter 6: The Calculus of Fourier series
Again, it suffices to show that the Fourier coefficients of Equation (6.23) may be
derived from the Fourier coefficients of Equation (6.22). The equations for the
Fourier coefficients corresponding to Equation (6.22) are
L
ao(t) = ZL IL u(x, t) dx, (6.24)
fL
an (t) = LL u(x, t) cos nLx dx, (6.25)
L
and
L
1 nrx
bn(t) = u(x, t) sin L dx. (6.26)
L -L
The equations for the Fourier coefficients corresponding to Equation (6.23) are
L
1 Du(x,t)
(6.27)
2L f L 8x '
_ 1 L au(x, t) nrx
an (t) cos dx, (6.28)
L -L ax L
and
1 au(x, t) nrx
sin dx. (6.29)
L ILL ax L
L
a ax't)
dx = 2Lu(x, t)
-L
- 1
[u(L,t) - 'u(-L,t)] = 0,
L
1
L
[u(Lt)cosmi - u(-L, t) cos flit + 1J u(x, t) sin mix
flit
L
dx
-L
1 nit L mix
L
[(u(Lt) - u(-L, t)) cos flit +
L- L
u(x, t) sin
L
dx
1 n7r L mix
Lu(x, t) sin dx (6.30)
L L - L
an (t) =
flit [fLu(x,t)sin
1
dx . (6.31)
L L
Using Equation (6.26) and substituting equivalent terms, Equation (6.31) becomes
L
_ flit 1 mix flit
an (t) L
L -L
u(x, t) sin L dx = L bn (t).
Thus, an (t), in Equation (6.23), depends on the coefficient bn (t), in Equation (6.22).
d flit- = flit flit-
Again, this is not unexpected since sin cos
dx L L L
Similarly, we find the Fourier coefficient /n (t) depends on the coefficient an (t) .
Thus, we have proved term-by-term differentiation of a Fourier series representation
of a continuous function with respect to the variable x. We state our results in the
following theorem:
Theorem 42. If u(x, t) is a continuous function on the interval [-L, L] with u(-L, t)
au(axx' t)
= u(L, t), and is a piecewise smooth function, then the Fourier series
representation of u(x, t) is continuous and it can be differentiated term-by-term.
202 Chapter 6: The Calculus of Fourier series
There are two more theorems stated subsequently, with proofs left as exercises.
These theorems develop term-by-term differentiation of Fourier Cosine and Fourier
sine series representation of a function u(x, t). As you may have guessed, knowing
these theorems is very important in the chapters that follow.
au(x t) .
Theorem 43. If u(x, t) is a continuous function on the interval [0, L], and Zs
ax'
a piecewise smooth function, then the Fourier cosine series representation of u(x, t)
is continuous and it can be differentiated term-by-term.
Theorem 44. If u(x, t) is a continuous function on the interval [0, L] with u(0, t) _
au(axx' t)
u(L, t) = 0, and is a piecewise smooth function, then the Fourier sine series
representation of u(x, t) is continuous and it can be differentiated term-by-term.
We have covered all cases where Fourier series may be differentiated term-by-
term. If the function u(x, t) in Theorems (42, 43, and 44) is a function of one
variable, that is, f (x), the theorems still hold. Next, we turn our attention to
term-by-term integration of a Fourier series.
EXERCISES 6.3
6.3.1. Show that the Fourier coefficient Bn (t) in Equation (6.29) depends on the
coefficient an (t) in Equation (6.25).
6.3.2. Prove Theorem 43.
6.3.3. Prove Theorem 44.
6.3.4. Given
a2 u 2 a2 u
5;=C
axe
subject to
u(0, t) = 0 and u(L, t) = 0
and
n-1
Prove, using term-by-term differentiation, that the solution satisfies the PDE
and BCs.
Section 6.3: Differentiation of Fourier series 203
6.3.5. Given
au a2 u
at axe
subject to
au(0, t) au(L't)
=0 and -0
ax ax
and
u(x,0) = f(x),
Prove, using term-by-term differentiation, that the solution satisfies the PDE
and BCs. State reasons for all differentiations.
6.3.6. Given
n=1
where 0 < x < L and 1(x) is a continuous function, ti means approximately,
and f' (x) is piecewise smooth. Suppose 1(0) = a 0 and f (L) = 0, and
show that
n7rx
1
[-a]+ Ln((_l)_]c05
n71 2 n 1
n=1
6.3.8. Consider
00
nix
cosh x = ao + an cos L, 0 < x < L.
n=1
cosh x =
00
- n 2
an cos
nix
, 0 < x < L.
L L
n=1
°° nor 2 n?fx
a0 + an cos = - (-i-) an cos
L n=1
L
n=1
a0=0andan=0.
This conclusion is clearly false. Determine the error in the logic.
6.3.9. Consider
00
nix
sinh x = a0 -F- an cos , 0 < x < L.
L
n=1
a0=0andan=0.
This conclusion is clearly false. Determine the error in the logic.
6.3.10. Given
au a2u
ax2
Section 6.3: Differentiation of Fourier series 205
subject to
au (L, t)
()
u 0 t= 0 and
ax
=0
and
u(x,0) = f(x)
(2n-1)71x
u(x, t) = bne-an of sin ,
n=1
2L
2
[(2n_1)71
n= 2L ,n=1,2,3,...
Prove, using term-by-term differentiation, that the solution satisfies the PDE
and BCs. Briefly discuss any assumptions you make.
a2u
=16 a2u
axe
ate
4n71t mix
an cos L + bn sin 4n71t
L
sin
L
n=1
2
An = n= 1,2,3,...
L
Prove, using term-by-term differentiation, that the solution satisfies the PDE
and BCs. Briefly discuss any assumptions you make.
6.3.12. Given
au a2 u
at ax2 '
206 Chapter 6: The Calculus of Fourier series
subject to
and
u (,),)
-L t = u L t and au(-L, t) _
a
au(L, t)
ax
u(x,0) = f(x),
u(x, t) = ao +
00
n=1
e- an cos
n?rx
L
+ bn sin
n7rx
L
_),
nor 2
An=
[-IL-I
,n=1,2,3,...
Prove, using term-by-term differentiation, that the solution satisfies the PDE
and BCs. Briefly discuss any assumptions you make.
mrx
Since sin L is an odd function,
L
mrx
ao sin L dx = 0.
-L
Therefore, we have remaining
L
mrx L°° n7rx n7rx mrx
I L
1(x) sin
L
dx =
-L n=1
[an cos
L
+ bn sin
L
sin
L
dx. (6.32)
Section 6.4: Integration of Fourier series 207
In Equation (6.32), once we know the function 1(x), the left integral is feasible.
However, to perform the integration of the infinite series in Chapter 5, you were
asked to believe the integral of the infinite sum is equal to the infinite sum of the
integrals, and the result was a formula for the term b. This would tend to mean
term-by-term integration of a Fourier series representation of any piecewise smooth
functions on the interval [-L, L] is possible. We confirm this suspicion with a
theorem.
Theorem 45. 111(x) is a piecewise smooth function on the interval [-L, L], then
the Fourier series representation of f (x) can always be integrated term-by-term.
The result is an infinite series, not necessarily a Fourier series, which converges to
the integral of the function f (x) for -L < x < L. Note: It does not matter if the
function f (x) has jump discontinuities.
The proof of Theorem (45) can be found in the book titled Infinite Series, by
Earl D. Rainville, published in 1967.
Two examples are given showing term-by-term integration of Fourier series.
These examples demonstrate the unique features of Theorem (45). Remember, you
are guaranteed an infinite series that converges to the integral of the function the
Fourier series represents.
EXAMPLE 6.3. Consider the function f (x) = x on the interval [-2, 2]. Since
f (x) is an odd function, the Fourier series representation of f (x) = x is the same as
the Fourier sine series representation of f (x) = x on the interval [0, 2]. Therefore,
°O -4(-1)n n7rx
x= sin (6.33)
nor 2
n=1
x2
Integrating Equation (6.33) will yield an infinite series for the function 2 . That
is,
_4(_1)72 fl717
T dT = / sin dT =
0 nor 2
0 n=i
00
-4(-1)n
nor 0
x
sin-dr.
n71T
2
(6.34)
n=1
Notice the integrals in Equation (6.34) are from 0 to x. We want the integrals this
way because we want a function after integration. Performing the integration yields
x
x2
_
°° -4(-1)n -2 coS n71T =
2 n=1
n71 [n71 2 Jo
°° 4(-1)72 2 mix
cos -1 (6.35)
n-1 nor nor \ 2
208 Chapter 6: The Calculus of Fourier series
x2
2
_ 00.
L (n)
8(-1)n
2 (cos 2
nix
- 1)
n=1
00
8(-1)n
(mu)2'
nix
2
2
- 8(-1)n
(mu)
a
00
8(-1)n n7rx 8(-1) n
_ 2
COS
2
n=1 (n7r) n=1 (n71)
or
x2 =
-
16(1)n
cos
n7rx
2
00
.- (nit)2 n-1
(1Z7C12
00
16(-1)n , 16(-1)n mix
cos 2 (6.36)
n= n=1 (n71) 2
We know the infinite series in Equation (6.36) converges to x2. But, as expected,
the infinite series in Equation (6.36) is not a Fourier series. Sometimes, we really
need a Fourier series solution, and with a little algebraic manipulation we can
transform the infinite series, like that in Equation (6.36), into a Fourier series.
EXAMPLE 6.4. Starting with the infinite series in Equation (6.36),
00 o0
2 16(-1)n 16(-1)n n7rx
X=- 2 -F- 2 cos
2
(6.37)
n=1 (mu) n=1 (n71)
We notice the second term on the right has a cosine function. The cosine function
indicates the possibility of transforming it into a Fourier cosine series. Thus, we set
up the standard Fourier cosine series for g(x) = x2,
00
n7rx
g(x) = x2 = ao + an cos
2
n=1
16( -1)n
Comparing similar terms, we find the term - 2
, in Equation (6.37), must
n=1 (nit)
16(-21) n
be the ao term in Equation (6.38), and the term , in Equation (6.37), must
(mu)
be the an term in Equation (6.38). From Equation (6.38), we determine an and ao.
We have
mix
2
an = L J g(x) cos
L
L
dx =
2
x2 cosn7rx
L dx
0 0
16(-1)n
(6.39)
Section 6.4: Integration of Fourier series 209
and
f2
a0 = 2J x2dx=3. (6.40)
0
2 4 °O 16(-1) n nirx
x=3+ (n7r)
2 (6.41)
n=1
The right side of Equation (6.41) is almost what we have in Equation (6.37). The
16(-1) n
an term in Equation (6.41) is identical to the term 2
in Equation (6.37).
(n71)
To complete the process, we must show
4 16(-1) n
3
n=1
2 3
n=1 n=1
which implies
00
4 16(-1) n
'
(6.42)
3
n=1
n2
16(-1)n n7rx
-3
x2=+ 2 cos
2
(6.43)
n=1 (n71)
Suppose we would like to integrate Equation (6.43) and express the solution as
the Fourier sine series representation of x3. This is done in the next example.
EXAMPLE 6.5. Consider
4 °° 16(-1) n n7rx
x2 = -+ 2 cos
2
(6.44)
3 (n71)
n=1
on the interval [0,2]. Integrate Equation (6.44) and express the solution as a Fourier
sine series representation of x3.
210 Chapter 6: The Calculus of Fourier series
x o0 x
4 16(-1)n n71Y
dr + 2 cos dT.
(n71) o 2
3 o n=1
This becomes
x3 4 °° 16(-1) n 2 n7rx
_-x+ 2 sin
3 3 n=1 (nr) [mu 2
-4 °O 32(-1)n mix
-x+ 3
n=1 (n7r)
3
sin
2
(6.45)
However, Equation (6.45) does not look like the Fourier sine series of 1(x) = x3.
Therefore, first we multiply Equation (6.45) by 3, which yields
°O 96(-1)n , n7rx
x3 =4x+ sin (6.46)
n=1 (n71)
Second, the right side of Equation (6.46) must be completely expressed as a Fourier
sine series on the interval [0, 2]. Therefore, we must express the 4x term in Equation
(6.46) as a Fourier sine series on the interval [0, 2]. We have
°O
n7rx
4x = bn sin ,
2
n=1
where
2
mix -16(-1 n
b= 4x sin 2 dx =
n7r
0
Thus,
4x=
°° (-)
161
n
sin
n7rx
(6.47)
n=1
mr 2
Section 6.4: Integration of Fourier series 211
where
b=
n L
2 L
sinn7rx
L dx = 2 n
(1 - (-1)).
0 n7r
yields
X
x=(1-(-1))
°° n) I- L cos nlrs
2
n7r
n_ 1
[n7r L j0
-2 n x - 1]
(1 - (-1)) [cos_Z_
n_1 (n7f)2
00 00
_
2L
2 (1 - (-1)n) - (n7r)2 (1 - nx(6.49)
(-1))cos--.
n=1 n n=1
212 Chapter 6: The Calculus of Fourier series
Equation (6.49) has the form of the Fourier cosine series for the function f (x) = x
2L_(1
where ao = - (-1)n). Finding ao in the usual way, we have
n=1 (nor)
ao=
1
L
L
xdx= L2 .
0
00
2L
Therefore the infinite series 2 (1 - ( -1)n) converges to L written as
n=1 (nit)
2
00
L
(nit)2 (1 - (-1)n)
2
n=1
for any positive number L.
Another use of term-by-term integration of Fourier series is determining the
Fourier coefficients, which is left as an exercise. This concludes our discussion
about integrating Fourier series. We move on to the Gibbs phenomenon.
EXERCISES 6.4
Hint: Find the Fourier sine series for the function f (x) = x, then integrate the
Fourier sine series to get the Fourier cosine series of f (x) = x2 where L = it.
Section 6.5: Fourier series and the Gibbs phenomenon 213
Hint: Find the Fourier sine series for the function 1(x) = x2, then integrate
the Fourier sine series to get the Fourier cosine series of f (x) = x3 where
L=7C.
representation of 1(x) is
x2 _ L+ 2 °O 4L2(1)72
cos (6.50)
3 n_1 n7r 2 \ L
214 Chapter 6: The Calculus of Fourier series
Now, consider the following sequence of partial sums of the Fourier series given in
Equation (6.50) with their corresponding Figures(6.4-6.9):
Sl
()x _ LZ+
3
1 4L2(-1)n
(n)2 cos
rn7rxl
\LI
L2 4L2
3(n)2 cos \ L /
n=1
S2(x) _ L2 + r. 4L2(-1)n
cos
3 L (n7r)2 \L
= 3-
2
4L2 rx 4L2
cos
( 27rx
(
ncos ( L )+ ( )2
Figure 6.4: The graph of f (x) = x2 Figure 6.5: The graph of f (x) = x2
and the partial sum Sl (x). and the partial sum S2 (x).
L2
-L
Figure 6.6: The graph of f (x) = x2 Figure 6.7: The graph of 1(x) = x2
and the partial sum S5(x). and the partial sum Slo(x).
S L2 + 5 4L2(-1)n nirx
3
n=1
(n cosL().
LZ 10
4L2(-1)n nirxl
Sio(x) = + cos
n=1
(n
Section 6.5: Fourier series and the Gibbs phenomenon 215
2f_ \n
2
Sis (x) = -a-- + :i:
15
4 n x,
cos (----)
n=1
Sioo(x) =
LZ ioo
+ :i:
4L2(-1)n
cos
nixn=1
(n7r)2 (-z--)
- L2
-L L -L L
Figure 6.8: The graph of 1(x) = x2 Figure 6.9: The graph of f (x) = x2
and the partial sum S15(x). and the partial sum Sloo(x)
It is easy to see that the sequence of partial sums of the Fourier series represen-
tation of the function x2 converge uniformly to the function x2. In fact, by partial
sum Sioo (x) the graph of the partial sum and the graph of the function are almost
indistinguishable, which means convergence is relatively fast. Note, each partial
sum, Sn, n = 1,2,3,..., is a trigonometric polynomial. This fact is very important
and the following theorem brings it to light.
Theorem 46. (Weierstrass approximation theorem) If the function 1(x) is contin-
uous and has period 2L, then it can be uniformly approximated by trigonometric
polynomials.
on the interval 0 < x < L. Please note the period of this function is still 2L.
Remember, you must do the odd periodic extension shown in Figure (6.10) to
generate the Fourier sine series. The Fourier sine series is
-L L
00
10 [1 - (_i)12] nirx
f(x) = 5 =
n=1
n sin
4
Because of the term [1 - (-1)n], we recognize that for even natural numbers, n =
2n, S2n (x) = 0. Therefore, the following sequence of partial sums and graphs-
Figures (6.11-6.16)-are for odd n.
-L
-5
Figure 6.11: The graph of the odd Figure 6.12: The graph of the odd
periodic extension of f (x) = 5 and periodic extension of 1(x) = 5 and
the first partial sum Si(x). the first partial sum S3 (x) .
3
10 [1 - (_i)12] 20
S3 (x) _
n?r
sin
,
4
=
?r
sin
7rx
4
+ 20
37r
371x
-sin----.
4
n=1
Section 6.5: Fourier series and the Gibbs phenomenon 217
Figure 6.13: The graph of the odd Figure 6.14: The graph of the odd
periodic extension of 1(x) = 5 and periodic extension of f (x) = 5 and
the first partial sum S7 (x) . the first partial sum S15 (x) .
10 [1 - (-1)n] nix
S15(x) _ sin ----
4
n=1
5 5
-L -L
-5
Figure 6.15: The graph of the odd Figure 6.16: The graph of the odd
periodic extension of 1(x) = 5 and periodic extension of f (x) = 5 and
the first partial sum S99(x). the first partial sum S201 (x).
10 [1 - (_i)12]
sin
nix
S99(x) _
n=1
nit 4
2:i: 10 [1 _(_i)12]
S2o1(x) =
n sinnix
4
-
n=
218 Chapter 6: The Calculus of Fourier series
By the partial sum, S201 (x), we see that the Fourier sine series representation
of the function f (x) = 5 is converging to 5 at points away from the jump discon-
tinuities at x = 0, x = ±L. At the points x = 0, x = ±L, we notice an overshoot
of the Fourier series. This overshoot is known as the Gibbs2 phenomenon. As
n --f oc, we could expect the overshoots to disappear. However, in reality they
don't. It can be shown that the Fourier series converges to the average value at any
jump discontinuity. It should be noted that in higher mathematics, we can show
that the Fourier series actually has weak convergence to the endpoints of a jump
discontinuity. This weak convergence is from the left and from the right at a jump
discontinuity. However, the concept of weak convergence is beyond the scope of this
course.
Does the Gibbs phenomenon occur when we solve a PDE? Consider the following
example.
EXAMPLE 6.8. Suppose we are given the heat problem
on 02 u
O`lt O`lx2 '
subject to
and
u(x, 0) = x2 + x - 1.
We see that the IC has the value of -1 at x = 0 and 5 at x = 2. From our previous
work in Chapter 5, we know that the general solution is
00
bri =
2
L f L , nirx 2
f (x) sin L dx = 2
0
L
(x2 + x - 1) sin 2 dx
f 2
(x2 + x - 1) sin 2 dx =
2 (1 + 7 (-1)"d)
n71
2Josiah Gibbs (1839-1903), an American physicist, brought the overshooting phenomenon to the
'
attention of the scientific community in a paper. However, it was first noticed by the mathematician
Henry Wilbraham and was rediscovered by the British during World War II while developing radar.
Section 6.5: Fourier series and the Gibbs phenomenon 219
x
x
1 1.5
Figure 6.17: The graph of f (x) = Figure 6.18: The Fourier sine series
x2 + x - 1. representation of 1(x) = x2 + x - 1.
Figure (6.19) shows the graph of u(x, t) for time 0 < t < 0.5. Please note how the
0.4 t
u(x, t)
boundaries are fixed at 0. Also, the surface of the function has the curve x2 + x - 1
for 0 <x <2 subject to the decreasing exponent. The function 1(x) = x2 + x - 1
is shown as a bold line where it was produced by a Fourier series partial sum of
n = 500. Note how the IC does not quite follow the curve of u(x, t) at either end
at time t = 0.
EXERCISES 6.5
For each set of BCs and ICs, solve the initial value problem; using your favorite
mathematical software, graph the solution for 0 < t < 2. Clearly indicate the
initial temperature distribution on your graph and determine if the Fourier se-
ries expansion of the initial temperature distribution is accurately represented
on the graph of u (x, t).
u(o, t)= o
(1) Bcs: and IC: u(x, 0) = x-4.
t) = o
u(0, t) = 0
(2) BCs: and IC: u(x, 0) = x3 +2x- 1.
0
8x
8u(0, t) _ 0
8x
(s) Bcs: and IC: u(x, 0) = x2 -1.
/3 \ 0
I u 2,t)=
8u(0, t)
=0
8x
(4) Bcs: and IC: u(x, 0) = x2 + 2x + 4.
8u(ir, t)
8x
-0
6.5.2. Consider the PDE
a2u(x, t) a2u(x, t)
=1 axe
ate
For each set of BCs and ICs, solve the initial value problem. Using your
favorite mathematical software, graph the solution for 0 < t < 1. Clearly
indicate the initial displacement on your graph and determine if the Fourier
series expansion of the initial temperature distribution is accurately repre-
sented on the graph of u(x, t).
Section 6.5: Fourier series and the Gibbs phenomenon 221
au (o, t)
-x; 0<x<.5
=0 J u(x, o) _
(4) BCs:
8x
and IC: x-1; .5<x<1
8u (1, t)
J dx =0.
at
6.5.3. Let a metallic rod 45 cm long be heated to an initial temperature which
is modeled by (x sin 2x + 3)°C. Suppose at t = 0, the ends of the rod are
plunged into an ice bath of 0°C and thereafter maintain this temperature.
Also, suppose no heat is allowed to escape from the lateral surface of the rod.
Note: You must model and explain your choice of boundary conditions for
this experiment.
(1) Solve the initial value problem if the rod is made of gold. Note: The
thermal diffusivity of these materials may be found in Appendix E.
(2) Using your favorite mathematical software, graph the solution for each
material, then determine the approximate solution for the following three
cases:
(a) Use only the first term in the series for u(x, t) to find the approxi-
mate temperature at x = 15 cm when t = 30 secs.
(b) Use the first three terms in the series for u(x, t) to find the approx-
imate temperature at x = 15 cm when t = 30 secs. Compare this
answer with the previous answer.
(c) Use the first 25 terms in the series for u(x, t) to find the approximate
temperature at x = 15 cm when t = 30 secs. Compare this answer
with the previous two answers.
(d) Explain, in a short essay, the differences in the above three approx-
imations.
(3) Determine the amount of time required for the temperature of the rod
to reach zero degrees at x = 15 cm.
6.5.4. Solve Exercise 6.5.3 if the rod is made of steel.
222 Chapter 6: The Calculus of Fourier series
1. Determine that the PDE and BCs are linear and homogeneous.
2. Assume that the solution is a product of two functions. One function is
possibly of time, while the other function is composed of spatial variables.
3. Substitute this product into the PDE and move all functions of the time
variable to one side, while placing all other components on the opposite side
of the equation.
4. Set both sides equal to the same separation constant, thus breaking the orig-
inal equation into two new equations and separating the variables.
5. Repeat steps (2), (3), and (4) on equations with more than one spatial variable
until you have an ODE for each spatial variable.
6. Substitute the product form of the solution into the BCs to obtain BCs for
the spatial ODEs.
7. Solve the ODE boundary value problem(s) to obtain the eigenvalues and eigen-
functions.
8. Solve the time ODE using the previously found eigenvalues.
9. Use the principle of superposition to get the general solution as an infinite
series.
223
224 Chapter 7: Separation of Variables: The Nonhomogeneous Problem
10. Use the ICs and orthogonality to determine the constants for the specific
solution.
u(O,t) = 0
J
u(L,t) = 0.
Also, there must be an IC, and we model it as
u(x, 0) = f (x).
The general technique for solving these types of problems is discovered by re-
membering how we solved a linear second-order nonhomogeneous ODE: First, solve
the homogeneous part. Second, solve the particular part. However, solving the
Section 7.2: Nonhomogeneous PDEs with Homogeneous BCs 225
0126
at -' 012 ZG
ax2
G'(t)q5(x) = kG(t)q"(x).
Separating variables and setting the resulting equation equal to a separation con-
stant, we obtain
G'(t) - qY'(x) -
kG(t) fi(x)
qY'(x) -
fi(x)
(L) =0.
We now have two simple ODEs: a spatial ODE with BCs and a time ODE. Up
to now, this is familiar, but here is where we enter new territory. Basically, we
throw away the time ODE and only solve the spatial ODE for its eigenvalues and
eigenfunctions.
The eigenvalues and eigenfunctions that form the solution to the spatial ODE
can be found in Chapter 5. They are
nor 2
eigenvalues: An =
L
n=1,2,3 , .... (7.5)
n7rx
eigenfunctions: q5( x) = cn sin L
226 Chapter 7: Separation of Variables: The Nonhomogeneous Problem
Step 2: Assume that the general solution for Equation (7.1) should be a Fourier
series. In particular, we assume that it will be a Fourier series expanded in the
eigenfunctions, Equation (7.5), multiplied by some function of time, t. Therefore,
assume the general solution for Equation (7.1) is
00
nzx
u(x, t) = cn (t) sin
L
Notice in Equation (7.6) the former constant, cn, is now required to be a function
of time, t.
Step 3: Start working on the particular part by expanding Q(x, t) from Equation
(7.1) in a Fourier series using the eigenvalues and eigenfunctions that were found in
Step 1. This means,
00
n7rx
Q(x, t) _ do sin L
n=
Since we are already using the name cn in Equation (7.6), we selected the constants
do for Equation (7.7).
We must determine the constants do . From Chapter 4, we know
2 L
d72=-
L Q(x, t) sin L dx.
0
From Chapter 5, we know that do must in fact be a function of time, t, and only
constant with respect to x. Thus, if we assume that we can perform the integral
and let an (t) = dn, we can write Equation (7.7) as
00
Since the BCs are homogeneous and we assume the rod to be continuous, we
have the Fourier sine series as a continuous function. Thus, if we differentiate the
Fourier sine series with respect to x, we have
00
nLx 11
cn (t) sin (nLx) I = cn (t) (sin C
8x 8x ll
00
nor n7rx
L cn (t) cos L
n=1
Since the BCs are homogeneous and we assume the heat flow to be continuous, we
have the Fourier cosine series as a continuous function. Thus, we can differentiate
with respect to x the Fourier cosine series. This yields
a2u a °°
ax2
=
ax n-1
[1cn(t)cos()]
12?C
L
n?Cx
00
_
n=1
L cn (t) (cos
())
ffl7r
\ L / 2 cn (t) sin (nLx) .
Step 5: Solve Equation (7.1) by replacing the terms of the equation by what we
have determined they equal in Steps 1 through 4. That is,
au _ a2u
k
at ax2 + Q(x,t)
becomes
00 00
00
nix
an (t) sin (-;-).
n=1
We now have the Fourier sine series expansions over the same interval, [0, L], in
all three terms of the equation. This means that the Fourier coefficients must be
equal. Therefore, we only have to solve the ODE
2
c(t) = -k () c(t) + a(t).
228 Chapter 7: Separation of Variables: The Nonhomogeneous Problem
nor 2
Since L = we may write
Equation (7.11) is a first-order ODE. To develop the solution for cn (t), we use
the integrating factor
t.
Note that the integral does not have a lower limit. We assume that when both sides
of the equation are integrated, we will get constants of integration on both sides of
the equation, which we combine and call b.
Step 6: Write the solution to Equation (7.1) and apply the IC, Equation (7.3).
We have
t
u(x, t) = [e_nt ekanTan(r) dr + e-kantbn sin n7rx (7.12)
L
n=1
Therefore, the complete solution to Equation (7.1), subject to the BCs, Equa-
tions (7.2), and IC, Equation (7.3), is
u(x, t) _
n= 1
[et t
ekanTan(r) dT +- e-kantbn sin n7rx
L
,
Jx ,
Q(x) _
2r -x r < x2
< ir, ,
subject to
u(o, t) = o
(7.15)
{ u(2ir, t) = 0
and
u(x, 0) = x2 - 27rx
8u(x, 0)
l = 0.
at
Find the time-dependent solution for u(x, t). For this problem, the BCs are linear
and homogeneous. However, the PDE in Equation (7.14) is not homogeneous.
Therefore, we need to follow the method developed in this section.
Step 1: Solve the related homogeneous problem. We basically ignore the term Q(x),
and we derive the eigenvalues and corresponding eigenfunctions for
012 u 012 u
at2 01x2
subject to the BCs, Equation (7.15). Remember, we are only solving for eigenval-
ues and corresponding eigenfunctions of the spatial problem after we separate the
variables. Thus, we must solve
o"(x) = -A(x). (7.17)
Equation (7.17), subject to the BCs indicated in Equation (7.15), has eigenvalues
and corresponding eigenfunctions of
nx
Pn (x) = bn sin
2
Step 2: Assume the general solution for Equation (7.14) has the form
00
Step 3: Expand the term Q(x) in a Fourier series using the eigenvalues and corre-
sponding eigenfunctions found in Step 1. That is,
00
nx
Q(x) _ an sin 2) .
n=1
In Equation (7.19), the an term is not a function of time, t. The reason for this is
that the function Q(x) is only a function of x. Solving for an yields
f2
an = J Q(x)sin(2) dx
0
2r
nor
8 sin 2
Therefore,
nit
00 8 sin nx
Q(x) _ -J rn2
2 sin --).
2
(7.20)
n=
a2 u a2 u
Step 4: Using Equation (7.18) find ate and axe . Remember we must be able to
justify term-by-term differentiation. For the derivatives with respect to time, t, we
expect time to be continuous. Thus,
00
a2u nx
ate
n (t)sin().
2
(7.21)
n=
Since the string is continuous and the BCs, Equation (7.15), are homogeneous, we
know that we can perform term-by-term differentiation with respect to the spatial
variable, x. It is
a2 u oo
ax2
= - -n 2 bn (t) sin
2
nx
2
(7.22)
n=1
Step 5: Solve Equation (7.14) by using Equations (7.20, 7.21, and 7.22). We have
00 00
()2
Since the Fourier series are all in the same eigenvalues and corresponding eigen-
functions over the same interval, we know the Fourier coefficients must be equal.
Therefore,
nit
fl\2 8 sin ( 2
b(t) = - () b(t) + itfl2
or, written in a more familiar form,
bii t fl \2 8 sin (2 )
() + () b(t) =
This is a second-order linear nonhomogeneous constant coefficient ODE, with a
solution of
b (t) =ate,
cos/
I
nt l
2
/\
)+dnsin)+ I
nt
2
32in
s
n4
7rn4
nit
2 )
Step 6: Write the solution for u(x, t) and apply the ICs (Equation (7.16)). We have
n=1
ate, cos 12 I + do sin (2
nit
)
32sin(_-)1
itn4 2 jsin(). nx+
2
u(x, 0) = x2 - 2?Cx =
00
n=1
[aj +
flit
32sin()]
2
n
ur4 s).
in(2
Thus,
nit
- -32 sin 2 2ir
nx
an -
in4
+ it-1 (x2_2itx)sin
2
dx
o
The general method and problem we have solved in this section, nonhomoge-
neous PDEs with homogeneous BCs, is the basis for the other nonhomogeneous
problems we will solve later in this chapter. So, it is well worth your time to fully
understand the method. My suggestion is to work through the problem by picking
a value for k, such as 1, and some simple function for Q(x, t), such as xt. After
solving the problem with values for k and Q(x, t), you will notice that we only add
a couple of steps to the solution method of separation of variables, from the chapter
introduction. At the end of this chapter, we summarize the steps to follow when
solving the nonhomogeneous problem by the technique of separation of variables.
EXERCISES 7.2
subject to
x, 0<x<ir
BCs: and IC: u(x,0) _
2ir - x it < x < 2ir.
Find the time-dependent solution for u(x, t), and using your favorite mathe-
matical software, plot your solution. By inspection, determine if the graph of
your solution is reasonable for the problem.
7.2.4. Consider the PDE
z 2 z
c + Q(x, t).
ate - ax2
Find the time-dependent solution for u(x, t) subject to the ICs
5u(x' 0)
u(x, 0) = f (x) and = g(x),
and the following sets of BCs:
234 Chapter 7: Separation of Variables: The Nonhomogeneous Problem
5u(0'
(2) t) = 0 and au(L' t) = 0.
8x 8x
(3) a 0' t) = 0 and u(L, t) = 0.
7.2.5. Consider the PDE
92u 52u
ate axe -}- x sin t,
=o
and to the ICs
-x, 0<x<1.5 8u(x, 0)
u(x, 0) _ and at =0.
5x-9 1.5<x<2
Find the time-dependent solution for u(x, t), and using your favorite mathe-
matical software, plot your solution. By inspection, determine if the graph of
your solution is reasonable for the problem.
7.2.6. Consider the PDE
Du D2u
16 -}- y sin t,
at a2
subject to the BCs
8u(0, t) _ 0
ay
and u (2 t) = 0 ,
and IC
u(y, 0) = y
Find the time-dependent solution for u(y, t), and using your favorite mathe-
matical software, plot your solution. By inspection, determine if the graph of
your solution is reasonable for the problem.
7.2.7. Consider the PDE
Du D2u D2u -t
y) e
Dx 2 + p`l y 2 + `x + ,
Dt
and IC
u(x,y,0) =x+y.
Find the time-dependent solution for u(x, y, t), and using your favorite mathe-
matical software, plot your solution for several different times t. By inspection,
determine if the graph of your solution is reasonable for the problem.
7.2.8. Consider the PDE
a2a a2u a2u
ate axe + a 2 + sin t cos xy,
y
subject to the BCs
au(ir, y, t) = 0
ax '
u ,y,t
2
=0
au(x, 0, t) _ 0
u(x, 2ir, t) = 0
ax '
' au(x, 0)
u(x,0) = andd =0.
at
2 (2ir x) 71 <x < 27r
71 2 -
Find the time-dependent solution for u (x, t), and using your favorite mathe-
matical software, plot your solution. By inspection, determine if the graph of
your solution is reasonable for the problem. This problem reflects vibrations
which arise from a gravity field in a medium with resistance proportional to
the velocity; the ends of the string are fixed at the same height.
236 Chapter 7: Separation of Variables: The Nonhomogeneous Problem
Find the time-dependent solution for u(x, t), and using your favorite mathe-
matical software, plot your solution. By inspection, determine if the graph of
your solution is reasonable for the problem. This problem reflects a vertical
flexible rod of length it (with its left end rigidly attached to a freely falling
lift, while its right end is free) that, having attained a velocity of 100 ft/sec,
stops instantaneously.
7.2.11. Given a one-dimensional rod of length it m. A heater moves along the sur-
face of the rod with constant velocity of 1 cm/min. Thus, a convective heat
exchange takes place. The flow of heat from the heater to the rod is 6e-at
watts where $ is the convection coefficient. Suppose the ends of the rod are
held at 0°C, and the initial temperature distribution is 3 sin(5x).
(1) Mathematically model this problem.
(2) Suppose $ _ .12, and solve the time-dependent problem.
(3) Graph the time-dependent problem.
(4) Solve the steady-state problem.
has some type of insulation, which constricts the flow of heat to a constant value of
A2. The mathematical model for this problem is
z
(7.24)
8t 8x2 >
8u(L, t) _ A2
8x
with IC
u(x,O) = f(x). (7.26)
In the last section, we said the nonhomogeneous PDE with homogeneous BCs
problem was the basis for all the other nonhomogeneous problems we will solve
in this chapter. However, how is a homogeneous PDE with nonhomogeneous BCs
turned into the basic nonhomogeneous problem? The answer is by creating a new
function. So what does this new function look like? To answer this question, we
must consider all the things that the new function must do.
The new function must have homogeneous BCs. It must replace the function
u(x, t) with no loss of information. Finally, the new function must turn the non-
homogeneous BCs into a source term that makes the PDE nonhomogeneous. The
question that still remains is how? To answer this question, consider what we have
learned so far in the solution techniques for PDEs, particularly PDEs that involve
temperature distribution in a one-dimensional rod. So far, two techniques have
been studied: separation of variables and steady-state. We know that we cannot
use the separation of variables technique yet. So, by default, that leaves us with the
steady-state technique. Let's investigate this method to determine what it gives us.
Basically, in the steady-state problem, we assume that u(x, t) = uSS (x) This
.
a 2uSS = d 2uSS .
assumption implies that cu ss = 0 and The general steady-state
at axe dx2
solution for Equation (7.24) is
u38 (x) = C1 X + C2.
(7.27)
We now want to create a new function, using Equation (7.28), that has homo-
geneous BCs. Suppose we create a function v(x, t) such that
v(x,t) = u(x,t) - uss(x). (7.29)
Then
v(0, t) = u(0, t) - u33(0) = T1 - T1 = 0
238 Chapter 7: Separation of Variables: The Nonhomogeneous Problem
and
av(L,t) _ au(L,t) - &u33(L)
=A2-A2=0.
ax ax ax
Therefore, v(x, t) is a function that has homogeneous BCs. Ideally, we'd like to
replace u(x, t) with v(x, t) in Equation (7.24). To do this, we must determine
au D2 u
u(x, t), , and axe . From Equation (7.29), we have
at
u(x, t) = v(x, t) + u33 (x). (7.30)
Using Equation (7.30) and Equation (7.28), we find that
au av 9u88 (x) _ av a(A2x + T1) _ av
- at (7.31)
at - at + at - at + at
and
a2Z.6 a22J a2uss (x) a2v a2 (A2x +T1) aax2v
. (7.32)
axe + axe
axe axe + axe e
Thus, u(x, t) and v(x, t) are interchangeable in Equation (7.24). However, instead
of ending up with a nonhomogeneous PDE with homogeneous BCs, we have a
homogeneous PDE with homogeneous BCs, as summarized below:
av a2 v
(7.33)
at - ax2 ,
subject to
v(0, t) = o
(7.34)
av(L, t) _ 0
ax
We know from our studies in Chapter 5 that Equation (7.33), subject to Equation
(7.34), has the solution
r (2n- 1) irx 1
sin (7.35)
L 21'
n=1
r ,n=1,2,3,....
n= [(2n_1)]2
2L
However, Equation (7.35) only gives us the solution for v(x, t), not u(x, t). Equation
(7.30) and Equation (7.35) determine the general solution for u(x, t). Thus, the
general solution is
u(x, t) = v(x, t) + u33(x)
To avoid any confusion as to the contents within the summation sign, this may be
rewritten as
00
u(x, t) = A2x + Tl - bne-t sin [[ (2n 1
. (7.37)
n= 1 J
Applying the IC
u(x,0) = f(x),
Thus, the specific solution to the problem in Equation (7.24), subject to the
BCs, Equation (7.25), and the IC, Equation (7.26), is
00
(2n - 1) 7rx
u(x't)=A2x+T1 + >bn e-pant sin 2L
n=1
and
Applying the first boundary condition, u (O) _ -1, yields c2 = -1. Applying the
second boundary condition, u(it) = 1, yields c1 = 2 . Therefore, the specific steady-
state solution is
2
u33 (x) _ -x - 1. (7.42)
71
Letting the function v(x, t) = u(x, t) - u33 (x), we find v(0, t) = 0 and v(ir, t) = 0.
Also,
av au
at -
2v 52u
.
axe axe
Thus, we have linear homogeneous PDE and BCs in terms of v(x, t). We can now
apply the separation of variables technique. We should recognize from Chapter 5
that
av a2 v
3
at = axe
subject to
v(0, t) = o
v(ir, t) = 0,
has, as a general solution,
00
bne-3n2t
v(x, t) _ sin nx.
n=1
Therefore,
00
u(x, t) = -x - 1+ bne-3n2t
sin nx.
71
n=1
Using orthogonality of the Fourier sine series provides the value for bn, which is
u(x, t)
00
00
98(n- 2-(_e_3n2tsinnx;
7rn (n 49)
n=8
242 Chapter 7: Separation of Variables: The Nonhomogeneous Problem
or it may be written as
6 n
u(x, t) _ x -1 + 98 n 4g) e-32t sinnx
00
+
98(n-(-1)) e -3n2t sinnx.
n=8
urn (n2 - 49)
In Figure (7.2), the solution for u(x, t) is shown for 0 < t < 0.05. Please note the
boundaries are at -1 when x = 0 and 1 when x = ?C. Also, the initial temperature
distribution is indicated by a bold-faced curve.
Now, a question that should come to your mind is: The method works great
for the heat equation, but how does it apply to the wave equation? This is a very
reasonable question considering the wave equation does not have a steady state
solution. But, the wave equation does have an equilibrium, and the equilibrium
state for the wave equation is equivalent to the steady-state solution of the heat
equation. Thus, this technique will work with the wave equation.
Although the homogeneous PDE with constant nonhomogeneous BCs is an in-
teresting problem, the homogeneous PDE with variable nonhomogeneous BCs is far
more interesting. We solve this problem in the next subsection.
the rod. The approach we will take is called construction. As the name suggests, we
will construct a function v(x, t) such that v(x, t) = u(x, t) - r(x, t), with v(0, t) = 0
av(L t)
and = 0. This, in turn, requires r(x, t) to have the following properties:
ax
r(0, t) = a(t)
(7.46)
ar(L,t)
8x
-
Using Equation (7.46) as a starting point, we realize that integrating both sides
ar(x t)
of = $(t) with respect to x yields r(x, t) = x$(t) + c(t). (Note: As
ax
far as the variable x is concerned the function c(t) is a constant.) Furthermore,
applying the first BC, r(0, t) = a(t), we find a(t) = c. Thus, it would seem
reasonable to assume that r(x, t) = x$(t) + a(t). This representation of r(x, t)
yields the necessary properties described in Equation (7.46). However, there are an
infinite number of functions r (x, t) that may be generated or picked. For instance,
rx
r (x, t= a(t) x cos L+ cos x- -$(t) sin will also be a continuous function
that satisfies the necessary properties in Equation (7.46). Although r1(x, t) will
work, it is a much more complicated equation than r(x, t) = x$(t) + a(t), which we
will now use. We may now construct v(x, t) and replace u(x, t) in Equations (7.43
and 7.44) to yield
at - at + (x$'(t) + a '(t)),
a2 u .
while axe yields
2L [(2n_1)7r12 n= 1,2,3,...
nlx = Cn, Slri [/Xx]
Step 2: Assume the solution for v(x, t) is a Fourier series in the eigenfunctions
just found. We have
where An =
[(2n1) 2
where
L 1
av a2 v
Step 4: Find and axe . Remember, this requires us to satisfy all of the con-
at
ditions necessary to do term-by-term differentiation of a Fourier series. Assuming
all the conditions were met, we obtain
av °O
_ c( t) sin [\/x] (7.52)
n=1
and
00
a2v
axe =-: n=1
AnCn (t)Slll [\/cx] . (7.53)
Step 5: Solve Equation (7.49) by replacing the terms of the equation with what
we determined they equal in Steps 1 through 4. Thus,
av a2 v
_ (x$'(t) + a (t) )
at ax2
becomes
00 00 00
c' (t) Slri [\/cx] = -1C Ac (t) Slri [\/ix] - Cln,(t) Slri [\/cx]
n=1 n=1 n=1
We now have the Fourier series expansions with respect to the same eigenvalues over
the interval [0, L], in all three terms of the equation. This means that the Fourier
coefficients must be equal. Extracting the coefficients yields the ODE
c(t) _ -1CAnCn(t) - an(t)
or
Note that the integral does not have a lower limit. We assume that when both sides
of the equation are integrated, we get constants of integration on both sides of the
equation, which we combine and call b.
Step 6: Write the solution to Equation (7.49). We have
t
v(x, t) _ -e-leant ekanTan(T) dT + bne-karat
sin nx
n=1
246 Chapter 7: Separation of Variables: The Nonhomogeneous Problem
We cannot apply the initial condition since this is the solution to v(x, t), and we
want the solution to u(x, t). But, from Equation (7.47) we know
Substituting, we get
u(x,t) =x$(t)+a(t) +
n-1
[_et t
dT + e-kantbn sin [x]
where bn is given by Equation (7.56).
The following example illustrates this method, and it provides a graphic repre-
sentation of the solution.
EXAMPLE 7.3. Find the time-dependent solution for u(x, t) if
8u 82u
(7.57)
8t 8x2'
Section 7.3: Homogeneous PDE with Nonhomogeneous BCs 247
subject to
u(0, t) = a(t) = e_t
(7.58)
/71 \
u(,t) =$(t)=t
and
u(x, 0) = cos 3x. (7.59)
The PDE in Equation (7.57) is linear and homogeneous. However, the BCs, Equa-
tion (7.58), are linear but not homogeneous. Thus, to use the separation of variables
technique, we must develop a new PDE in terms of v(x, t), where
v(x, t) = u(x, t) - r(x, t). (7.60)
This means the function r(x, t) must equal e-t when x = 0. It must equal t when
x = . The BCs in Equation (7.58) suggest that r(x, t) = a(t) + x (3(t) - a(t)) =
2 L
e-t + 2x t - e_t . (Note: Although r x t is one of the easier continuous functions
to work with, there are an infinite number which may be found.) Thus, r(0, t) = e-t
andr r2,t =t.
Since v(x, t) = u(x, t) - r(x, t), we have v(0, t) = 0 = v
r2 We must
(,t). now
replace u(x, t) in Equation (7.57) with v(x, t). Using Equation (7.60) yields
Thus,
au
at
=-+-=
av ar av - e -t + 2x 1 +
at at at ?C (
e
_t
)
and
subject to
v(O,t) = 0
(7.62)
v ,t2 =0,
2x
where the source term Q(x, t) equals e - t -
r ( 1 + e - t). We now have a linear P DE
with linear and homogeneous BCs.
Step 1: Solve the related homogeneous problem, which is
av a2 v
at ax2
subject to
v(0, t) = o
r =0.
v 2,t
Solving the related homogeneous problem for eigenvalues and eigenfunctions yields
n = 4n2
,n=1,2,3,...
n (x) = sin 2nx J
Step 2: Assume the solution for v (x, t) is a Fourier series in the eigenfunctions
described earlier. We have
Step 3: Expand Q(x, t) in a Fourier series using the eigenfunctions found in Step 1.
This yields
00
e
-t - 2x (1 + e-)
t
= an(t) sin 2nx.
71
n=1
an (t) = -
o
2
(e_t - x (1 + e-t) sin 2nx dx
=-(e-t+ (-i)).
2
Section 7.3: Homogeneous PDE with Nonhomogeneous BCs 249
Therefore,
o0
e
-t - 2x (1 + e-t_t _ an (t) sin 2nx
?f
n=1
00
av a2v
Step 4: Using Equation (7.63), find and axe. Remember, this requires us to
at
satisfy all of the conditions necessary to do term-by-term differentiation of a Fourier
series. Assuming that all the conditions are met, we have
av °O
_ bn (t) sin 2nx (7.65)
at n=1
and
00
a2v
_ -4n2bn(t) sin 2nx. (7.66)
axe n=1
Step 5: Solve Equation (7.61) using Equations (7.64, 7.65, and 7.66). Thus,
av _ (92?)
+e
_t
- 2x (1+e _t )
at axe
becomes
00 00
00
In Equation (7.67), all the Fourier series expansions are over the same interval [o,
2
and are expanded in the same eigenfunctions. Since Fourier series representation
of a function is unique, we may solve the ODE formed by the Fourier coefficients,
which is
Therefore,
v(x, t) =
00 14ne_t + (4n2
n2(4n2 -1)71 - 1)(-1)n
+ dne -4n2t sin 2nx.
n= 1
Thus, the general solution for u(x, t) is
u(x,t) = r(x,t) + v(x,t)
=e_t + 2x
r (t - e
-t )
00 4n2e-t
1)(-1)n
+ + (4n2 - +dne -4n2t sin 2nx. (7.69)
n=1
n (4n 2 -
2
Equation (7.69) indicates the general solution for u(x, t). We now find the specific
solution for u(x, t) by applying the IC (Equation (7.59)), u(x, 0) = cos 3x. Thus,
u(x, 0) = cos 3x
_ 2x °° + (4n2 - 1)(-1)n
4n2e-t
-4n2t sin 2nx,
1- r + n=1 n2(4n2 _ 1)71 +dne
or
00
cos 3x - 1 +
n=1
n2(4n2 - 1)+ dnJ sin 2nx. (7.70)
4 2
u(x, t) = e-t + 2x (t - e -t )
r
00
Figure 7.3: The graph of the solution for u(x, t) for 0 < t < 2.
A graph of the solution is given in Figure (7.3) for 0 < t < 2, and the initial
temperature distribution is indicated by a bold-faced curve. Note how the BCs
affect the surface as time progresses.
In the previous two sections, we solved the problems of the nonhomogeneous
PDE with homogeneous BCs and the homogeneous PDE with nonhomogeneous
BCs. Both of these problem types are useful in solving physical models. More
importantly, they help us solve the more general problem of a nonhomogeneous PDE
with nonhomogeneous BCs. The nonhomogeneous PDE with nonhomogeneous BCs
applies to many more physical models. We address this general problem in the next
section.
EXERCISES 7.3
8u(0, t) au(L, t)
(1) = a and
8x
=3.
8x
t)
(2) u(0, t) = a and =8.
_
(3)
8u(0, t)
ax - a(t) and 8u(L,
ax
t)
- ,Q(t)
subject to
7.3.3. Liquid is input at time t = 0 by the amount of 3 drops/sec into the right end
of a tube of length 2. The left end of the tube is connected to a large tank
in which liquid pressure remains invariant. Assuming that, until the change
of input at the end x = 2, the pressure and input to the tube are constant,
find the change of input into the tube for t > 0 and the change in pressure in
the section x = 2 for t > 0. The boundary value problem that governs this
system is
ap _ aw
ax at - 2w
0<x<2,0<t,
ap aw
at - ax
subject to
p(O,t) = 0 and w(2, t) = 3 drops/second
and
w(x, 0) = 0
p(x,0) = 0.
Section 7.3: Homogeneous PDE with Nonhomogeneous BCs 253
(Hint: Eliminate p(x, t) in the equations previously listed and obtain a second-
order wave equation in w(x, t). If you are unsure how to eliminate p(x, t),
review Chapter 2.)
7.3.4. Consider the mathematical model
a2u a2u
=4
at2 ax2
subject to
u(0, t) = 6 and u(ir, t) = 1
and
au(x, 0)
u x 0 = 6 cos x ? x and =0.
Find the time-dependent solution for u(x, t), and using your favorite mathe-
matical software, plot your solution. By inspection, determine if the graph of
your solution is reasonable for the problem.
7.3.5. Consider the mathematical model
a2u 2 a2u
=C
ate axe
subject to
u(0, t) = coswt and u(L, t) = 0
and
au(x, 0)
u x 0= 0 and =0.
at
Find the time-dependent solution for u(x, t).
7.3.6. Consider the mathematical model
a2 u a2 u
= 16 axe
ate
subject to
u(0, t) = 0 and u(5, t) = sin 3t
and
subject to
Du(L, t
u 0 t = 0 and = sin wt
(' ) ax
and
u(x, 0) = f (x).
Find the time-dependent solution for u(x, t).
7.3.8. Consider the heat equation
au 1.14a2u
at - axe '
subject to
au(0, t) = au(7, t)
sinwt and =0
ax ax
and
-x, 0<x<2
u(x, 0) = -.5x2 + 4.5x - 9, 2 < x <- 5
1, 5<x<7.
Find the time-dependent solution.
7.3.9. Consider the PDE
av
a2v
axe
-LC a2v
ate - (RC + GL)
at
- GRv = 0,
8v(8, t)
ax = 0 (perfect insulation)
and
av (x, 0
(')
v x 0 = 0 and
at
=0.
Section 7.3: Homogeneous PDE with Nonhomogeneous BCs 255
_ + sx
subject to
u(o, c) = o
BC: and IC: u(x, 0) = 9xex-4.
u(4, t) = 36
Find the time-dependent solution for z (x, t), and using your favorite mathe-
matical software, plot your solution. By inspection, determine if the graph of
your solution is reasonable for the problem.
This problem reflects the input of liquid into the right end of a tube of length
10 ft dropping at t = 0 by an amount of 5 gals/sec. The left end of the tube
is attached to a large tank in which the liquid pressure remains invariant. If
you assume that until the change of input into the right end, the pressure
and input were constant, we can find the change in input for t > 0 using the
equations listed previously, and the change in pressure in the section x = 10
for t > 0 by the following equation
Dp _ Dz
+ 2z .
Dx Dt
jC ax +
z
252u
Q(x,t)> (7.7z)
subject to BCs
u(0, t) = A
(7.73)
8u(L, t) = B(t)
8x
and ICs
u(x,0) = f(x)
(7.74)
8u(x,0)
at
Equations (7.72-7.74) contain parts of every problem we have discussed in this
chapter to this point. We have a nonhomogeneous PDE, and the nonhomogeneous
BCs contain both a nonzero constant and a function of time t. Therefore, to solve
this problem a solution strategy is mapped out in the following steps:
1. Replace u(x, t) with a function v(x, t) which has homogeneous BCs. This
8u(L , t)
requires constructing a function r(x, t) so that u(0, t) -r(0, t) = 0 =
ax
ar(L, t)
ax
2. Solve the related homogeneous PDE in v(x, t) for eigenvalues and eigenfunc-
tions.
3. Assume a Fourier series solution for v(x, t), where the coefficient in the Fourier
series is a function of time t.
4. Perform an eigenfunction expansion of the time-dependent function, Q(x, t),
using the eigenfunctions found in Step 2.
5. Determine the derivatives of the Fourier series solution for v (x, t).
6. Replace the PDE by its equivalent Fourier series representation and determine
the ODE, which must be solved.
Section 7.4: Nonhomogeneous PDE and BCs 259
Using Equation (7.78), we can replace u (x, t) in Equation (7.72) with v (x, t). This
a2u a2u
requires the determination of ate and axe , since
a2u a2v
xB (t) (7.7s)
ate =ate +
and
a2u a2v
axe ' (7.80)
Dx2
260 Chapter 7: Separation of Variables: The Nonhomogeneous Problem
2 2 2
ate
=c + Q(x, t)
becomes
a2v a2v
c2 axe +Q(x,t) - xB ii(t), (7.81)
ate =
subject to BCs, Equation (7.77).
Step 2: Solve the related homogeneous problem in terms of v(x, t) for eigenvalues
and eigenfunctions.
The related homogeneous problem is
a2v a2v
5=C 2
axe
subject to BCs, Equation (7.77). After separating variables and applying the BCs
to the spatial problem, we conclude that
r (2n - 1J
)=
n I
ZL
n=1,2,3 ,.... (7.82)
r (2n-1)irxl
cPn,(x)= bn, Sln
2L
Step 3: Assume a Fourier series solution for v(x, t), where the coefficient in the
Fourier series is a function of time t.
We have
[(2n - 1)irxl
v(x, t) _ b(t) sin
n =1
2L
where, after integration, the constant do becomes a function of time t. Thus, as-
suming we can perform the required integration, we have
Step 5: Determine the derivatives of the Fourier series solution for v (x, t).
Remember, we must satisfy all the conditions that allow term-by-term differ-
entiation of a Fourier sine series, which is covered in Chapter 5. Assuming these
conditions are satisfied, we have
00
av(x,t)
ax
_ Ld
n=1
nbn(t)COS 1 nx) .
a2v(x, t) °O
at2
= L1 Ul) Sill V AnJJ.
n=1
Step 6: Replace the PDE by its equivalent Fourier series representation and
determine the ODE that must be solved.
Using Equations (7.84, 7.85, and 7.86), Equation (7.81)
a2v a2v
c2 + Q(x, t) - xBii(t)
at2 = ax2
becomes
00 00
Equation (7.87) equates one Fourier series to the sum of two Fourier series
all expanded in the same eigenfunctions over the same interval. This means the
constants must be equal. Extracting coefficients yields
bn(t) = bnh (t) + bnP (t) = cYn cos [c/t] + ,3n sin [c'5t] + bnP (t). (7.89)
Step 8: Replace the coefficient in the Fourier series representation for v(x, t)
with the solution from the ODE.
From Equation (7.83), we have
00
tl
[an cos [C ant] + ,Qn Sin [ct] + bnp (t)] sin (x)
Step 9: Determine the solution for u(x, t) from the solution for v(x, t).
Equation (7.78), which is
+
n=1
Lan cos[ct] + n sin [ct] + bnp (t)J sin (x)
Step 10: Use the ICs and orthogonality to determine the constants in the general
solution of u(x, t).
Equations (7.74) tell us
u(x,0) = f(x)
8u(x,0) _ g(x)
at
Section 7.4: Nonhomogeneous PDE and BCs 263
t
8u(x, 0)
= g(x) = xB' (0) +
n=1
°°
[/3n + b'ap (0)] sin (x).
Therefore, in a similar method as described earlier, we have
fL
/3n = L [g( x) - xB'(0)] sin ( x) dx - b(0). (7.93)
We now have the complete solution for u (x, t) from Equation (7.72) subject to
BCs, Equation (7.73), and ICs, Equation (7.74). It is
u(x, t) = xB(t) + A
+
n=1
Lan cos [ct] + /3n sin [ct] + bnp (t)] sin (x),
where an and /3n are given in Equation (7.92) and Equation (7.93), respectively,
and An is given in Equation ( 7.82) .
EXAMPLE 7.4. Consider
au a2u
- xt, (7.94)
at = axe
subject to
au(0, t) _ 0
8x
(7.95)
8u(7r, t)
8x
-t
264 Chapter 7: Separation of Variables: The Nonhomogeneous Problem
and
11x .
u x 0) = sin (7.96)
) 2
Solve for the time-dependent solution using the ten steps developed in this section
(Nonhomogeneous PDE with Nonhomogeneous BCs).
Step 1: Replace u(x, t) with a function v(x, t).
x2t
Let v(x, t) = u(x, t) - r(x, t) where r(x, t) = which implies u(x, t) = v(x, t) +
r(x, t). Therefore,
au av x2
+ (7.s7)
at at 271
and
a2 u a2v t
ax2 ax2 + 71.
av x2 a2v t
at + ax2 + - xt
or
- x t--
av a2v 1 x2
(7.99)
at ax2 -
which has the following BCs
av(0, t) _ o
8x
av(71,t)
-0
ax
av a2 v
(7.101)
at = ax2
subject to BCs, Equations (7.100). We have solved for the eigenvalues and eigen-
functions several times in the past chapters. The eigenvalues are
(7.103)
an
ancoscosnxnx , n> 0.
Step 3: Assume a Fourier series solution for v(x, t) where the Fourier coefficients
are functions of the variable t.
The Fourier series solution is
1 x2
C -x) t- 2 = bo(t) + bn(t)cosnx,
n=1
where
o \ /
1 7C 71
and
x2
b(t)=-r2 o
-?r1 - x t - 271
cos nx dx
x2 _ 1 7C 71
t 271 7r 2 t- 6
+
00
2
-i [(1 - (-1)") t - (-1)] cosrtx. (7.105)
n=1
266 Chapter 7: Separation of Variables: The Nonhomogeneous Problem
Step 5: Determine the derivatives of the Fourier series solution for v (x, t).
Using Equation (7.104) and assuming the required conditions hold, we have
00
8v(x, t)
at - a o (t) + an (t) cos nx (7.106)
n=1
and
00
a2v(x,t)
= -n a (t) cos nx. (7.107)
8x2 n=1
Step 6: Replacing
av a2v 1
_x t_ x2
at = axe +
with its equivalent Fourier series representation yields
From Chapter 4 we know that two Fourier series that are equal must have equal
coefficients. Therefore, Equation (7.108) can be reduced to the ODEs
ao(t)=
r ---
2
t--6 1 7r 71
(7.109)
and
an(t)
2
71n4 (1 (-1 )fl) Ct - 2
n
I - (_i)fl] +dne-n"zt (7.113)
Step 8: Replace the coefficients in the Fourier series representation for v (x, t).
Replacing ao (t) and an (t), given in Equations (7.111 and 7.113), in Equation (7.104)
yields
2
v(xt)=(--)---+do+ 7rt
(7.114)
00
{ n 2
4
[(1 - (-1)n) (t n (_1)n + dne2
-n t cos nx.
Step 10: Determine the constants in the general solution of u(x, t) by using the ICs
and orthogonality.
Equation (7.96) indicates the initial temperature distribution in the one-dimensional
rod is u(x, 0) = 0. Therefore,
2 = d0+
u(x,0) =sin 5x
which implies
d0=-j
1 sin
11x dx_2
r o 2 1171
and
dn = 2 11 x
cos nx dx +
2 1 (i)) + (_i)fl]
71j0
sin
2 7rn4 n
44+ n4 r
1) )
n 1
+ (-1)nJ . (7.116)
r (4n2-121)
268 Chapter 7: Separation of Variables: The Nonhomogeneous Problem
2
urn
1- (-i)) n t- n - (_i)fl]+ d e-net
1 cos nx
n=1
au(x' O t)
= 0 and u(x, 27, t) = x°C, (7.120)
ay
Section 7.4: Nonhomogeneous PDE and BCs 269
and IC
r(x,y,t) = x. (7.122)
ar(x, y, t) _ 1 ar(x, y, t)
Since Equation (7.122) yields r (0 ,y,t) = 0, = 0 and
ax ay
r(x, 27r, t) = x. Thus, v(x, y, t) becomes
and
Rewriting Equation (7.118) in terms of Equations (7.123, 7.124, and 7.125) yields
and
av(x, 0, t)
= 0 and v(x, 27r, t) = 0. (7.128)
ay
270 Chapter 7: Separation of Variables: The Nonhomogeneous Problem
at = ax2 aye
and
and
where
r (2n-1
T =
L 2
and
r(2m-1)12
L 2 ,m1, 2,3,.
j
Therefore,
12
2 2 2 (2n-1) 2 (2m-1
n, m = 1, 2,3, .. .
2 + 2
Second, we must expand the new source term of Equation (7.126) in terms of the
eigenvalues and eigenfunctions found for the homogeneous problem. We have
°° °° 1(2m - 1)y [(2n - 1)x
sint = bnm(t) cos sin
4 2
n= 1 m= 1
Section 7.4: Nonhomogeneous PDE and BCs 271
where
2
2 2n 1)xl (2m
bn-m, (t = 2
si n
Slri t Sl
sin 2 ]cosL 4 1)y]dxdY
n
o n
o
16(_1)m+1 sint
ir2(2m - 1)(2n - 1
= bnm Sin t,
00 00
a2v(x,y,t) 2n - 1) 1`2 (2m-1)yl r(2n- 1)x1
ax2 L -art(t)
- n=1 m=1i
L 2 J
cos L 4
]sinL 2
and
y
y,t)
z
. 00
I
n=1 m=1
00
I -anm(t) I
1(2m_1)
L J
2 cos I (2m 4 1)yJ sin
L
(2n 2 1)x
L
Next, we put it all the pieces together and solve the resulting ODE. We have
00 00
00 00
n=1 m=1
-anen(t)
1(2n - 1)12
L 2 J
cos
r (2m - 1)yl
L 4 J
sin
L2
[(2n - 1)x1
°°
+ :i:°° -Clnm, t r(2m - 1)12 1(2m_l)yl (2n-1)x1
() I 4
j cos 4 J sin 2
n=1 m=1 J
+ bnm Sin t,
which yields the ODE
1)12
anm(t) _ -Cbnm(t) [r(2n2 1)J2 -Clnm(t) (2 + bnm Sin t (7.136)
L J
Therefore,
becomes
a2 sin t - cost
v(x,y,t) bnm
n=1 m=1 [( { A2+1
+ Cnme_ \2t
cos f (2m-1)yl s in f (2n-1)x 11
L 4 J L 2 JJ
We now may state the general solution for u(x, y, t), which is
u(x, y, 0) = 0 = x + v(x,y,0)
x+ r({-b+1}
nm
n=1 m=1 L
(2m4
+ Cnm)cos 1)yJ sin L(2n
L 2 1)x1J
or
00 00
(2m4 Lr(2n_
+Cn,m,l cos I 1)yJ sin 2 1)xJ
n=1 m=1 / L
which means
2
f2 f (-x) (2m (2n
Cnm = cos 4 1)yJ sin 2 1)x dxdy + + 1.
J
I
Jo o L L J
Thus,
32(-1)m+1(_1)n bnm
Cnm = (7.137)
(2m-1)(2n-1)22 + A2+1
Section 7.4: Nonhomogeneous PDE and BCs 273
CnmC-a2tCOS
(2m4 Lr(2n_
+ I 1)yJ sin
L 2 1)xJJ
where Cnm is given by Equation (7.137).
In the next section, we summarize the solution technique, called the eigenfunc-
tion expansion technique for solving a nonhomogeneous PDE problem by separation
of variables.
EXERCISES 7.4
subject to
u(x,O) = f(x).
Solve for a time-dependent solution given the following nonhomogeneous bound-
ary conditions:
au (0' au (L'
(1) t) = A and t) -B
au (L, t)
(2) u 0 t = A and =B.
ax
(3) u(0, t) = A(t) and u(L, t) = B(t).
au(L, t)
(4) u 0 t= A and = B (t).
a
7.4.2. Consider the wave equation
a221 207221
Q(x, t),
ate - 07x2 +
subject to
(1) Set up the mathematical model. Note: The thermal diffusivity of pure
iron may be found in Appendix E.
(2) Give a series solution of your mathematical model.
(3) Using your favorite mathematical software, plot your solution. By in-
spection, determine if the graph of your solution is reasonable for the
problem.
(4) Check that this solution satisfies the equation, the BCs, and the ICs.
and
and IC
u(x, y, 0) = 0.
7.4.8. Solve the wave equation in a rectangular sheet with Dirichlet BCs and a time-
periodic forcing function, cos t. Assume that the sheet is initially at rest with
displacement
(3) Using your favorite mathematical software, plot your solution. By in-
spection, determine if the graph of your solution is reasonable for the
problem.
(4) Check whether the solution satisfies the equation, the boundary condi-
tions, and the initial conditions.
7.4.11. The space shuttle heat shields are shaped as rectangular boxes with curved
surfaces matching the curvature of the body of the shuttle. The greater ma-
jority of the heat shields are surrounded by other heat shields that are fitted
tightly together. One surface is against the body of the shuttle, while the
opposite side faces space.
The typical heat shield may be modeled adequately by a rectangular box with
no curvature. The model is assumed to have perfect insulation on its four
sides, while the side against the shuttle is assumed to be held at 14°C. The
temperature of the side open to space at time of re-entry is proportional to
the time in seconds, required for re-entry. On re-entry into earth's atmosphere
this function is modeled by
7.5 SUMMARY
In this chapter, the problems solved by the eigenfunction expansion technique were
all linear PDEs with linear BCs. It is important to realize that this method only
works on linear PDEs with linear BCs.
We summarize our eigenfunction expansion technique as follows:
278 Chapter 7: Separation of Variables: The Nonhomogeneous Problem
(a) Assume the solution is a product of two functions. One function is pos-
sibly of time, while the other function is composed of spatial variables.
(b) Substitute this product into the PDE and move all functions of time to
one side while placing all other components on the opposite side of the
equation.
(c) Set both sides equal to the same separation constant, breaking the orig-
inal equation into two new equations, thus separating the variables.
(d) Repeat steps (a), (b), and (c) on equations with more than one spatial
variable until you have an ODE for each spatial variable.
(e) Substitute the product form of the solution into the BCs to obtain BCs
for the ODEs.
(f) Solve the ODE boundary value problem(s) to obtain the eigenvalues and
eigenfunctions.
9. If a constructed function was used, obtain the solution for the original function
of the original PDE.
10. Use the ICs and orthogonality to determine the constants for the specific
solution.
Chapter 8
The Sturm-Liouville
Eigenvalue Problem
8.1 INTRODUCTION
The Sturm-Liouville eigenvalue problem is a second-order homogeneous ODE with
variable coefficients subject to homogeneous boundary conditions. You should re-
member second-order ODEs produce eigenvalues and eigenfunctions for the solution
space where the ODE is defined. The study of this class of ODEs is important in
itself. However, applying this class of ODEs to PDEs is our primary interest.
Consider the following examples:
EXAMPLE 8.1. Given the heat equation
0721 C72 2l
at ax 2
subject to
u(o, t) = o
u(L, t) = 0
and
u(x, 0) = f (x),
we know the separation of variables technique applies. Letting u(x, t) = p(x)G(t),
taking the appropriate derivatives and separating the PDE yields the time equation
G'(t) = -AkG(t)
and the spatial eigenvalue problem
p"(x) = -Ap(x),
279
280 Chapter 8: The Sturm-Liouville Eigenvalue Problem
subject to
(p(0) = 0
p(L) =0.
EXAMPLE 8.2. Given the wave equation
a2u
=c
29u
ate axe
subject to
8u(0, t) _ 0
8x
8u(L,t) _ o
ax
and
u(x, 0) = f(x)
8u(x, 0)
at = 9(x)
we again apply the separation of variables technique resulting in the time equation
cp"(x) = -A,o(x),
subject to
We are not concerned with the solution for these examples. What we are con-
cerned with is the spatial problems, which resulted from using the separation of
variables technique. In particular, we notice the spatial equations, Equation (8.1)
and Equation (8.3), are identical. However, the BCs, Equation (8.2) and Equation
(8.4), are different.
Suppose we rewrote the BCs in the following way:
aip(0)+a2p'(O)=O
a3,o(L) + =0.
Section 8.2: Definition of the Sturm-Liouville Eigenvalue Problem 281
Then, we need a2 and a4 to equal 0, and a1 and a3 to equal 1 for Equation (8.2)
to be true; for Equation (8.4) to be true, we need a1 and a3 to equal 0, and a2 and
a4 to equal 1. Thus, Equation (8.5) is really a generalized expression of the BCs in
Equations (8.2 and 8.4). Therefore, we may restate the spatial eigenvalue problem
given by Equation (8.1), subject to the BCs, Equations (8.2 and 8.4), as the general
spatial problem
subject to
ai(0) + a2(p'(O) = 0
c(x)n(x)
at - ax 1Ko(x)1
ax, + q(x)u(x, t), (8.8)
subject to
Ou(O, t)
-K0(0) = -hiu(0,t)
(8.9)
KO(L)OLt) = -h2u(L,t)
282 Chapter 8: The Sturm-Liouville Eigenvalue Problem
and
u(x, 0) = f (x).
Since the PDE and BCs are homogeneous, the separation of variables technique ap-
plies. Letting u(x, t) =cp(x)G(t), taking the appropriate derivatives, and rewriting
Equation (8.8) yields
dx
The spatial equation may be written as
=0. (8.11)
dx
Applying u(x, t) = (x)G(t) to the BCs yields
-Ko(0)p'(O) _
-h2(L)
after separation. The BCs also may be rewritten as
h1 cP(0) - K0(0)'(0) = 0
(8.12)
h2(L) + Ko (L)co'(L) =0.
Thus, we have the spatial equation, Equation (8.11), subject to the BCs, Equation
(8.12).
At this point, we usually solve the spatial problem for eigenvalues and eigen-
functions. However, the spatial problem described by Equations (8.11 and 8.12) is
not our usual spatial problem. Therefore, we must determine if a solution exists,
and if we can find it.
Equations (8.11 and 8.12) may look somewhat messy, but, by making the fol-
lowing substitutions, Equations (8.11 and 8.12) become more compact. First, let
K0(x) = p(x). Second, let c(x)p(x) = a(x). Finally, since h1, K0(0), h2, and
K0 (L) are all constants, we let them equal a1, a2, a3, and a4, respectively. These
substitutions yield the spatial problem
subject to
al(0)+a2'(0) =0
(8.14)
a3(L) + a4cp'(L) =0.
If p(x), q(x), and a(x) are real continuous functions on the interval 0 < x < L;
p(x) and a(x) are always greater than 0 on the interval 0 < x < L; and ai, i = 1
to 4 are real numbers, then Equations (8.13 and 8.14) are known as the regular
Sturm-Liouville eigenvalue problem on the interval 0 < x < L. A more
generalized version of the regular Sturm-Liouville eigenvalue problem is stated on
the interval a < x < b. However, for our purposes we generally use the interval
0<x<L.
The fact that Equations (8.13 and 8.14) are regular Sturm-Liouville eigenvalue
problems doesn't mean we have a solution. It means we know the class that ODE
Equations (8.13 and 8.14) belong to, and there are four very important theorems
which apply to any Regular Sturm-Liouville Eigenvalue Problem. These theorems
are as follows :
Theorem 47. There exists an infinite number of eigenvalues, An, n = 1, 2, 3, .. .
Furthermore, the eigenvalues, An, n = 1, 2, 3, ..., have the following properties
All eigenvalues, An, n = 1, 2, 3, ..., are real numbers.
There is a smallest eigenvalue, usually denoted A1.
There is no largest eigenvalue. This means An - oo as n -f 00.
For the eigenvalues An and Am, we have An Am unless n = m. Therefore,
the eigenvalues may be ordered in the following manner:
Ai <A2 <A <... <An <...
Corresponding to each eigenvalue, An, n = 1, 2, 3, ..., there is an unique
eigenfunction, denoted co (x), n = 1, 2, 3, ..., which has exactly n - 1 zeros
on0<x<L.
We will only prove selected parts of Theorem (47). For a complete set of proofs,
I refer the reader to An Introduction to Partial Differential Equations by Michael
Renardy and Robert C. Rogers. We prove that all eigenvalues are real numbers,
and eigenfunctions corresponding to each eigenvalue are unique, after we state and
prove the next theorem.
Theorem 48. The eigenfunctions, (Pn (x), n = 1, 2, 3, ..., corresponding to differ-
ent eigenvalues, An, n = 1, 2, 3, ..., are orthogonal relative to the weight function
a(x). This means
L
Proof. Let cPn (x) and Pm (x), with corresponding eigenvalues An Am, be solutions
to
subject to
ai'P(0) + a2'P(0) = 0
a3cp(L) + =0.
Thus, we have
ai'Pn(0) + a2pn(0) = 0
a3pn(L) + =0,
and
and
d {p(x)(x)] com(x).
-f L dx dx. (8.23)
L
= f 0
dx
L
+J 0
dx+
which becomes
(An - Am)J L dx
0
= (s.24)
-p(L)Pn(L)Prn(L) +1()On(O)Pm(O),
286 Chapter 8: The Sturm-Liouville Eigenvalue Problem
= p(L)
p(O) (8.25)
(k3cPn(L) _
and
(kl(Prn(O) _ -a2(O)
We can now prove, from Theorem 47, that all eigenvalues are real and that
eigenfunctions corresponding to each eigenvalue are unique.
Proof. Suppose the eigenvalues, An, n = 1, 2, 3, ..., are not real numbers. Hence,
they are complex. Thus, each eigenvalue, An, n = 1, 2, 3, ..., has the form
a + bi. Let A be a complex eigenvalue and cp(x) be the corresponding eigenfunction.
The function, cp(x), must be complex since the differential equation defining the
eigenfunction is complex. Thus, (x), with its corresponding eigenvalue A, is a
solution of
subject to
ai(O) + a2'(0) = 0
(s.27)
a3(L) + =0.
Section 8.2: Definition of the Sturm-Liouville Eigenvalue Problem 287
Please note the functions p(x), q(x), and a(x) are real functions. Since A is a com-
plex eigenvalue with corresponding eigenfunction cp x ,then A (complex conjugate)
is an eigenvalue with corresponding eigenfunction (x). This statement is devel-
oped from a similar theorem about complex roots of quadratic equations. Thus,
(x), with its corresponding eigenvalue A, is a solution of Equations (8.26 and 8.27).
Hence, we have
subject to
al(0)+a2'(0) =0
a3(L) + CY4cP'(L) =0.
(A_)fa(x) (x)2 dx = 0
I2
since (x)co(x) = I cp(x) > 0. Also, a(x) is defined to be greater than 0 for all x on
the interval. Thus,
L
Equation (8.31) may only equal 0 if cp(x) - 0 (identically equal to 0), which we
know can't be true since we may only have n - 1 zeros for 0 < x < L. Hence, we
must have (A - A) = 0. Therefore, A = A, which implies A is real.
Finally, we prove eigenfunctions corresponding to each eigenvalue are unique up
to an arbitrary constant multiplier.
Proof. Suppose cp1(x) and cp2 (x) are eigenfunctions corresponding to the same eigen-
value, A. Then, we have
subject to
(O) + (0) = 0
(8.33)
(L) + a4( L) _ 0,
288 Chapter 8: The Sturm-Liouville Eigenvalue Problem
and
dx
(x)(x)] + q(x)2(x) + o, o < x < L, (8.34)
subject to
al2(0) + a2(p(0) = 0
(8.35)
(8. 35)
a4(L) = 0.
If we multiply Equation (8.32) by cp2(x), then subtract from it Equation (8.34)
multiplied by cpl (x), we have, after some algebraic manipulation,
L
_ Pi (x) (x)] o -f 0
(x) dx+
which becomes
f 0
L (x)
0
(x) dx,
or just
p(x) 0.
Section 8.2: Definition of the Sturm-Liouville Eigenvalue Problem 289
By definition, p(x) is greater than 0 for all x on the interval, 0 < x < L. Therefore,
we may divide both sides of Equation (8.38) by p(x). This yields
which implies
llx = C(p2(x)
Thus, cot (x) is a constant multiplier of co1(x). Therefore, eigenfunctions corre-
sponding to each eigenvalue are unique up to an arbitrary constant multiplier. This
completes the proof.
Theorem 49. Any piecewise smooth function may be represented by a Fourier se-
ries of the eigenfunctions con (x) . This means
f (x) ancPn(x),
n=1
where
L
f(x)p(x)a(x) dx
0
Cln = J fL
cpn(x)o(x) dx
0
J
Furthermore, the infinite series converges to f (x) wherever f (x) is continuous on
f (x f (x ) where f(x) has
- -
0 < x < L, and it convey9 es to ) 2
2
discontinuity.
Sometimes the eigenfunctions, con (x), n = 1, 2, 3, ..., are referred to as forming a
"complete" set.
The proof of Theorem (49) is beyond the scope of this course. A generalized
proof may be found in Introductory Functional Analysis with Applications, Seventh
Edition by Erwin Kreyszig.
290 Chapter 8: The Sturm-Liouville Eigenvalue Problem
+ fL [p(x) ((x))2 - dx
L
f cpn(x)Q(x) dx
called the Rayleigh quotient, relates any eigenvalue, An, to its corresponding
eigenfunction, (pn(x), for n = 1, 2, 3, .. .
Proof of the Rayleigh quotient is provided in the next section.
A more general expression of the above theorems results if we state the theorems
with the bounds a <x <b instead of 0 <x <L.
EXERCISES 8.2
subject to
(1) = 0 and (b) = 0, 1 <b.
(1) Show that multiplying Equation (8.42) by 1 and performing some al-
gebraic manipulation allows you to put Equation (8.42) in the regular
Sturm-Liouville form.
(2) Show that A > 0.
8.2.2. Consider the non-Sturm-Liouville differential equation
P"(x) + A 3(x)co(x) = 0.
Multiply this equation by H (x) . Then, determine H (x) so that the equation
may be reduced to the standard Sturm-Liouville form:
P1 = x
3x2 - 1
P2 =
2
5x3 - 3x
P3 2 ,and
35x4 - 30x2 + 3
P4
8
(1) Show by direct calculation that they are orthogonal to each other on
-1<x<1.
(2) Show by direct calculation that the integral of the square of each of the
Legendre polynomials on -1 <x < 1 is equal to
2
2n + 1'
where n is the subscript of the respective Legendre polynomial.
292 Chapter 8: The Sturm-Liouville Eigenvalue Problem
Lo (x) = 1,
L1(x) = 1-x,
L2 (x) = x2-4x+2, and
L3 (x) = -x3 + 9x2 - 18x + 6.
(1) Show by direct calculation that they are orthogonal to each other on
0 <x < oc with weight function e-x.
(2) Show by direct calculation that the integral of the square of each of the
Laguerre polynomials on 0 <x < Do with weight function e-x is equal
to
(n!)2
Ho(x) = 1,
Hl (x) = 2x,
(1) Show by direct calculation that they are orthogonal to each other on
-oc <x < oc with weight function e-X2.
(2) Show by direct calculation that the integral of the square of each of the
e_x2
Hermite polynomials on -oc <x < oc with weight function is equal
to
2nnk./,
where n is the subscript of the respective Hermite polynomial.
8.2.9. Consider
(x'(x))'+A()(x) 1
subject to
subject to
al(p(0)+a2(,o'(O) =0
(8.44)
a3(L) + a4(p'(L) _ 0,
we show each eigenvalue, An, n = 1, 2, 3, ..., is related to its eigenfunction, c ' (x),
n = 1, 2, 3..... Thus, we show
L
f [((p ,(x))2 - q(x)cp2(x), dx
L
cp2(x)Q(x) dx
Jo
Proof. Let cpn(x) be a solution of Equation (8.43). Then, Equation (8.43) becomes
dx
(x)(x)] (x) + 0,
f0
L dx = f L
0
-d (x)(p(x)] Pn(x) dx. (8.47)
294 Chapter 8: The Sturm-Liouville Eigenvalue Problem
Rearranging the left side of Equation (8.47), and noting An is a constant, yields
f L dx + An
0
L (x)(x) dx = f L -dx (x)(x)] Pn(x) dx. (8.48)
0
Using integration by parts on the right side of Equation (8.47), Equation (8.48)
becomes
An IL
J0L 0
L
_ f
0
p(x) ((p (x))2 dx. (8.49)
+f L [p(x) ((p(x))2 dx
L
f co(x)o(x) dx
Thus, we have shown the Rayleigh quotient relates each eigenvalue An with its
corresponding eigenfunction (pn (x). D
One use of the Rayleigh quotient is to directly prove that An, n = 1, 2, 3, ... is
greater than or equal to 0 if
1) 0 and
subject to
(8.51)
o'(1) = 0.
the point of this example is finding a continuous function satisfying the boundary
conditions, which is not cp1(x). First, we find the lower bound of the eigenvalues.
From Equation (8.50), we immediately determine p(x) = 1, (x) = 1, and q(x) = 0.
This, in turn, means the Rayleigh quotient becomes
1
f ((x)) dx
(8.52)
p(x) dx
L
Using the BCs stated in Equation (8.51), we have cpn (x)cpn (x) Io = 0. Thus, since
(p(x))2 > 0 and SPn (x) 0 for all n, we know Equation (8.52) is greater than
or equal to 0. Therefore, the lower bound for the first eigenvalue is 0. To find an
upper bound, we must find a continuous function that satisfies the BCs, Equation
(8.51). Consider the trial function
ft(x) = x2 - 2x.
or
3 1
43 - 4x2 + 4x
0<A1< - x4 + 4x3 1
0=2.5.
5 3 IO
..Therefore, 0 < Al < 2.5. We note the exact value of Al = ti 2.467 < 2.5.
4
Next, a slightly more complicated example using the more general form of the
Rayleigh quotient is considered.
296 Chapter 8: The Sturm-Liouville Eigenvalue Problem
subject to
(8.54)
cp(2) = 0,
find the lower and upper bound for the first eigenvalue.
f
x
fin 1
x- x<2,
- x('(x))-
>0on1 Pn 2 > 0on1 - x<2foralln
- ,
2
and f(x)
n 1x dx > 0, we must determine ifpn n >- 0. 2
2
-x(x)'(x) I 1
0 < Al <
-x (-2(x2 + 1) + 5x) (-4x + 5) + f x (-4x + 5)2 dx ti 17.05.
2
(_2(x2+1)+5x)2 1 dx
1
x
Therefore, 0 < Al < 17.05. The large gap between 0 and 17.05 may mean our trial
function was not very good. If we had chosen as our trial function x3 - 5x + 2,
then our upper bound would become approximately 11.52, which decreases the gap.
Section 8.3: Rayleigh Quotient 297
That is, 0 < Al < 11.52 < 17.05. A far better approximation occurs with the trial
function
J x, 1<x<1.5
fe(x) _
-3x+6, 1.5<x<2 ,
which, when used in the Rayleigh quotient, yields an upper bound of 0.583734.
Today, one of the major uses of the Rayleigh quotient is a numerical estimate
on the beginning bounds of eigenvalues when solving a variable coefficient PDE
numerically with a computer. We now develop a general solution to a variable
coefficient PDE.
EXERCISES 8.3
8.3.1. Use the Rayleigh quotient to obtain a reasonably accurate upper bound for
the lowest eigenvalue of the following:
(1) cp"(x) - x2cp(x) + Acp(x) = 0 with cp'(0) = 0 and cp(1) = 0.
do(0) _ d2(p(1)
dx - 0, and dx2 = 0.
Show that the eigenvalues are less than or equal to 0. Also, answer the problem
if A = 0?
8.3.5. Determine all negative eigenvalues (if any) for
(p"(x) + 5o(x) + A(p(x) = 0,
subject to
cp(0) = 0 and cp(7r) = 0.
c(x)n(x)
at - ax IKo(x)l
ax] - au(x, t), (8.55)
subject to
u(0, t) = o
(8.56)
u(L, t) = 0
an d
u(x, 0) = f (x). (8.57)
Since Equation (8.55) and BCs, Equation (8.56), are linear and homogeneous,
we know separation of variables is a valid solution method. Thus, letting u(x, t) =
(p(x)G(t) yields the time equation of
G' (t) = -AG(t), (8.58)
Section 8.4: The General PDE Example 299
with spa ti a l BC s
(p(O) = 0
(8.61)
We know from the description of the problem that Ko (x) and c(x) p(x) are positive
functions for the length of the rod. Therefore, Equations (8.60 and 8.61) are of the
regular Sturm-Liouville type where Ko (x) is p(x), -a is q(x), and c(x) p(x) is o(x).
Thus, we are guaranteed the existence of eigenvalues and eigenfunctions, although
we may not be able to identify them. Setting up the Rayleigh quotient, with the
appropriate functions replaced, results in
f(x)c(x)p(x) dx
Applying the BCs, Equation (8.61), tells us An is greater than 0. Thus, we have
a lower bound on our first eigenvalue. Using the solution to the time equation,
Equation (8.59), and what we know about the solution to the spatial equation,
yields the general solution for u(x, t), which is
u(x, t) = (8.62)
n=1
We find the specific solution by applying the initial condition. Thus, we have
u(2,0) = f (x) -
n=1
or simply
.f(x) _
Thus, we have
f
L
f dx = fA()()()()
n=1
_ An L L cpn(x)cp,n,(x)c(x)p(x) dx.
n=1
By orthogonality,
O,n#m
L
Pn(x)m(x)c(x)p(x) dx = fr. (8.64)
L cpn(x)c(x)p(x) dx, n = m.
Jo
` 4n -
f f(x)p(x)c(x)p(x) dx
0
L
L (8.65)
(pn(x)C(x)P(x) dx
Jo
Therefore, the solution to Equations (8.55, 8.56, and 8.57) is Equation (8.62) where
An is given by Equation (8.65).
EXERCISES 8.4
8.4.1. Given
a2u ar aul
(x)ax -Bu,
p(x) at2 - ax T e
subject to
and
au(x, 0)
u(x, 0) = f (x) and at = g(x),
determine the general solution to include the Rayleigh quotient for An, the
orthogonality relationships, and the equations for the coefficients of the gen-
eralized Fourier series. You may call the eigenfunctions cp(x).
Section 8.4: The General PDE Example 301
8.4.2. Consider
subject to
u(0, t) = Tl and u(L, t) = T2
and
u(x,O) = f(x).
(1) Let v(x) be a solution of the problem
[T(x)v']' - /3(x)v = -Q(x),
subject to
v(0) = Tl and v(L) = T2.
If z(x, t) = u(x, t) - v(x), find the boundary value problem satisfied by
z (x, t).
(2) Consider the same PDE, only now subject to the BCs
au(O,t)- andau(L,t)
au(0' t) =T1 + a 2u(L' t) =T2.
ax 1 ax
Find the general solution.
(3) Using the methods of this problem solve
au a2 u
3,
at =axe -
subject to
u(0, t) = 2 and u(7r, t) _ -5,
and
u(x,O) = cosx.
302 Chapter 8: The Sturm-Liouville Eigenvalue Problem
subject to the IC
u(x,0) = f(x) (8.67)
and BCs
8u(0, t)
+ hou(0, t) = 0
where the an are the eigenvalues found when solving the spatial equation.
When solving the spatial equation, Equation (8.70) subject to the BCs, Equation
(8.72), we must remember to consider all three cases for A. That is, A <0, A = 0,
andA>0.
Case 1: A < 0: Let A = -s, where s > 0. Then Equation (8.70) becomes
(8.74)
Cl =
- C2
H0
Applying the second BC of Equation (8.72) to Equation (8.75), we get the tran-
scendental equation
s-H0
'vJ
tanh(,/TT.l -
Letting z = \/L, we can separate the transcendental function into two functions
f1(z) = tanh z
and
z2 - HoL
f2(z) = z (Ho - HL)
304 Chapter 8: The Sturm-Liouville Eigenvalue Problem
flz)
z2 - HoL
Figure 8.1: Gra ph of .fi(z) = tanh z and fadz ) -
z(Ho - HL)
Remember that s > 0; then z > 0. We graph the two functions and look for inter-
section points. Figure (8.1) shows one possible intersection point. The intersection
point indicates the first eigenvalue. It is
z2
(s.77)
C2 X
C1 = - H
0
Applying the second BC of Equation (8.72) to Equation (8.81) gives us the tran-
scendental equation
fl(z) = tanz
and
Z(HL - H0)
f2(z)=
z2 + HLHoL
We graph the two functions and look for intersection points. Figure (8.2) gives a
C2 A
possible representation of the intersection points. Also, we know that Cl =- H0 .
z(HL - Ho)
Figure 8.2: Graph of fl(z) = tanz and f2(z) =
z2+HLHoL
We now have the complete solution to the spatial problem. Therefore, the general
306 Chapter 8: The Sturm-Liouville Eigenvalue Problem
solution to Equation (8.66) subject to the BCs, Equation (8.68), may be given by
= Dle-'t (_cosh(x) + Ho
(s.84)
z2
where Dn = Cndn, D1 = C1d1, and Al = - L . We may find Dn by using the IC,
Equation (8.67), and the orthogonality of the eigenfunctions, and it is
L
(pn(x) f (x) dx
Dn = J° L (8.85)
3ir
it < z3 = X L < 2
Sir
2ir < z4 = a4L <
2
7ir
3ir < z5= VX5 L < 2
Sir
4ir < z6= XL < 2
(2n-3)ir
(n - 2)ir < zn,XL < 2
Section 8.5: Problems involving Homogeneous BCs of the Third Kind 307
Using Table (8.86), we may develop upper and lower bounds for each An starting
with n = 2. In general,
C(n - 2)ir12
L )
<< 1 (2n -
2L
(s.87)
(n - 2)irl 2
Again , using Figure (8.2) , it is interesting to note that , as n -- oo , an --
L J
We finish this section by presenting an example.
EXAMPLE 8.6. Consider heat conduction in a one-dimensional rod with perfect
lateral insulation of length 2ir and no internal source. Suppose the boundary at 2ir is
maintained at the constant temperature of 0°C, and the boundary at 0 is subject to
au(0' t)
Newton's law of cooling where it is known to have the relationship u(0, t) _
ax
Also, the initial condition is known to be
x+1 0<x<ir
J \'J
'-'1
-(it + 1)x
+ 2 (ir + 1) 7r < x < 27r.
r
The mathematical model for this problem is
au(x,t) a2u(x,t)
subject to
au(0, t)
u(0, t) =
8x
u(2ir, t) = 0
and the IC, Equation (8.88). The PDE and BCs are linear and homogeneous.
Therefore, we may apply the separation of variables technique. If we let u(x, t) _
G(t)cp(x), and we take the correct partials and substitute into Equation (8.89), we
have
Separating Equation (8.91) and setting the equations equal to an arbitrary constant
yields the time equation
(8 .93)
308 Chapter 8: The Sturm-Liouville Eigenvalue Problem
Applying the same technique to the BCs produces the BCs for the spatial problem,
which are
(O) =
(8.94)
(2ir) = 0.
Equations (8.93 and 8.94) are of the regular Sturm-Liouville type where p(x) = 1 =
o(x), q(x) = 0, c1 = 1 = a3, c2 = -1, and c4 = 0. Thus, we may use the Rayleigh
quotient to find a lower bound for A. The Rayleigh quotient for this problem is
2ir
f0
dx
f[(x)]2dx
We would like A > 0. Since ['(x)]2 > 0 and [(x)]2 > 0, we need only check
-cp(x)cp'(x). Since cp(2ir) = 0, then
2ir
(x) = (0) (0).
=
[()]2 >0.
Thus, A > 0. Therefore, we need only solve the eigenvalue problem for A > 0. We
know that cp"(x) _ -A(x) has the solution
(x) = cl cos V5x + c2 sin fx. (8.95)
or
ZI Z3
Z2
Z4
Z5
Using Figure (8.3), we may set up a table of the intersection points of the graphs
of the two functions f 1(z) = tan z and f2 (z) = -f--. The intersection points are
27r
the actual points that the two functions are equal. However, we are interested in
bounding the intersection points. Bounding the intersection points will yield upper
and lower bounds of the eigenvalues. Thus, the table of bounding values for each
zn is as follows :
Lower Bound zn = Xn 27r Upper Bound
7r
< zl = X 27r < 7r
2
37r
< z2 = %27r < 27r
2
57r
< z3 = X 27r < 37r
2
310 Chapter 8: The Sturm-Liouville Eigenvalue Problem
7ir
< Z4=\/)27r < 47r
2
Sir (8.99)
< z5 = X527r < 5ir
2
(2n- 1)ir
< zn = /Xri2ir < nir
2
Using Table (8.99), we may develop the table for An, which is really what we want.
Therefore, Table (8.100) contains upper and lower bounds for each An:
1 < Al < 1
16 4
9
1
16 < A2 <
25 9
< As <
16 4 (8.100)
49
A4 < 4
16 <
81 < < 25
A5
16 4
(2n- 1)2 n2
16
< An < 4
It i s i nteres ti ng t o no t e th a t , as n - 00, -- (2 n - 1)
We have an id ea o f
An
2
16
the bounds on the eigenvalues and of the eigenfunctions, n (X). Also, we know the
solution to the time problem Equation (8.92), which is
G(t) =
u(x, t) = (8.101)
n=1
Section 8.5: Problems involving Homogeneous BCs of the Third Kind 311
The complete solution is Equation (8.101), with the equation for an, which is
f2n
f (x)cpn(x) dx
an =
J
2 dx
EXERCISES 8.5
cP"(x) + ) p(x) = 0,
subject to
cp'(0) =0 and cp'(1) + 2cp(1) =0.
(1) Prove that A > 0.
(2) Does A = 0?
(3) Determine all eigenvalues graphically. Obtain lower and upper bounds.
Estimate the large eigenvalues.
8.5.2. Consider the eigenvalue problem
01u 01226
- 9u,
at = 9ax2
u(x, 0) = f(x).
312 Chapter 8: The Sturm-Liouville Eigenvalue Problem
au 0.402 u
= axe
at
au( a0,t)
u(0, t) - = 0 and u(1, t)+ a ax't = o
u(x, 0) = f (x).
Note: You may call the relevant eigenfunctions co (x) and assume that they
are known.
012 u 012 u
01x2
ate
a ao, t)
= u(o, t) and
u( 0, t _ -u(400, t)
au(x, 0) = O.Olx
u x 0 =0 and - 0.000025x2.
au(0, t) t)
u(0, t) = and u(7r, t) -
8x 8x
and
au (x, 0)
u x 0 = x2 2x 1 and =2x+2.
Section 8.5: Problems involving Homogeneous BCs of the Third Kind 313
subject to
u (0 , t) =
8u(0, t)
ax an d u (L , t) _ - 8u(L,
ax
t)
and
au(x, 0
u(x, 0) = 1(x) and
at
(1) Graphically determine the eigenvalues.
(2) Estimate the large eigenvalues.
(3) Determine the solution.
(4) What happens after a long time?
8.5.10. Consider the PDE
au(x, t) _ a2u(x, t) au(x, t)
at axe + ax
Solve the initial value problem subject to the BCs
au (7r,
u( 0't)= 0 and
ax
t) =0
and IC
u(x, 0) _ - sinx.
In five sentences or less, give a physical description of this problem.
8.5.11. Consider the PDE
a2u(x,t) __ a2u(x,t) au(x,t)
at2 5x2 + ax
Solve the initial value problem subject to the BCs
a O,t) =0
and
(31\)2
and =x2- .
u(x, 0) = 1 and
au(x, 0) -x
- 2
.
at
In five sentences or less, give a physical description of this problem.
8.5.13. Consider the PDE
au(x,t) __ a2u(x,t) au(x,t)
9 4
at axe + ax
Solve the initial value problem subject to the BCs
au(0, t) au (2ir, t)
at
= 0 and t)_
ax
=0
and IC
u(x,0) = -cosx.
In five sentences or less, give a physical description of this problem.
Chapter 9
Solution of Linear
Homogeneous
Variable-Coefficient ODE
9.1 INTRODUCTION
In Appendix C, the general second-order linear homogeneous variable-coefficient
ODE is given as follows:
S(x)u"(x) + K(x)u'(x) + H(x)u(x) = 0. (9.1)
Airy's u" - xu = 0,
A few examples where these equations arise in problems are astronomy, quan-
tum physics, electromagnetism, propagation of electromagnetic radiation through a
317
318 Chapter 9: Solution of Linear Homogeneous Variable-Coefficient ODE
we assume that S(x), K(x), and H(x) are polynomials having no common factors.
Then xo is said to be an ordinary point of Equation (9.3) if S(xo) 0. If x0 is
not an ordinary point, then it is a singular point. For instance, every real number,
x0, is an ordinary point of
u"+xu=0.
Whereas, xo = ±1 are singular points of
(1-x2)u" -xu'+n2u= 0.
If x0 is an ordinary point of Equation (9.3), then Equation (9.3) may be rewritten
as
lim(x - x
X-4xo
S is finite
and
lim (x - xo)2
x-xo
5 is finite. (9.7)
Section 9.2: The General second-order ODE 319
H(x) = x.
Hence, for x0 = 0, we have S (xo) = 0;
K (x)
lim - x 0) x) = x fix = > a finite constant;
x-xo S(x) x -O O x2
H(x)
lim x - x0) 2 = lim x2 x = x a finite constant.
x-xo ( S(x) x-0 x2
Therefore, Euler's equation is an example of an equation that has a regular singu-
larity at x0 = 0. Legendre's equation has regular singular points at x0 = ±1, which
is left as an exercise.
This text examines those variable coefficient ODEs that have only ordinary
points or regular singular points, and are relevant to PDEs. ODEs that have ir-
regular singular points are extremely difficult to solve, and their study is left to an
advanced course in ODEs.
EXERCISES 9.2
9.2.1. Show that the following equations have ordinary or regular singular points:
(1) u"-xu=0.
(2) x2u" + xu' + (x2 - n2) u = 0.
(3) (1 - x2) u" - xu' + n2u = 0.
(4) u" - 2xu' + gnu = 0.
(5) (1 -x2)u" - 2xu' +n(n+ 1)u = 0.
9.2.2. Determine if the following equations have an ordinary point, a regular singular
point, or an irregular singular point at 0 and 1:
(1) xu"+(1-x)u'+xu=0.
(2) x2 (1 - x2) u" + 5xu' + 8u = 0.
(3) x5(1-x2)u"-3xu'+u=0.
(4) 2x4 (1 - x2) u" - 2 u' + 3x2u = 0.
(5) u" + 1+x u' + 4 (1 + x) u = 0.
(6) (x + 2) u" + 2xu' - (1 - x2) u = 0.
320 Chapter 9: Solution of Linear Homogeneous Variable-Coefficient ODE
You can determine the solutions by power series methods. However, sometimes the
solution may be determined by a proper guess, which is the way most second order
constant-coefficient linear ODEs were originally solved.
For our guess, we'll let u(x) = xr, where x 0, since 0 is a singular point of
Equation (9.8). Then, u'(x) = rxr-1 and u"(x) = r(r - 1)xr-2. Substituting u(x),
u' (x), and u" (x) into Equation (9.8) yields
-
x2r(r 1)xr-2 + flxrxr-1 + xxr =
0,
r= -(-1)±-1)2-4x --ryf8,
2
where
2 and
(a-1)2-4x
2
Therefore, the roots of the characteristic equation may be expressed as
rl = -(-1)+(-1)2-4
2
x =-7+8 (9.10)
and
r2 =
-(-1)--1)2-4 (9.11)
2
Case 1: (/3 - 1)2 - 4x > 0, produces two real and unequal roots.
Case 2: (/3 - 1) 2 - 4x = 0, produces real and equal roots.
Case 3: (/3 - 1) 2 - 4x <0, produces complex roots.
The following examples illustrate the type of solution for each case.
EXAMPLE 9.1. Consider
x2"xu'
u - -6u=0.
If we assume a solution u(x) = xr, where x > 0, then the characteristic equation
becomes
r(r - 1) - r - 3 = 0,
which is
r2-2r-3=0.
Thus, we have Case 1: two real and unequal roots, where the roots are -1 and
3. By our assumption, u(x) = xr, we have the two solutions u1(x) = x-14 and
u2 (x) = x3. Therefore, we know that the general solution is
u(x) = C1x-1
+ C2x3. U
equation becomes
r(r - 1)+5r+4=r2 +4r+4 = (r+2)2 = 0.
Thus, we have real and equal roots r=-2 with one solution u1 (x) = x-2. The second
solution may be found by using reduction of order, and it is u2 (x) = x-2 ln(x), which
means the general solution is
u(x) =x2 (c1 + c2 ln(x)) .
By far, the most interesting case is Case 3. To arrive at the proper solution for
Case 3, we must remember some basic mathematics. Since you already know that
the solution has complex roots, you probably expect sine and cosine functions to
appear. The question is how? Remember that
xr = er In X
From this simple formula, we arrive at the mathematically cleanest solution. We
now proceed with the example.
322 Chapter 9: Solution of Linear Homogeneous Variable-Coefficient ODE
rl =
2
r2 =
-1 2- 2
Thus,
u1(x) = xr1
-1+t
= x 2
-l lnx if lnx
= e 2 e 2
= x 2 e iylnx
-1
2
Similarly,
-1 -2ylnx
u2 (x) = x 2 e 2
Thus,
-1 iyln x -iyln x
x2 cl e 2 + c2 e 2
Note: In Examples (9.1, 9.2, and 9.3) we assumed x > 0. For the complete
solution given in Proposition (52) for all x 0, we must use absolute values.
x2u" + /3xu' + xu = 0.
We can solve Equation (9.12) in any interval such that x 0 by substituting u(x) =
xr and its derivatives into Equation (9.12), solving for the roots rl and r2 of the
resulting characteristic equation
r2 + ( - 1)r + X = 0.
If the roots of the characteristic equation are real and unequal, then
EXERCISES 9.3
9.3.2. Consider the interval x > 0. Let x = ey. Knowing that Euler's equation is
show that
du 1 du
(9.16)
dx x dy
and
d2 u _ 1 d2 u l du
(9.17)
dx2 x2 dy2 x2 dy
Next, using Equations (9.16 and 9.17), show that Euler's equation becomes
d2 u du
d2+(/3-1)d +u=O. (9.18)
y y
9.3.3. In Example (9.2) it was said that a second solution to the real and equal roots
case may be found by reduction of order. Given Euler's equation
and the fact that rl = r2 are real roots of the characteristic equation
ar(r-1)+br+c=0.
Thus, ul (x) = xnl is one solution of Euler's equation. Use the method of
reduction of order to show that u2 (x) = xnl In x is the second solution.
Section 9.4: Brief Review of Power Series 325
ex
00 n2n+1
sin x
(2n+1)!
00
/-l\nx2n
l
cos x
n_O (2n)1
00
1
1-x
-1<x< 1, and
n=0
-1)n-}-lx2n
n
-1<x<1.
n=0
You should have encountered these series in your study of Calculus. However, you
may not have encountered a general definition of a power series. It is as follows:
Definition 53. An infinite series of the form
00
where xo and b0, b1, ... , bn, ... are constants, is called a power series in x - x0.
Knowing the general definition of a power series only helps us if we know when
and how a power series converges. The power series given in Equation (9.19) is said
to converge if
N
lim bn(x - x0)n
N-oo n=0
exists; otherwise it is said to diverge. If x = x0, the power series must converge.
For values of x x0, we may not know whether the series converges. However, if
the power series converges, the following theorem applies:
Theorem 54. If the power series
>bn(x_xo)
n=0
326 Chapter 9: Solution of Linear Homogeneous Variable-Coefficient ODE
(2) the power series converges absolutely and uniformly for all values of x.
(3) there exists a positive number R (called the radius of convergence) such that
the power series is absolutely convergent for all Ix - xo I < R, and divergent for
x - xo > R.
The proof of this theorem may be found in the text Infinite Series by Earl D.
Rainville. Also, note that in the radius of convergence, the power series is often
written as
Since we are developing a power series solution for a general second-order ODE,
it would seem useful to know if a power series may be differentiated. This is given
in the following theorem. The proof may also be found in the text Infinite Series
by Earl D. Rainville.
Theorem 55. If the power series
00
then
00
bn (x _ xo )n-m
n=no
may be rewritten as
00
bn+m (x - xo) n .
n=no-m
Section 9.5: The Power Series Solution Method 327
In the next section, we show how power series may be used to solve Equation
(9.1) at an ordinary point.
EXERCISES 9.4
(3)
(9)
(5)
u'(x) _ (n + 1)bn+lxn.
n=0
Replace the assumed solution and the derivative in the original equation, and we
have
or
b = 2bn (9.23)
Equation (9.23) is called a recurrence relation. This means that all coefficients may
be expressed in terms of b0. For example,
b1 = 2b0,
b2 = b1=2b0,
b3 = 2 b2 = 2 (2bo ),
3 3
b4 = Zbs = 3Zbo)>
Section 9.5: The Power Series Solution Method 329
3 3 '
since
u"-xu=0 (9.24)
Then
Substituting Equations (9.25 and 9.26) for u and u", respectively, in Equation (9.24)
yields
Here, we cannot simply combine the sums. To express Airy's equation in powers of
x -2, we must express x in front of the second sum in powers of x -2. This is done
20. B. Airy (1801-1892) was the Astronomer Royal of England. He made many contributions
to the study of series and integration.
330 Chapter 9: Solution of Linear Homogeneous Variable-Coefficient ODE
00 00
n=0
(n + 2) (n + 1)bn+2 (x - 2)n = 2
n=0
bn (x 2)n +
n=1
bn-1(x - 2)n.
Equating like powers of x - 2 yields
(n+2)(n+ 1)bn+2 = 2bn+bn-1,
which implies
b2 = bo,
Thus, we have
b2 = bo,
b3
= b1 + bo
3 6'
b2 bl b1
b4 = + = b° +
6 12 6 12'
b3 b2 b1
b5 = + = + b°
10 20 30 15'
Thus,
Section 9.5: The Power Series Solution Method 331
/ (x- 3
(x- 4
(x- 5 l
+ bl I (x - 2) + 302) + ...
3 2) + 122) + J
= boul(x) + biu2(x),
where
=/ (x 3 4 5
ul (x) I 1 + (x - 2)2 +
6 2) + (x 6 2) + (x 152) + ... )
and
(x - 2)3 (x - 2)4 (x - 2)5
u2(x) _ C(x - 2) + + ... f
3 + 12 + 30
and bo and b1 are arbitrary. Figure (9.1) shows the graph of Airy's function from
-10 to 5. Notice the oscillatory nature for x <0, whereas for x > 0 the function is
unbounded.
Example (9.5) demonstrates the following theorem:
Theorem 56. If xo is an ordinary point of
u"(x) + ic(x)u'(x) + e(x)u(x) = 0, (9.28)
K(x) H(x)
then icx
( ) =
Sx and e(x) = Sx have Taylor series expansions at xo> then> the
general solution of Equation (9.28) is given by
where c1 and c2 are arbitrary, and u1(x) and u2 (x) are linearly independent se-
ries solutions, which have Taylor series expansions at x0. Also, the Taylor Series
expansions of u1(x) and u2 (x) have radius of convergence at least as large as the
minimum of the Taylor series expansions of ic(x) and e(x).3
Unlike Airy's equation, which may be expanded in powers of any real number,
Legendre's equation, given by
has singularities at ±1. In the next section, we look at Legendre's Equation and its
solution.
EXERCISES 9.5
9.5.1. Find a solution to the following ODEs using the Power series method at
x0 = 0:
(1) u' = 5u.
(2) u"-2u'-3u=0.
(3) u'-2u=0.
(4) u"+4u'+4u=0.
(5) u' + 6xu = 0.
(6) u"+4u=0.
(7) u'-3u+2=0.
(8) u"+u'+u=0.
9.5.2. Find the power series solution for the following ODEs at the indicated value
of x0. That is, find the series solution in powers of x - x0.
(1) (1 + x2) u" + 3xu' + u = 0, xo = 1.
(2) (1 + x2)u" + 6xu' + 2u = 0, x0 = -1.
(3) (1 + 2x2)u" + 3x2u' - u = 0, x0 = 1.
2
u" - 2xu' + nu = 0,
like Airy's equation, in which every real number xo is an ordinary point. Find
a series solution to Hermite's equation. The solutions are called Hermite
polynomials.
2dne_x2
Hn(x) _ (_1)nex
dxn
Generate the first five Hermite polynomials using Rodrigues' formula.
(1) Use the recurrence formula Hn+1(x) = 2xHn (x) - 2nHn_ 1(x) to find the
next five Hermite polynomials.
(2) Show that the Hermite polynomials H2 (x) through H6 found in Part (1)
satisfy the recurrence formula Hn (x) = 2nHn _ 1(x) .
9.5.9. This problem shows how a piecewise smooth function may be expanded in a
generalized Fourier series in Hermite polynomials. The weight function for
the Hermite polynomials is e-x2. Also,
00
e-x2Hn (x) dx = 2m!/.
Let
Show that
1 e_x2
An = f (x)Hn (x) dx.
2nn!
9.5.10. Let
Find the generalized Fourier series in Hermite polynomials. Using your fa-
vorite mathematical software graph the function on the interval (-10,10),
and the generalized Fourier series in Hermite polynomials for n = 10, 25, 50.
334 Chapter 9: Solution of Linear Homogeneous Variable-Coefficient ODE
9.5.11. Let
Find the generalized Fourier series in Hermite polynomials. Using your fa-
vorite mathematical software graph the function on the interval (-5, 5), and
the generalized Fourier series in Hermite polynomials for n = 10, 25, 50.
9.5.12. Find a power series solution to Weber's equation
x2
u"+ n+ 21 - 4 u=0,
where n = 0, 1, 2, .. .
Since S (x) = (1 - x2) is nonzero at 0 we may use the methods we have already
developed in the preceeding sections. Here, the radius of convergence is the open
interval (-1, 1). Thus, we assume that the solution has the form
= (m+1)bm+ixm
m=1 m=0
and
00 00
u"(x) = m(m -
1)bmxm-2
Substituting into Equation (9.30), Equation (9.31), and the correct derivatives of
Equation (9.31) yields
00 00
00
+ n(n + 1) bmxm ,
m=0
which becomes
00 00
- 2b2x2 + .
(9.33)
- 2blx - (2)(2)b2x2 +
n(n + 1)
ba = - 2 bo (9.34)
(3)(2)b3+[ri(n+1)-2]bi =0
(n-1)(n+2)
bi (9.35)
b3 (3)(2)
b3
(n+2)(n-
(3)(2)
(n+3)(n-2) (n+3)(n+1)n(n-2)
(9 37)
.
b o,
b4 (4)(3) b2 - (4)(3)(2)
(5)(4) b3 (5)(4)(3)(2)
C (n+2)(n-1) 3 (n+4)(n+2)(n-1)(n--3)
+ bl x - f ...
(3)(2) x + (5)(4)(3)(2)
where
and
u1(x) = 1 - (n+1)n x2 I
2.
and
K 2k!((2n-2k)!
P(x) _ (-l)k
- k)!( -
2k)!xn-2k
k=o
n n
where K = or 1, whichever is an integer. Thus, a simplified solution to
2 2
Equation (9.30) is
u(x) = c1Pn(x).
P0(X)
Figure 9.2: The plot of the first five Legendre polynomials, Pn (x).
Po = 1,
P1 = x,
3x2 -
P2
2
5x3 - 3x
P3 2 ,and
35x4 - 30x2 + 3
P4
8
338 Chapter 9: Solution of Linear Homogeneous Variable-Coefficient ODE
Qo
l In
2
l+x
1-x '
-1+Zln(i+xl
Q1
l
Figure 9.3: The plot of the first five Legendre functions, Qn(x).
and
=::
1
Pn (x)tn . (9.40)
1 - 2xt + t2 n=0
dx =0, if m n
-1
and
1 2
[P(x)]2 dx =
_1 2n± 1
These are left for you to show in the exercises. This short introduction to Legendre
polynomials completes our coverage of this topic. If you are interested in learning
more, I suggest you study the text Special Functions by Earl D. Rainville. In the
next section, we examine Bessel's equation.
EXERCISES 9.6
9.6.1. Generate the first five Legendre polynomials using the formula
n
Pn(x)=
l -p--(
2
x2 x2 - l n.
dxn
Hint: Use the generating function and differentiate with respect to t. Then
multiply both sides of the equation by 1 - 2xt + t2.
4Olinde Rodrigues (1794-1851), a bench economist and mathematician.
340 Chapter 9: Solution of Linear Homogeneous Variable-Coefficient ODE
9.6.8. Let
-1-x, -1 <x <0
x-1, 0<x<1.
Expand g(x) in a generalized Fourier series in Legendre polynomials. Use
your favorite mathematical software to graph the function, g(x), and the gen-
eralized Fourier series in Legendre polynomials for n = 10, 25, 50, and 100.
9.6.9. Let
h(x) = sin x, - 1 <x < 1.
Expand h(x) in a generalized Fourier series in Legendre polynomials. Use
your favorite mathematical software to graph the function, h(x), and the
generalized Fourier series in Legendre polynomials for n = 10, 25, 50, and
100.
9.6.10. Let
f(x)=cosx, -1<x<1.
Expand 1(x) in a generalized Fourier series in Legendre polynomials. Use
your favorite mathematical software to graph the function, 1(x), and the
generalized Fourier series in Legendre polynomials for n = 10, 25, 50, and
100.
Section 9.7: Method of Frobenius and Bessel's equation 341
9.6.11. Let
IXIr>bflXn, bo 0. (9.41)
n=0
Under certain conditions, which we state later, Equation (9.41) always yields one
solution. The second solution may or may not exist. The problems we study
generally have two solutions. The solutions are either immediately available by
knowing that two roots exist or a second solution is "easily" (Remember: Everything
is relative) constructed. Since it is easier for the presentation, and it does not effect
the method for determining r and all bn, we only consider x > 0. Thus, Equation
5G. Frobenius (1849-1917), a German mathematician who also made important contributions
in group theory.
342 Chapter 9: Solution of Linear Homogeneous Variable-Coefficient ODE
bnxn+r, bo 0. (9.42)
n=0
u(x) = bnxn+r
(9.44)
n=0
and find the appropriate derivatives of Equation (9.44), then replace them in Equa-
tion (9.43). We have
u'(x) = (n + r)bnxn+r-1
(9.45)
n=0
and
00
1)bnxn+r-2
Note: Equations (9.45 and 9.46) seem to not follow the rule for differentiation of
a power series. However, when n = 0 in Equations (9.44) we have boar and the
derivatives of this first term are rboxr-l and r(r - 1)boxr-2
Section 9.7: Method of Frobenius and Bessel's equation 343
Replacing in u(x), u' (x), and u" (x) in Equation (9.43) with Equations (9.44, 9.45,
and 9.46) yields
00
-(1 _ x) bnxn+T ,
n=0
which becomes
00 00
3(n r)bnxn+T
0= 2(n + r)(n + r -
1)bnxn+T
+ +
n=0 n=0
00 00 00
Canceling out the common x'' term in Equation (9.47) and combining like sums
yields
0= [2r(r - 1) + 3r - 1] bo
+ >1(2n + 2r - 1)bn_1xn.
[2r(r - 1) + 3r - 1] = 0,
called the indicial equation. It tells us that r = 1 -1. The second and third lines
2'
of Equation (9.49) yield the relationship of the bn's. We have
Thus,
ul(x)
n=0
u(x) = Clx2
f- 2+ 35x2 - 385x3 + .
1 1
+CZx-1 1-x+2x2- 6x3+...
and
00
u2( x) I 1 /
n=1
(2) If r1 = r2 = r, we have
00
and
u(x) = bnxn+r
n=0
and substituting u(x) and first and second derivatives of u(x) into Equation (9.51)
yields
(m + r) (m + r -
m=0
00 00
which becomes, after some algebraic manipulation and canceling the common x''
term,
00
0=
[(m+r+n)(m+r-n)]bmxm
m=0
+ bmxm+2. (9.52)
m=0
0= (r+n)(r-n)bo+(l+r+n)(l+r-n)blx
+(2+r+n)(2+r-n)b2x2+
+box2+
Thus, the indicial equation, which comes from the coefficient of the x0 term, is
which means r1 = n> 0 and r2 = -n. Hence Ti - r2 = 2n > 0. The next equation
is
(1+r+n)(1+r-n)b1 =0,
which must equal 0 for all choices of nonnegative integers n and values of r. There-
fore, b1 must equal 0. The recurrence relation is given by the equation
(2+r+n)(2+r-n)b2+bo =0,
which can be rewritten as
bm
m 1 b m_2 (9.54)
(m+r+n)(m+r-n)
Since b1 = 0 and n is a nonnegative integer, the recurrence relation indicates b3 =
b5 = = 0. Thus, m is always even, so we may replace it with 2m in Equation
(9.54), which becomes
(9.55)
(2m + r + n)-1(2m + r - n) b2m 2
b 2m _
b 2m
2m 1
(2m + 2n)2m 2m 2 b_= 22m(m + n) b_ m=1,2,3,...
2m 2,
1
Section 9.7: Method of Frobenius and Bessel's equation 347
Thus,
bo
b2
22(1+n)
-b2 bo
b4
22(2)(2 + n) 24(2)(2 + n)(1 + n)'
-b4 -bo
b6
22(3)(3 + n) 26(3)(2)(3 + n)(2 + n)(1 + n)'
(- 1)mb
22mm!(1 + n)(2 + n)(3 + n) (m + n)
Therefore, if we let n = 0, we have
r 00 (_1)mX2ml
ui(x)=bo 1+ 22m(m!)2j'
L m=1
which is the Bessel function of the first kind of order zero, and is denoted as Jo (x) .
Since n = 0, we have case (2) of Theorem 57, a double root. Hence, the second
solution has the form
-0.4
Figure (9.4) shows the graph of Bessel's function of the first kind for n = 0, 1, 2, 3.
Bm can be determined by taking first and second derivatives of u2 (x) and replacing
u2 (x), u2 (x), and u2 (x) in Equation (9.51). The solution obtained, after some
algebraic manipulation, is
Y0(x) Y1(x)
Y2(x)
(ry + In 2) Jo(x) In x +
°O
(1)1G
2(m!)2"''
m=1
where
Gm = l+ 1 +...+l+l
m m-1 2
and 'y = 0.577215664..., which is called Euler's constant. Yo (x) is called Bessel's
function of the second kind of order zero. Therefore, the general solution to Equa-
tion (9.51) when n = 0 is
Figure (9.5) shows the graph of Bessel's function of the second kind for n =
0,1,2,3.
EXERCISES 9.7
9.7.1. Use the method of Frobenius to find the indicial equation for each of the
following ODES:
(4) x2u" + xu = 0.
(5) x(u" - 2u - lu) = 0.
(6) x2u" - x2u' + 3xu = 0.
9.7.2. Another set of polynomials is Laguerre's polynomials. They can be found by
using the method of Frobenius on Laguerre's equation
x2u" + xu' + x2 - 14 u = 0.
f 1 x [Jo(Ax)]2 dx =
J0
t [Jo(t)]2 dt = a {[Jo(A)]2 + [JOF(A)]2}.
Hint: Consider the ODE y"+ t, +y = 0, where y =Jo(t). Multiply the ODE
by 2t2y'.
9.7.5. The Hypergeometric equation, sometimes called Gauss hypergeometric equa-
tion,
x(1-x)u"+[c-(a+b+1)x]u'-abu=0,
is solved by the method of Frobenius.
(1) Show that the hypergeometric equation has a regular singular point at
x=0.
(2) Show that the hypergeometric equation has a regular singular point at
2
(10.1)
2 15(5u)
u - r ar r ar
a2u 1
59-+--=0;
52u
(10.2)
+ r2 2
O2 u=
1 a r2 - 1 t9 / sin 8 au +
1
52 u
r2 ar ar + r2 sin 9 ae ae r2 sin2 8 acp2
351
352 Chapter 10: Classical Problems of PDEs
where a is the radius of the circular plate. Figure (10.3) shows the cylindrical plate.
It appears we only have one boundary condition. However, Equation (10.4) is
in polar coordinates, which allows us to make some assumptions.
Section 10.2: Laplace's equation 353
2
u=
15(5u)
r
152u
=0,
r ar Sr + r2 582
subject to
u(r, -7C) = u(r, 7C)
and IC
Also, we must note that when r = 0, Equation (10.6) would become unbounded.
Therefore, it is usual to assume that as r -* 0, we have u (r, 8) < oc, which means
that Equation (10.6) is bounded as the radius approaches 0.
The single boundary condition, which is nonhomogeneous, immediately tells us
that the "time-like" condition is in terms of r. Therefore, we proceed as before,
354 Chapter 10: Classical Problems of PDEs
1d dR(r) 1 420(8)
O(e) R(r) = 0, (10.9)
r dr r dr + r2 d92
r d (rd1T)) = AR (r)'
dr dr
which may be written as
r
2 d2R(r)
dr2
+r dR(r)
dr
-AR(r)=0,
subject to the condition
After separating the BCs, Equation (10.7), the homogeneous equation, Equation
(10.13), is subject to the BCs
(10.14)
= O'(ir).
Equation (10.13), subject to BCs, Equation (10.14), is well known to us and has
the solution
Ao = 0, e0(9) = ao
An = n 2 (10.15)
,n=1,2,3,...
On (8) = an cos n9 + bn sin n9
x2y"+/3xy'+xy=0,
Section 10.2: Laplace's equation 355
Letting R(r) = r8, and finding the appropriate derivatives yields the characteristic
equation
[s(s_l)+s_n2]r8 =0
or s = ±n. Thus, the solution is
R(r) = clrn + c2r-n, n = 1, 2, 3,... (10.16)
[s(s-1)-i-s]r8 =0,
which yields s = 0 or a solution of R(r) = c3. However, we know that a second-order
ODE has two solutions. Let us rewrite Equation (10.11) as
d
rd r dR(r) = AR(r) = 0.
r dr
Thus, we have
d (rd')) 0
dr dr '
Thus, our complete solution for Equation (10.11) for the case Ao = 0 is
R(r) = c3 + c4 In r. (10.18)
Applying the condition given by Equation (10.12), R(r) <00 as r -* 0 implies that
c4 = 0. Therefore, the solution for Equation (10.11) for the case Ao = 0 becomes
For the general solution to Equation (10.6), subject to the BCs, Equation (10.7),
we have u(r, 8) = O(8)R(r) = aoc3 = Ao given by Equations (10.15 and 10.19) for
A = 0, and u(r, 8) = O(8)R(r) = (aa cos n8 + ba sin n8) (clrn) = Aarn cos n8 +
Barn sin n8 given by Equations (10.15 and 10.17). Summing all of our solutions,
we arrive at
A0 = ff(9) d8 (10.21)
1
Aa ff(9)cosne dB, (10.22)
?ran
and
ir
1
Ba f (B)sin nB dB. (10.23)
7rctia J-
We now solve Laplace's equation in spherical coordinates.
2 1a 2 au 1 a au 1 52n
u r2 Sr r ar + 2i959 sin 8
a8
+
r2i29
sin 2
= 0. (10.24)
Section 10.2: Laplace's equation 357
However, we will consider the problem where u(r, 8, cp) = u(r, 8). This means we
are considering Laplace's equation in spherical coordinates, which is independent of
cp. This is not an unrealistic assumption, since many problems in electrostatics are
solved in this manner. It is done by setting up the problem in such a way that it is
independent of the variable cp. Therefore, Equation (10.24) becomes
Vu_ 15/25u\
r 1
a
si n 8- = 0. (10.25)
r2 ar ar + r2 sin 8 88 a8
Like the polar coordinate system, as r -f 0, u(r, 8) < oo. Again, this means
that the function u is bounded as r approaches 0. Also, we assume the boundary
of the sphere is now a function of 8. Thus,
u(a, 8) = f(9), (10.26)
where a is the radius of the sphere. This is the "time-like" condition, which will be
used to find the coefficients of the general Fourier series. Since Equation (10.25) is
homogeneous, we let u(r, 8) = R(r)O(e) and proceed with separation of variables.
Thus, Equation (10.25) becomes
R
1r r2 R' ' O + r sin 8
sin 80' ' = 0
which when separated becomes
1 r2R''=- 1
sin 8
Therefore, we have two ODEs,
(r2R)' _ -AR (10.27)
and
(sin 80')' = -a0 sin 8. (10.28)
358 Chapter 10: Classical Problems of PDEs
We will work with Equation (10.27) first. After completing the differentiation,
it becomes
Equation (10.29) we have seen before; it is Euler's equation. (See Chapter 9.) The
usual solution method is to assume the solution has the form R = r'3, and then,
take the appropriate derivatives and determine the characteristic equation. Using
our assumed solution, we have
p2+p+A=0,
which has solutions
-lf 1-4a
p 2
-1 /1
pl- 2 +U4 A (l0.30)
and
A. (10.31)
C2
=C1rn+ .
=C1rn+C2r-n-1
R(r) rn+ 1 (10.33)
Therefore,
d0 .
sin 80 = sin 8 = - sine 9
d9 dw '
applying a trigonometric identity followed by substitution yields
- sine ed = cost 8 - 1
d _ w2 - 1 d0
Thus,
d d0 d r d0 dw
(w -1)- =;
2 2
(sinee')' = I (w -1)
dB dw dB
d de
[(w2 - 1) sin B.
dw J
-
Hence, Equation (10.35) becomes
dw (
1-wed
) dw
+n(n+1)0=0. (10.36)
Equation (10.37) is Legendre's equation. (See Chapter 9.) We know it has solutions
Since Legendre functions of the second kind are not finite at ± 1, cos 0 = 1, and
cos it = -1, we have Qn (cos 9) undefined at 0 and it. Therefore, we must have
Bn = 0. Thus, Equation (10.38) becomes
Combining Equations (10.34 and 10.39) we have our solution for u(r, 9), which
is
00
u(r, 9) = R(r)O(9) = cnrnPn(cos 9). (10.40)
n=0
Thus, the coefficients are found by using orthogonality, and the equation for them
is
ir
cn =
2n+1 f (9)Pn (cos 9) sin 9 d9.
2(x '2 J0
Note:
2
Pn (cos 9) sin 9 d9 =
2n+1
EXERCISES 10.2
and
(< a, a a constant in SZ
has at most one solution.
10.2.4. Let 1 be a bounded region. Show that the Neumann problem
au
V2u+ au = f in SZ and = 9on 5SZ
an
has at most one solution if a <0 in SZ.
10.2.5. Let 1 be a bounded region. Show for the Neumann problem
V2
u =fin 1 and au =9on 5SZ .
an
Show that any two solutions differ by a constant.
10.2.6. If we write Laplace's equation in cylindrical coordinates, then
assume that it is axially symmetric (no dependence on 0), we arrive at
a2u(r, z) 1 a2u(r, z)
= 0.
are + r az2
Use separation of variables where u (r, z) = R(r)Z(z) and show that R and Z
satisfy
rR"(r) + R'(r) + ArR(r) = 0 and Z"(z) - AZ(z) = 0.
362 Chapter 10: Classical Problems of PDEs
8u (r, )
(2) u ( r, 0) = 0, ae 4 = 0 , u ( a, B) = s i n 8 .
Du (r, 0) au (r, 4
(3) = 0, = 0, u(a, 0) = cos 0.
a0 a0
10.2.14. Suppose u (x, y, z) satisfies Laplace's equation. Show that the value of u (x, y, z)
at any point (x, y, z) is approximately equal to the average of its values at the
six surrounding points (x ±6, y, z), (x, y ±6, z), and (x, y, z ± b). Hint: Calcu-
late the Taylor series expansion of u(x + b, y, z) to the term 63, similarly for
the other five points.
10.2.15. In the text, we solved the first boundary-value problem for Laplace's equation
inside a sphere of radius a, which is independent of cp. Solve the first boundary-
value problem for Laplace's equation outside the sphere. Note: The first
boundary-value problem is a Dirichlet BC.
10.2.16. Complete the following:
(1) Solve Laplace's equation for the general second boundary-value problem
inside a sphere of radius a, which is independent of the variable cp. Note:
The second boundary-value problem is a Neumann BC.
(2) Solve Laplace's equation for the second boundary-value problem inside
a sphere of radius a, which is independent of the variable cp, where the
BC is
au
= a cos 0.
an
364 Chapter 10: Classical Problems of PDEs
10.2.17. Solve Laplace's equation for the second boundary-value problem outside a
sphere of radius a, which is independent of the variable cp.
10.2.18. Solve Laplace's equation for the general Sturm-Liouville BCs inside a sphere
of radius a, which is independent of the variable cp.
10.2.19. Find the potential in the interior and exterior of a sphere of radius a if the
upper half of the sphere is charged to a potential of qi and the lower half of
the sphere is charged to a potential 2.
10.2.20. Find the potential in the interior and exterior of a sphere of radius 2 when
one-half of the surface of the sphere is charged to potential o = 0 and the
other half of the sphere has potential of 0.
u(x, t)
We derived the wave equation for a tightly stretched perfectly flexible horizontal
string in Chapter 3. Those equations also allowed us to model longitudinal vibra-
tions of a uniform flexible rod. (See Problem 4.4.9.) However, transverse vibrations
of a thin beam require a new set of equations since, unlike transverse vibrations of a
tightly stretched perfectly flexible horizontal string, the thin beam offers resistance
to bending.
Consider an elastic beam with uniform rectangular cross section, which lies with
the x-axis centered down the beam of length L, as shown in Figure (10.5).
Section 10.3: Transverse Vibrations of a Thin Beam 365
Given a variable load, w(x, t), is on top of the beam, which produces a small
downward deflection of the beam, we have for any small section (x, x + Ox) of the
beam a bending moment, M (x, t). The bending moment is usually given as
M(x, t) = -E(x)I(x)C, (10.41)
where E(x)I(x) is the flexural rigidity and is composed of the Young's modulus
of elasticity, E(x), which depends on the material and the moment of inertia,
1(x). Also, the curvature, C, of the beam under a variable load w(x, t) per unit
length x can be found from calculus and is given as
D2u(x, t)
C= axe 3 . (10.42)
2 2
1 (Du(xt)
+ Dx
When we assume a small slope in the curvature, which is the usual assumption, we
have
Du(x, t) 0
ti
Dx
Newton's law of motion, applied to a small section of the beam, Ox, is given as
the sum of the forces in the u direction equal to mass of the beam times acceleration
of the beam in the u direction,
2
(x, t)
F =may = ma t2't) where m =
9
and the sum of the moments of bending equal to the moments of inertia times the
angular acceleration,
M(x,t) = 1w.
We assume the angular acceleration, w, to be 0. We make this assumption be-
cause we ignore angular acceleration just as we ignored horizontal vibrations in the
derivation of the wave equation for a vibrating string. (See Chapter 3.) Thus,
M(x,t) = 0.
366 Chapter 10: Classical Problems of PDEs
M MX+&
SX+&
We also have shear forces acting on the cross section of the small section of the
beam, Ox, shown in Figure (10.6). We denote shear forces as S(x, t). The sum of
the forces in the u-direction under a continuous load z7(x, t) on the small section of
beam, Ox is
DS(x, t)
F= S x t- (S(x t
Dx
tv x t Ox = 0. (10.44)
If we take the sum of the moments counterclockwise about the point x, we arrive
at
Replacing M (x, t) in Equation (10.47) with its equivalent value found in Equation
(10.43) yields
Taking the derivative of Equation (10.48) with respect to x and adding uv (x, t) to
both sides gives us
a2 _ a2u(x, t) DS(x, t) a2u(x, t)
1 x2
E(x)I(x)
1x2
)+zz7(xt)= ax + z7(x, t) = m 92
which becomes
82u(x, t) 82 D2u(x,
CE(x)r(x) -tz7(x't). (10.49)
m(x) 8t2 + 8x2
If the beam is uniform, then E(x) and 1(x) are constant. Also, if the load uv(x, t)
is constant and does not vary with time, t, Equation (10.49) becomes
D2u(x, t) D4u(x, t)
+ c2 ax4 9, (10.50)
ate
where c2 = EI. Since the gravity, 9, is small compared with the internal forces
m
of the beam, it may be neglected in most applications. Thus, Equation (10.50)
becomes
D2u(x, t) D4u(x, t)
+2
which is the homogeneous equation for transverse vibrations in a beam.
82u(0, t) 82u(L, t)
=0, and =0,
8x2 x2
and
0)
u(x, 0) = f (x) and au(t' = g(x). (10.53)
Equations (10.51 and 10.52) are linear and homogeneous. Therefore separation
of variables applies. If we let u(x, t) = p(x)G(t) and replace it in Equation (10.51)
we find that Equation (10.51) becomes
cp(x)G"(t) + c2cpiv (x)G(t) = 0. (10.54)
An easier form of Equation (10.58) is found by using the noncomplex and complex
forms of Euler's equations found in Chapter 5. Thus Equation (10.58) becomes
Cl = -C3.
Applying the third BC of Equation (10.57), cp"(0) = 0, yields
C1 = C3.
Applying the second and fourth BCs of Equation (10.57), cp(L) = 0 and cp"(L) = 0,
yields
and
0 = 2C2 sinh(A) 4 L,
and eigenfunctions
n7rx
cpn (x) =sin L (10.64)
n7r 4
Gn (t) - (i;-) cG(t),
which has the solution
u(x, t) = Al cos __ 2
t] + A2 sin
2
'n /nffS\
n=1
n=1
Therefore, the complete answer for transverse vibrations in a simply supported thin
beam is given by Equations (10.66, 10.67, and 10.68).
Other possible boundary conditions are
1. The cantilever beam, Figure (10.8), which has one end rigidly fixed and the
other as a free end. An example of this model is an airplane wing. It has the
Section 10.3: Transverse Vibrations of a Thin Beam 371
Figure 10.8: The cantilever beam: rigidly fixed on the left end and free on the
right end.
following BCs:
( u(0,t) =0
Rigidly fixed at x = 0 end
8u(0, t)
l 0
0
Free end at x = L end
0.
2. Beam rigidly fixed at both ends, Figure (10.9). A very good example of this
type of beam is the World War II fighter, Lockheed P-38 Lightning. The P-38
was designed as a twin engine fighter where the engines were attached to long
booms, the wings extend from the outside part of the boom, and between the
two booms was the central nacelle, which contained the pilot and armament.
372 Chapter 10: Classical Problems of PDEs
u(0, t) =o
Rigidly fixed at x = 0 end
{ 8u(0, t)
0
Ox
u(0, t) = o
Rigidly fixed at x = L end
Ou(0,t)
0 .
m
x
A 'r
EXERCISES 10.3
10.3.1. Solve the simply supported beam problem of length 27r when u(x, 0) = x(27r -
au(x, 0)
x) and =0.
at
10.3.2. Solve the simply supported beam problem of length 7r when u (x, 0) = x and
au(x, 0) _ 0
at
10.3.3. Solve the simply supported beam problem of length 1 when u(x, 0) = 0 and
au(x, 0)
at
-1-x.
-1-x2
10.3.4. State the BCs for a beam which is simply supported on one end and rigidly
fixed on the other end. Also, try to determine a physical application for these
BCs.
10.3.5. Given a cantilever beam where the right end is free, we know that after sep-
arating variables the spatial equation is
Show that Equation (10.69) satisfies the cantilever beam BCs if (\) 4 L is a
1 1
10.3.6. Given a beam that is rigidly fixed at both ends, we know that after separating
variables the spatial equation is Equation (10.69). Show that Equation (10.69)
satisfies rigidly fixed beam BCs if (,\) 4 L is a root of cosh 7) 4 L cos 7) 4 L = 1.
Section 10.4: Heat Conduction in a Circular Plate 373
_ (O2u(xYt) +
02u(x, y, t)
(10.70)
axe 82
y
subject to the homogeneous Dirichlet BCs
u(0, y, t) = u(L, y, t) = u(x, 0, t) = u(x, H, t) = 0
and
u(x, y, t) = f (x, y).
If you notice, I used the Laplacian (V2), then wrote out the partial derivatives with
respect to x and y. This makes the transition from a rectangular plate to a circular
plate transparent. The only thing that changes is the coordinate system. In a
circular plate, the preferred coordinate system is the cylindrical coordinate system,
which is called polar form in two dimensions, Figure (10.10).Thus, Equation (10.70)
becomes
8u(r, B, t)
_ k02u(r, B, t)
at
k
rl 8 (rO?(1',O,t)1 + 1 82u(r,B,t)1
(10.71)
Lr 8r or J r2 802 '
374 Chapter 10: Classical Problems of PDEs
subject to
u(a, B, t) = 0
and
Just as in Section 10.2.1, it appears that we only have one boundary condition.
However, as in Section 10.2.1 the other BCs come from the geometry of the problem,
and they are
u(r, -ir, t) = u(r, ir, t)
and (10.74)
limn-,o Ip(r,9) 00
and
subject to
= e(7r)
and
subject to
R(a) = 0
and (10.82)
(10.83)
m = 1,2,3,...
Om(e) = Am cos me + Bm sin me
We must solve Equation (10.81), restated as
r2R(r) + rR(r) + (\r2 - m2)R,,, (r) = 0, (10.84)
subject to
R,,,, (a) = 0
and (10.85)
I
< 0.
found in Chapter 8, and we know the general solution, which, for Equation (9.51),
is
Since we know that limo Rm (r) < 0 and as r * 0, Ym (V5r) 00, we must
have C1 = 0. Therefore, we are left with
For each Bessel function of the first kind of order m, we have an infinite number of
Os. Hence, just like a sine or cosine function, we have an infinite number of possible
solutions. Since each Bessel function of the first kind of each order m is a solution to
a linear homogeneous ODE, we know the sum of all the solutions is also a solution.
Thus, letting ymn represent the Os of Bessel function of the first kind of each order
m, we have
which implies
Amn
where
fo R,,,(r)Jm (/Xr) r dr
Cn -
fo Jm 2 r dr
which is
We now can give the general solution for u(r, 0, t), which is
yields
+ Jm (/r)mn
Emn sin m0. (10.92)
m=1 n=1
Thus,
and
f ", f o g(r, ('/Xr) sin (mO) r dr dB
E,,,, _ , m > 0. (10.95)
f fo [J,,,, (\/Ar) sin (mO)]2 r dr dB
378 Chapter 10: Classical Problems of PDEs
Therefore, the solution to the circular heat problem is Equation (10.89) where
the coefficients are given by Equations (10.93, 10.94, and 10.95).
A similar method is used to solve the wave equation for vibrations of a drumhead.
EXERCISES 10.4
au(r, 0, t) Ou(r, t)
(3)
ae
-0
- ae
,
- 0 u(1, 0, t) = 0.
For each case briefly explain what occurs when we take u(r, 0, t).
10.4.3. State and solve the general vibrating circular drumhead problem.
10.4.4. Solve the vibrating circular drumhead problem when the radius r = 1, the
boundary is fixed, the initial velocity is 0, and the initial displacement is
f (r) sin 20.
10.4.5. Solve the vibrating circular drumhead problem when the radius r = 7r, the
boundary is free, the initial velocity is 0, and the initial displacement is
f (r) cos 0.
10.4.6. Consider vertical vibrations of a wedge 0 < 0 < with radius 2. Determine
4
the solution if the BCs are as follows:
(1) u(r, 0, t) = 0, u (r, , t = 0, u(2, 0, t) = 0.
4
au(r,0,0,t) _ 0 au(r,0,3,t)
()3 ,0,z,t)=0.
az az
10.4.9. Consider
V2u=0
in a cylinder. Find the solution with the following BCs:
(1) u(r, 0, 0) = 0, u(r, 0, 5) = r sin 20, u(1, 0, z) = 0.
8u( B, 0)
(3) _ 0, u(r, B, r) = r cos B, u(1, B, z) = 0.
=r202au(r,0,2)=0u0.50z=0.
(4) 0,0)
Oz
Does the solution always exist?
380 Chapter 10: Classical Problems of PDEs
where = 27rv and v is the frequency of the radiation. The derivation of = 27rv
is far beyond the scope of this text. However, it is easy to show that G(t) = e-it
is a solution of
G"(t) = -2G(t).
The function ( is the solution of the time-independent equation
02 = -s2 c= . (10.98)
'Herman Helmholtz (1821-1894) started his professional life in physiology, then became inter
ested in mathematical physics. His studies were primarily in acoustics and he published results in
a work titled On the Sensations of Tone.
Section 10.5: Schrodinger's equation 381
E=mc2. (10.100)
E = hi', (10.102)
where h is Planck's3 constant and v is the frequency of the radiation. This implies
that
e-it = e-2lrivt =
e
which implies that the eigenvalues depend on the total energy, E. This is shown
later to be exactly the case.
Replacing energy, E, in Equation (10.101) with the formula for energy given in
Equation (10.102), we find that
(10.103)
Since a photon has speed and energy, it has momentum. Momentum for a single
particle of mass (a photon) is given by the equation
p = mv, (10.104)
p = Inc. (10.105)
2Albert Einstein (1879-1955), a brilliant physicist, developed the theory of general relativity,
which brought the field of differential geometry back to center stage in the mathematical world.
3 Max Planck (1858-1947) a German physicist who in 1901 took the first steps toward quantum
mechanics.
382 Chapter 10: Classical Problems of PDEs
In the mid 1920s, de Broglie's hypothesis was presented. It suggested that waves
were associated with material particles, with wavelength, A, and momentum, p,
where
A-h_ h _c (10.107)
p Inc v
Thus,
by h
p (10.108)
c A'
2irmc
Using Equations (10.105 and 10.107) we have c = Replacing s in Equation
h
2irmv
(10.99) by yields
h
V2 + (2 c1J 2 _ +
0. (10.110)
22(P=0. (10.111)
From physics, we know total that the energy, E, is equal to the sum of the
kinetic energy, Ek, and the potential energy, E. Also, we know the kinetic energy
mc2
is equal to 2 . Thus,
mc2
E=Ek+Ep= 2 +Ep,
Section 10.5: Schrodinger's equation 383
which means
mc2
2 =E-E. (10.112)
8ir2 m 8ir2 m
where q (x) _ - h2 E( x) and A = h2 E. Thus, the eigenvalues are based on
the total energy. Also, each eigenvalue, En, n = 1,2,3,..., has a corresponding
eigenfunction, Spn, n = 1,2,3,...
EXERCISES 10.5
where k is the spring constant and in is the mass. Note: ku represents the
restoring force. Dividing through by in in Equation (10.115) yields
u+wou=0
where wo 2 = -' which implies k = mwo2.
m
(1) Solve the ODE.
384 Chapter 10: Classical Problems of PDEs
(2) Derive the potential, Ep, by integrating the restoring force, ku. Note:
We must choose the constant of integration so that Ep = 0 when the
spring-mass system is at equilibrium, u = 0.
(3) Replace Ep in Equation (10.114).
(4) Multiply Equation (10.114) by h
2irmcao
h +ASP- 2irmwo
u2
2irmwo SP
SP=0.
+ (A - cp = 0. (10.116)
(7) We now look for solutions to Equation (10.116) on (-oo, oo) by substi-
tuting e 2 2 v(x) into Equation (10.116) and dividing out e 2 2 .
(8) Now let A = 2n + 1 and our equation becomes identical to Hermite's
equation given in Chapter 10.
(9) Solve and graph several Hermite polynomials on the interval (-5, 5).
10.5.2. PROJECT: Consider nuclear transport
(1) The Euler differential equation is one of the few differential equations
with variable coefficients that can be solved using a change of variables.
The change of variables necessary to transform the Euler differential
equation into a linear constant-coefficient differential equation is x =
ez . With this transformation, find the complete solution to the following
equations:
(a) x2 y" - xy' + y = x5.
(b) x2 y" - xy' + 2y = 1 + (lnx)2.
(2) You are working in the joint counter-terrorism unit with other federal
agencies. While searching for the plutonium brick (which was found
using information you supplied about the maximum size of the brick),
airport security forces discovered that the same group of terrorists were
trying to smuggle a 6.0 cm radius solid sphere of Pu-239 into the U.S.
disguised as a child's toy ball. The FBI wants to ship the ball back to a
Department of Energy (DOE) lab for evaluation. Since there is a Coast
Guard cutter docked at a nearby marina, they plan to send the Pu ball
packed in a crate on board ship. An FBI agent goes and gets a big crate
Section 10.5: Schrodinger's equation 385
full of small foam packing peanuts. (In the center of which he/she plans
to place the Pu sphere for shipping.)
The steady-state neutron diffusion equation with neutron multiplication
is:
v.7f SP
D V2 ( - a Sp + =0,
where
10.5.3. PROJECT: It is July 1945, and the future is uncertain. You are designing
the Little Boy atomic weapon. Little Boy is a "gun-type" weapon that uses
two identical subcritical cylindrical halves. One cylinder of U-235 remains
stationary at one end of the device, while the other (identical) subcritical
cylinder is fired using a chemical explosive as a projectile down a gun barrel
directly into the stationary cylinder. Provided other components are installed
correctly in the device, a nuclear yield results. Using the time-dependent
neutron diffusion equation with neutron multiplication you find that
1
a Sp Sp =
vo at
where
2 10(0(;o
r8r r8r+8z2
For boundary conditions, the origin can be taken to be the center of the
cylinder. The flux can be assumed to be 0 at the extrapolated boundaries,
and the initial (t = 0) flux profile is given as
p r +a t =0 a
± =+
a
+ 2D cylinder half height
2 ' 2 2
and
Note that Sp > 0 everywhere in the cylinder. Also, the following nuclear data
Section 10.5: Schrodinger's equation 387
Er = P(t) dt.
8ir2m
h2
(E-E)p=0
may be developed in spherical coordinates if the potential Ep depends only
on a distance r from some fixed point in space. Given the spherical r, e, and
as shown in Figure (10.11) where
show that
1 2
V2
r a (r2 8+ r Sin e ae
(sin0
-) + sin 8 (10.117)
(1) Show that Equation (10.117) may also be written in the form
2 a2 (,O 2( 1 La2 SP
e
a
+ 1 a2 SP
ae2 + cot ae
2
v 'P ar2 + r ar + r2 sine a q52
(2) Using Schrodinger's equation in spherical coordinates, show that sepa-
rating variables results in
i a eaW(e,
c5)1 + i aw(e, ) + aW(e, q) = 0. (10.119)
Sin eae ` ae J sine B a2
Section 10.6: The Telegrapher's equation 389
and
I:(q5) = 0. (10.121)
ti
(4) Show that I(q) = m = ±1, ±2, ±3, ±... is a solution of Equation
2
(10.121). Note: r = m What conclusions can you make about this
solution?
2
(5) Solve Equation (10.120) where r = m Also, set x = cos 0, O(8) _
y(x); only consider the case where m = 0 and A = n(n + 1), n =
0,1,2,3,... Graph the solution for several values of n. (Hint: After
replacing x = cos 0 and O(0) = y(x), your new equation should be
Legendre's equation; see Chapter 9.)
v ( x + d x, t) - v ( x, t ) _ -R dx i( x, t ) - L dx a2
(x,
t . ( 10 . 123 )
Similarly, the current at p2 is equal to the current at pl minus the current loss
through leakage to the ground and the current loss due to the capacitance. Thus,
according to Kirchoff's second law we have
8v(x, t)
i(x + dx,t) = i(x,t) - G dx v(x,t) - C dx at (10.125)
a2z(x,t)
ax2
- Lc a2i(x,t)
at2
+ RC ailx,t>
at
-G ax v(x,t)
(10.129)
Section 10.6: The Telegrapher's equation 391
v(x, t)
In Equation (10.129) if we substitute Equation 10.124( ) for the resulting
equation is the telegrapher's equation for i (x, t), which is
522(x't) Rca2(x,t)
= (10.131)
8x2 rat
The cell membrane acts or takes on the roll of the capacitor, denoted C = q
v
where q is the charge across the capacitor and v is the voltage potential. Since
we have both intracellular and extracellular space, we must consider voltage and
resistance of both. For the intracellular space, the voltage and resistance are denoted
as v2 (x) and v2 (x + dx) and Ri dx. For the extracellular space, they are denoted
as ve (x) and ve (x + dx) and Re dx. Also, there are two currents to consider;
the transmembrane current denoted It dx, which is the current running across
the membrane, and the axial current denoted Ia, which has components in both
the intracellular and extracellular space, denoted Iai and Iae , respectively, and
considered linear functions of the voltage. There also exists an ionic current, denoted
II (Note: When a neutral atom loses or gains one or more electrons, the loss of
electrons results in a positively charged ion and the gain of electrons results in a
negatively charged electron-gains and losses may occur during chemical reactions),
which occurs on the cell membrane. The different functions are as follows:
We start with derivation with the axial currents. The equations are similar to
Equation (10.122), and they are
8vz(x, t)
-Iai (x, t)Ri
ax
(10.134)
8ve(x, t) -Iae (x, t)Re.
ax
Thus,
1 8vz (x, t)
Iai (x, t) = Ri ax
(10.135)
Z 5ve(x,t)
Iae (x, t)
Re ax
We use Kirchoff's second law to calculate the change in axial current due to trans-
membrane current. We must have
Iai (x, t) - Iai (x + dx, t) = Itdx = Iae (x + dx, t) - Iae (x, t). (10.136)
We also have that the total axial current IaT (x, t) = Iai (x, t) + Iae (x, t) if no other
sources of current exist in the line. Therefore, using the expressions for Iae and Iae
given in Equation (10.135), we have
I _
aT
- 1 avi (x, t) _ 1 ave (x, t)
Ri ax Re ax
or
1 1 5v(x,t)
(x, t) 1 (5Ve(X,t) 5v(x,t)
(x, t)
Ri + Re 8x + Re 8x 8x
(x, t) _
1 5v(x,t) 1 5VT (x, t) - Re I aT. (10.141)
Ri ax Ri + Re ax Ri + Re
1 5v( x, t)
Since IaZ (x, t) _ - Equation (10.141) becomes
Ri ax
_ai
I ( xt) = 1 (x
A , T ax t _ RP.
RZ + Re
)
Rz + Re
IaT . (10.142)
which becomes
(c') II +
_ a 1
Z + Re
avT (x, t)
(10.144)
Note: Equation (10.144) is very similar to the telegraph equation for submarine
cables, Equation (10.131). Thus, Equation (10.144) is the telegraph or line equation
for a neuron.
If we assume that RZ and Re are constants, then Equation (10.144) becomes
5VT(x, t) _
II_p(RZ+Re)
1 a2VT(x, t)
(10.145)
C at + axe
Equation (10.145) may be normalized by taking into account certain constraints on
the membrane resistivity. Ignoring the extracellular resistance, and then nondimen-
sionalizing the resulting equation. Thus, we get a line equation for a neuron, which
is somewhat easier to work with; it is
C7v
a2 Z1
at - axe +f (v, t), (10.146)
where f (v, t) is a function of voltage and time. (Note: This is in the true tradition
of mathematical modeling. If you try to keep everything you sometimes end up
with an unworkable model, or a model that predicts nothing or the wrong answer
quite accurately. It follows the old military axiom: He who protects everything,
protects nothing at all.)
Many times f (v, t) can be modeled as -v when we are working with passive
activity, such as dendrites. For many other cells, the activity is called passive only
if the membrane potential is sufficiently small. Thus, Equation (10.146) becomes
C7v
a2 Z1
at - ax e - v. (10.147)
EXERCISES 10.6
a 0, t)
=0 and 0 (10.150)
396 Chapter 10: Classical Problems of PDEs
and
(1) Solve the Telegrapher's equation for both sets of BCs, Equations (10.149
and 10.150).
au(x' t)
22)[()
(2) Multiply Equation (10.148) by 2 and derive the differential iden-
tity
a au au a au 2 au\ 2
+ but - 2a au 2
+( =0.
ax at ax - at I
(3) Prove that if u(x, t) satisfies Equation (10.148) and either BCs, Equation
(10.149), or BCs, Equation (10.150), then
l t-
(4) State and prove the uniqueness theorem for the initial value problem
given by Equations (10.148, 10.149, and 10.151).
10.6.2. Given the Telegrapher's equation
1 a2u(x,t) a2u(x,t) au(x,t)
C2 at2 = ax2 -a at - 3u (x, t).
au(x, t) au(x, t)
Let u 1(x t) = u(x t), u2(x,t) = ax and u 3(x t) =
an and show
at
that ul (x, t), u2 (x, t), and u3 (x, t) satisfy the following system of three equa-
tions:
aul (x, t)
u3 (x , t) 0
at
au2(x, t) au3(x, t)
0
at at
au (x,t) _2 thL2(x,t)
+au3(xt)+ui(xt)) = 0.
at C ax
subject to
u(O,t) = 1 and u(L,t) =cost
Section 10.6: The Telegrapher's equation 397
and
u(x,0) = x3 - 2x + 1 and
tht' 0) =0.
x, 0<x<4;
7r
4,
7r
x<4,
37r
and
37r
7r - x, -j-< x< 7r.
Further suppose that the rod has no insulation on its ends and is plunged into a
bath, which is held at the temperature of 0° F.
(1) Find the temperature distribution of the rod for all time.
(2) Consider the point x = and estimate the error made in replacing the series
2
by its partial sum. Then, determine the time required for which the ratio of the
sum of all terms, starting with the second, to the first term, is less than s > 0.
x, 0<x<4r
ir ir < 3ir
x< (10.156)
4' 4- 4
3ir
it - x, < -x< ir.
4
Section 10.7: Interesting Problems in Diffusion 399
u(x, t) = brie-2.39n2t
sinnx. (10.158)
n=1
Using the IC, we solve for a specific solution, which is
00
4 nor nor
uxt= n 7r
(sin
2
cos
4
e-2.39n2t
sinnx. (10.159)
n=1
(__-4--)
sin cos
nor
2
nor e-2.39n2t
sin
nor
2
. (10.160)
The series on the right side of Equation (10.161) satisfies the alternating series test.
Therefore, we know that if s is the sum of the series and sn is the nth partial sum,
we have
I s - sn I <- zn+1
where zn+1 is the n + 1 term of the series. Therefore,
r2,t 00
4(_1)m
Sln
2m+1 it
cos
2m+1 it e-2.39(2m+1)2t
(2m+1) it 2 4
m=n+1
?r
where Is - sn I = Rn 2, t
We can now estimate the ratio of the sum of all the terms of the series starting with
the second term to the first term. This yields
R0 (,t) 2
4 (:) e-2.39(3)21
2
1 e-2.39(26)t
<
1 e-2.39(52)t <.
9 -9
400 Chapter 10: Classical Problems of PDEs
Thus, for t > t = -1 In 9Ewe have the ratio of the sum of all terms starting
with the second term to the first term less than > 0.
EXERCISES 10.7
10.7.1. Given a rod made of nickel with perfect lateral insulation and perfect insula-
tion on the end x = 0. Suppose the other end of the rod, at x = 15 ft, has a
convective heat exchange with a medium whose temperature is 0°F.
(1) Find the temperature distribution in the rod if the initial temperature
distribution is 100°F.
(2) Estimate the error made in replacing the sum of the series, representing
the point x = 15 by its partial sum.
2
insulation on the lateral surface of the rod and that the ends of the rod are
held at 0°C. Find the temperature distribution if the initial temperature
is uniformly 100°C throughout the rod.
(2) Suppose the cone is made by rotating the curve y = 3e-x, 0 < x < 5
about the x-axis. Suppose this cone is made of copper and is uniformly
heated to a temperature of 300°C. Further, suppose the lateral surface
area and the end x = 5 is perfectly insulated and the end x = 0 is held
at 0°C. Find the temperature distribution in the cone for all time.
10.7.4. Given a parallelepiped made of asbestos with 0 < x < it ft, 0 < y < 2ir
ft, and 0 < z < 3ir ft, find the temperature distribution if the sides of the
parallelepiped are maintained at 0°F and the initial temperature distribution
is 100°F. Find the time at which a steady-state will occur at the center of the
parallelepiped with relative accuracy > 0.
10.7.5. Consider a spherical shell made of mild steel with inner radius r1 = it m and
outer radius of r2 = 2 m. Suppose the inner and outer surfaces of the sphere
have a convective heat exchange with a medium whose temperature is 0°C,
and that the initial temperature of the spherical shell is r2 + 1 for r1 <r <r2.
10.7.6. Find the temperature distribution of a rod of length 2ir m, that has perfect
lateral insulation, consisting of two homogeneous materials: the first half of
the rod is made of aluminum, the second half of the rod is made of silver.
Suppose the left end of the rod is held at 0°C and the right end has a heat
flow of sin t watts, where t is time, and the initial temperature of the rod is
sin x, 0 < x < it
6(cosx-1),
10.7.7. A spherical vessel filled with gas moves uniformly for a long time with ve-
locity v0, and then at time t = 0 it is stopped instantaneously and remains
stationary. Find the vibrations of the gas in the vessel.
10.7.8. Find the vibrations of a gas in a spherical vessel of radius r = 3 produced by
small deformations of the wall, beginning at time t = 0, if the velocities of the
wall are radial and equal
6Pn(cos 8) cos(wt).
Fourier Integrals
and Transform Methods
11.1 INTRODUCTION
In Chapter 3, we discussed d'Alembert's solution to the two-dimensional wave equa-
tion. The particular solution developed was for an infinite string. So far this is the
only method we discussed for infinite boundaries. In the current chapter, we will
develop methods that allow us to solve linear second-order PDEs with semi-infinite
and infinite boundaries. These methods involve transforms.
You were first introduced to transforms in your ordinary differential equation
course. There you solved linear initial-valued ODEs with Laplace transforms. The
method involved transforming an initial-valued ODE into a space where the equa-
tion could be solved algebraically. Once the algebraic solution was found, the inverse
Laplace transform is applied returning you to the space of the original problem. In
ODEs, tables of Laplace transforms were used to simplify the process. Later, in this
chapter, we will again visit Laplace transforms. Here, we use them as another solu-
tion method for initial-valued PDEs with finite and semi-infinite boundaries. Again
tables of Laplace transforms are provided to simplify the process. However, unlike
solving ODEs by Laplace transforms, solving PDEs by Laplace transforms does not
generate an algebraic equation to be solved but an ODE one to be solved. Before
we solve PDEs with Laplace transforms, we will introduce the Fourier integral.
The Fourier integral is a natural extension of Fourier series. By extension, we
mean the representation of a piecewise-smooth function with a Fourier integral
where the domain of the function is semi-infinite or infinite. This is quite different
than Fourier series where the domain is typically [-L, L]. However, as in Fourier
series representation of a function, there are certain restrictions on a function rep-
resented by a Fourier integral. These restrictions are discussed in some detail.
Once the Fourier integral is developed, the Fourier sine and cosine integrals are
discussed. The Fourier sine and cosine integrals are the natural extensions of the
Fourier sine and cosine series, respectively. Next, we introduce transform solution
403
404 Chapter 11: Fourier Integrals and Transform Methods
which is shown in Figure (11.2). Please note that Figure (11.2) exhibits the Gibbs
phenomenon, and that at x = ±2 the Fourier series converges to -9.
The Fourier series representation of x3 - 3x2 - 2x+ 3 as a function is considered is
periodic on [-oo, oo]. Figure (11.3) shows this periodic nature of the Fourier series
function using x3 - 3x2 - 2x + 3 on [-2, 2] as the original function represented.
Thus, we speak of a Fourier series involving "periodic" functions. The question
which arises is; what do we do with a nonperiodic function on [0, oo] or [-oo, oc],
such as
0, -2<xorx>2
f (x) _ (11.2)
x3-3x2-2x+3, -2<x<2.
Section 11.2: The Fourier Integral 405
Figure 11.3: Graph of Fourier series function on [-oo, 00] where n = 150.
00
where
ao
2L
1 f L
L
1(x) dx,
fL
an L f (x) cos (nLx) dx,
L
and
1 L
bn 1(x) sin (L ) dx.
L J L
nor
If we let 8 = , and replace ao, an, and bn in Equation (11.3) with their integral
L
representations, we have
f (x) dx +[fl(x)
1 L o0 1 L
f (x) =
2L f L
- cos (ex) dx cos (ex)
L
n=1 I'
+
[fL 1(x) sin (Ox) dx sin (Ox)}. (11.4)
L
fL
+ e 1(x) sin (Ox) dx sin (Ox)
J L
1 00 L
1 L
2L, 1(x) dx +-ir IfJ 1(x) cos (Ox) dx OB cos (Ox)
n=1 L
fL
+ 1(x) sin (8x) dx OB sin (Ox)
L
J
[fL
L
1(x) dx + L
1(x) cos (Ox) dx cos (Ox)
Second,
{ 1(x) cos (8x) dxJ cos (Ox) + 1(x) sin (Ox) dxJ sin (Ox) } D8
ll J
I
LJ J
F(B) = 1(x) cos (Ox) dxJ cos (Ox) + I 1(x) sin (8x) dxJ sin (Ox)
LJ 0o L J o0
1 F(B) dB,
Jo
which becomes
J- f (x) dx
exists. Then, at every point x where the one-sided derivatives of f (x) exist, the
function f (x) may be represented by the Fourier integral
Please note, in Theorem (58), that the absolutely integrable function, 1(x),
eliminates all periodic functions except 1(x) = 0. Also, a more convienient form of
Equation (11.9) is
where
A(B) _ (11.11)
f:f()cos(o)d
and
0, -1 <xorx> 1
f (x) = 1, -1<x<0
2, 0<c<1,
shown in Figure (11.4). The Fourier integral of the function, 1(x), given in Equation
2-
2.5
1.5
0.5
x
-10.5
-1
(11.13) is
where
A(B) _ f d
J
(J °i d+ 12 d}
{f° Jo J
3 sin(B)
eir
and
B(e) - f
1/00
f d
(J ° sin(O) dx + sin(e) d }
i
{f° Jo 12 J
1-cos(O)
eir
Therefore,
3
2.5
2
1.5
-3
A(B) _
J
d
- 1
p2
- 2 cos(ec) d
4 sin(2B)
Bir
and
B(B) _ fg(e)sin(O)d
l2 - 2
1 J2 d
=0.
Therefore,
g(x) = J
/'°O
0
-e 4 sin(2e)
cos(Bx) d8. (11.16)
Section 11.2: The Fourier Integral 411
2 2
x
-3
Figure (11.7) shows the graph of the Fourier integral representation of g (x) . The
vertical bars from -2 to 0 at x = ±2 are a result of the mathematical software being
used.
0, x<-lorx>l
h(x) = -1, -1<x<0
1, 0<x<1.
Figure (11.8) shows the graph of the Equation (11.17). The Fourier integral of the
2
1.5
1
0.5
x
-1 1
-0.5
-1.5
-2
where
1 f°O
d
lf - d+f1
o
de}
and
B(B) = 1 d
-00
J °
_{f1
l i
- d+
Jo
1 sin(9)d}
J
2r1- s(8)1
l J
Therefore,
°° 2 1 - cos(B)
h(x) = sin(ex) de. (11.18)
0 a
Figure (11.9) shows the graph of the Fourier integral representation of h(x). The
vertical bars from -1 to 0 at x = -1 and 1 to 0 at x = 1 are a result of the
mathematical software being used.
x
1
0.5
-1.5
-2
Fourier's integral,
f(x) = f
0
{A(O) cos(ex) + B(e)Sin(ex)} de, (11.19)
where
A(8) = 1
J f()
d (11.20)
Section 11.2: The Fourier Integral 413
and
x= 1
is another form of Fourier's integral.
o f f() cos B x- d d8 (11.22
We develop Equation (11.22) from Equations (11.19, 11.20), and 11.21) in the
following manner. First, Equations (11.20 and 11.21) for A(e) and B(e) are
and
by the cosine of the difference of two angles' identity and the fact that the cosine
function is an even function. Finally, we place Equation (11.28) in its proper spot
in Equation (11.26), which yields
f(x) _ (11.29)
ir J J
414 Chapter 11: Fourier Integrals and Transform Methods
f()eZ°(x-)
1(x) _ (11.31)
and
oo oo
f(x) - 2ir f eiexde f (e)e-ZB (11.32)
Equations (11.31 and 11.32) are known as the complex form of Fourier's integral,
and their derivation is left as exercises.
Since we have considered a Fourier integral as the limiting case of a Fourier
series, it is natural to assume that the limiting case for the Fourier sine and cosine
series are Fourier sine and cosine integrals. This topic is discussed in the next
subsection.
f(x) = f
0
{a(e) cos(ex) +B(9) Sin(ex)} de, (11.33)
where
A(B) = 1
f:fcosod (11.34)
and
B(B) = 1 (11.35)
0, -2<xorx>2
g(x) _ (11.36)
-2, -2<x<2
was represented by a Fourier integral. However, since g (x) is an even function, B (O)
in Equation (11.35) equaled 0. Thus, A(e) is an integral of g(x) cos(Bx), which is
Section 11.2: The Fourier Integral 415
the product of two even functions. Therefore, the integral A(B) could be written as
A(8) _
itJ0
d
o J -z d
2 f2
J- 2 (11.37)
Equations (11.38 and 11.37) are an example of the Fourier cosine integral. For-
mally, the Fourier cosine integral for even functions, 1(x), is given as
In Example (11.4), we examine the Fourier cosine and sine integrals with the
function 1(x) = 2e-x, where 0 <_ x.
shown in Figure (11.10). First, we find the Fourier cosine integral of 1(x). This
416 Chapter 11: Fourier Integrals and Transform Methods
means, we compute
b
4
- lim e- cos(Be) d
7r b-9oo U
-6 \
{B sin(b9) - cos(bB)}
7r (1 + 92 + e 1 + 92
41 (11.44)
r 1 + 9V
Figure 11.11: Graph of The Fourier cosine integral of 1(x) = 2e-x, where 0 < x.
Section 11.2: The Fourier Integral 417
f (x) = fA(e)cos(ex)dO
Jo 1+92 cos(ex) dO
cos(9x) dB, (11.45)
0 1 + B2
which is shown in Figure (11.11). Now, we find the Fourier sine integral of 1(x).
This means, we compute the following:
B(9) = 2 ff()sin(9)d
o2e- sin(9) d
fe _ sin(g) d
b
4
lim
7r b-9oo 0
4 /1+92
9 -b \
{sin(bO) +9 cos(b9)}
n i. 1 + 92
49 (11.46)
7r l+92 .
Thus, the Fourier sine integral of 1(x) = 2e-x is
Figure 11.12: Graph of The Fourier sine integral of 1(x) = 2e-x, where 0 < x.
P004 e
sin(9x) d9
J0 ?r 1 + 92
4 °° e
sin(9x) d9, (11.47)
?r o 1 + 92
418 Chapter 11: Fourier Integrals and Transform Methods
where the variable of integration in the second integral was changed to for conve-
nience. Also, an important form of the Fourier sine integral is
Again, the variable of integration in the second integral was changed for convenience.
Before we proceed with the Fourier transform, our next section discusses Laplace
transforms and how they apply to PDEs.
EXERCISES 11.2
11.2.1. Find the Fourier integral representation of the following functions and graph
the solution using an algebraic software package:
0, x>33
(1)
(3)
(4) f(x)=
Section 11.2: The Fourier Integral 419
x, 1<x<0
(5) x2, 0<-x<1
0, x>1.1.
sin x, IxI < it
(6) 1(x) {
0, x > 7r.
(8) f(x)=
xI> 2.
x2-1, -3<-x<0
x-1, 0<-x<l
(9) f(x)=
-x2, 1<x<3
(xI> 3.
0, x<-1
(10) f(x)=1 -x-l, -1<- x<0
_e-x, 0<- x.
11.2.2. Starting with Fourier's integral, Equation (11.9), show that
-
1 °° °°
1(x) = f ()ei9(x-) ddb
is another form of Fourier's integral.
11.2.3. Starting with Fourier's integral, Equation (11.9), show that
1 oo oo
1(x) = 27r _
J
eZex
d8
_00
f ()e-Zed
d
420 Chapter 11: Fourier Integrals and Transform Methods
x2-1, (xI< 1
0, (xI> 1.
2 0<x<1
-
f cos (Ox' sin
e
0,
x=1
1<x.
11.2.5. Find the Fourier sine and cosine integral representation of the following func-
tions and graph the solution using an algebraic software package:
1, 0<x<1
(1)
{ 0, 1<x.
x3, 0<x<1
(2) 1(x) { 0, 1<x.
ex, 0<x<1
(3)
{ 0, 1<x.
25-x2, 0<x<5
{ 0, 5<x.
Section 11.3: The Laplace Transform and PDEs 421
x, 0<x<2
(s) f(x)= 4-x, 2<x<4
0, 4<x.
1
1<x<2
x
(6) f(x)
={ 0, otherwise.
(7) f(x)=e_x+e_2x, 0<x.
f (x) = 2 fcos(Ox)dOff(cos(O)d
?
J0 f() sin(O)
f (x) _ sin(Ox) dB
Jo
is another form of Fourier's sine integral.
Before we solve a PDE with Laplace transforms, we'll state several important
theorems, list some known properties, and refresh our memories by solving an ODE.
(Note: The proofs for all the properties may be found in any standard ODE text.)
Next, we discuss the error function, which is also known as the probability integral.
Finally, we solve PDEs using the Laplace transform.
f f (t) dt = lim
6-.00 0J
b f (t) dt,
where b is a positive real number. If the integral exists from 0 to /3 for each /3 < b,
and if the limb+00 exists, then the integral is said to converge to the limiting value.
If any part the previous statement is false, then the integral is said to diverge.
Given this brief description of an improper integral, we state the primary theorem
for Laplace transforms.
Theorem 59. Suppose that f is a piecewise-smooth function on the interval 0 <
t < T for any T E g+ and f (t) < beat when t > M and where a E Ilk and M and
k E Ilk+. Then, the Laplace transform is given by
The following list contains several more properties of Laplace transforms. You
should be familiar with all of them.
Properties of Laplace transforms
G{af(t)}=aG{f(t)}.
£ {f(t) + g(t)} = £ {f(t)} + £ {g(t)}.
Section 11.3: The Laplace Transform and PDEs 423
f(t')dt' = F(s).
( o /
£{tf(t)} - _dF(s)
£ {f(t)}
t
= fF(s') ds'.
Theorems (59 and 60) and the other properties of Laplace transforms provide
the necessary tools for solving initial-value ODE problems. Consider the following
example:
EXAMPLE 11.5. Consider
u"(t) + 2u'(t) - 3u(t) = 0, (11.52)
subject to
u(0) = 1 and u'(0) _ -2. (11.53)
The Laplace transform of Equation (11.52) is
G {u"(t)} + 2G {u'(t)} - 3G {u(t)} = 0.
Using Theorem (60) and the initial conditions, Equation (11.53), the Laplace trans-
form of Equation (11.52) becomes
s2G {u(t)} - su(0) - u'(0) + 2 [sL {u(t)} - u(0)] - 3G {u(t)} = 0
or
It should be noted that transforms come in pairs. Theorem (59) gives a definition
of the Laplace transform. The inverse Laplace transform requires knowledge of
contour integration. However, the inverse Laplace transform is
2 f°°e_u2du.
erf (x) _ e-"z du = 1 -
J J0
Also, there is a complementary error function, denoted erfc (x), where
It should be noted that erf (0) = 0 and erfc (0) = 1. Following is a proof of erfc
(0) = 1, from which we find that erf (0) = 0.
Proof. Let
e-x,
00
2
I= d
2
Jo
where we changed the variable of integration in the second integral. I2 may also
be written as
4 fOC OC
dx dy, ( 11.55 )
o J
transferring to polar coordinates where r2 = x2 + y2 and 0 = tan- 1 y . Also,
x
changing to polar coordinates requires a change to the limits of integration. Thus,
Section 11.3: The Laplace Transform and PDEs 425
for r, we have 0 <r < oo, and for 0, we have 0 <0 < Thus, Equation (11.55)
2
becomes
4 2 00
_r,2 4/1:1 1 00
e rdrd8=_
-2e_T2 de
0 J
0 0 0
2
=2 d8
it o
Figure (11.13) shows the graph of erf (x) and erfc (x).
In Figure (11.13), please note that as x - Oo, erf (x) - 1 and erfc (x) - 0. Also,
in Appendix F, there are several useful Laplace transforms of the complementary
error function. These transforms will be needed as you solve the problems at the
end of this section and chapter. Next, we look at Laplace transforms and PDEs.
b
1
= sin 7rx lim -1 e-(s+3)t
( )b-+oo [s+3 0
- sin(irx) 1l = sin(irx)
lim U
s+3 t-+L J s+3
EXAMPLE 11.7. Find the Laplace transform of cos(x+t). By Definition (11.56),
we have
f
b
b b
= lim
b-> o0 f e-St cos x cos t dt -
0
e-St sin x sin t dt
b b
= cos x lim a-St cos t dt - sin x lim a-St sin t dt
b- +o0 0 b- +o0 0
e-St b
e-st b
cos x 1e-Sb
lim (-s cos b + sin b) + s]
s2 + 1 boo
Section 11.3: The Laplace Transform and PDEs 427
sin x
lira [e-Sb (-s sin b - cos b) +1]
s 2 -I- 1 b- +o0
{ a ua 2' t)
j =Sec {u(x, t)} - Su(x, o) -
auat, °)
0)
= s2U(x, s) - su(x, 0) - au(x,
at
(11.58)
and
,oo
£1
82u(x, t)
e
-st a2u(x,
axe
t)
dt
8x2 0
p00 a2
JO
foo
dx2
[eu(x,t)] dt d=2U d ,s (11.59)
0
respectively.
A close examination of Equations (11.57 and 11.58) indicates that the Laplace
transform is ideally suited for PDEs with initial conditions, in particular, the heat
and wave equations in one spatial dimension where the spatial variable has either
finite or semi-infinite BCs. Also, Equation (11.59) informs us that the Laplace
transform of a linear second-order PDE is a linear second-order ODE. The first two
examples demonstrate this important concept.
EXAMPLE 11.8. Suppose that we have a very thin rod with perfect lateral insu-
lation of length L. Further suppose that the rod has an initial temperature distri-
bution of a sin
x and the ends are held in a 0° bath. Find the time-dependent
L
428 Chapter 11: Fourier Integrals and Transform Methods
and
u(x, 0) = a sin L)
7rx
.
First, we find the Laplace transforms of all equations. We have, by Equations (11.57
and 11.59),
£1J t)1 l
= Sc {u(x, t)} - u(x, o) = Su(x, S) - u(x, o)
= sU(x, s) - a sin L
7rx
(11.60)
and
G f 82u(x, t) d2U(x, s)
(11 . 61)
8x2 } dx2
For the particular part, we use the method of undetermined coefficients. This
method may be found in Appendix C. The solution for the particular part is
aL2 irx
Up(x, s) = sL2
2 sin L (11.65)
Section 11.3: The Laplace Transform and PDEs 429
2
irx
= cl sinh(/x) + c2 cosh(x) + SLa+ ir2
sin (-i-). (11.66)
2
(lrXL
sin (11.67)
U(x' S) sLa + 2 1
Finally, we find the inverse Laplace transform of Equation (11.67), which yields the
time-dependent solution u (x, t). Note: A table of inverse Laplace transforms may
be found in Appendix F. We have
2 l
u(x, t) = G-1 {U(x, s)} = G-1 I sin (L)
SLa+ 2 J
2t '7X
= ae-T sin L (11.68)
EXAMPLE 11.9. Consider a perfectly flexible string of length L with the ends
attached to frictionless sleeves, which move vertically up and down. Further, sup-
pose the string has no initial displacement. However, the initial velocity is given as
7rx
cos (). Find the displacement u (x, t).
L
The physical model may be described by the mathematical equations
a2u(x, t) a2u(x, t)
ate - axe
subject to
a a(0, t)
= 0 and =0
and
First, we find the Laplace transforms of all equations. We have, by Equations (11.58
430 Chapter 11: Fourier Integrals and Transform Methods
and 11.59),
l I
=s U(x,s)_cos._
x (11.69)
and
c a2u(x, t) d2u(x, s)
(11.70)
{ 8x2 } dx2
Also, the BCs become
8u(0, t)
8x
_ dU(0, s)
dx
= 0 and C
8u(L, t)
ax
Next, we combine Equations (11.69 and 11.70). This yields a nonhomogeneous
_ dU(L, s) _ 0.
dx ()
11.71
d2Uh(x, s) 2
Uh(x, s) = 0,
dx2
which has the solution
For the particular part, we use the method of undetermined coefficients. This
method may be found in Appendix C. The solution for the particular part is
L2 irx
Up(x, s) = s2L2 2 cos (11.74)
L
Combining Equations (11.73 and 11.74) yields
L2 (irx)
= cl sinh(sx) + c2 cosh(sx) + s2L2 + 2 cos . (11.75)
Section 11.3: The Laplace Transform and PDEs 431
Finally, we find the inverse Laplace transform of Equation (11.76), which yields the
time-dependent solution u(x, t). We have
u(x,t) = £'{U(x,s)} =
L2 xl 1
cos (L / J
'{ L2 + ir2
L
sinrirt
(L cos l L (11.77)
= sU(x, s) (11.78)
and
G 82u(x, t) kd2U(x, s)
(11.79)
{ 8x2 } dx2
432 Chapter 11: Fourier Integrals and Transform Methods
Next, we combine Equations (11.78 and 11.79). This yields a homogeneous linear
second-order ODE, which is
We usually write the solution of Equation (11.81) in terms of hyperbolic sine and
hyperbolic cosine functions. However, in this case, it is easier to write the solution
in terms of exponential functions. Therefore, we have
svT
U(x, s) =
s + c2 + (11.83)
Applying the second BC, lim0 U(x, s) = 0, from Equation (11.80) to Equation
(11.82) implies c2 = 0. Thus, the solution is
U(x S) = (11.84)
sf
Finally, we find the inverse Laplace transform of Equation (11.84), which yields the
time-dependent solution u (x, t). We have
The last two sections of this chapter deal with the Fourier transform. In the
next section, we discuss the Fourier transform and the properties of the Fourier
transform, which are similar to those of the Laplace transform. In the final section,
we use Fourier transforms to solve PDEs.
EXERCISES 11.3
erfc (i).
11.3.3. Show that
e-a/
is Laplace transform of
-a{erfc
()}.
11.3.4. Show that
e_'
is the Laplace transform of
{erfc (bj
J
434 Chapter 11: Fourier Integrals and Transform Methods
2 (1)flx2fltl
(4) erf (x) _ 7= n_o
n!(2n + 1)
11.3.7. Solve the following boundary value problems
(1)
8u x,
at t 2
4a axe t), x > 0, t > 0,
subject to u(0, t) = 4; lim u(x, t) = 0 and u(x, 0) = 10.
au(t,t)
a2a(2,t)
(2) _ ko , x > o, t > o,
subject to u(0, t) = a; lim u(x, t) _ Q and u(x, 0) = y.
auat (x,t)
azax2,t)
(3) 2 , x >0, t >0,
subject to a t = 0; lim u(x, t) = 0 and u(x, 0) _ -40.
(o,
0716 2 , a2u(x,t)
8u(0, t)
subject to ax - a; lim u(x, t) = ,3 and u(x, 0) = 'Y.
(5)
8u x,
at
t 2
aax2, t)
, x > 0, t > 0,
subject to a 0, t) _ t)
= 0 and u(x, 0) = 25.
Section 11.3: The Laplace Transform and PDEs 435
au(t,t)
a2ax2,t)
(s> = ko , x > o, t > o,
subject to au(0, t) = a' lim au(x' t)
_ ,6 and u(x 0) = ry.
(7x (7x
au(t,t) azaxe'
(7) _ .22
t> >
x >0, t >0,
subject to u(0, t) = 0; lim a t _ -10 and u(x, 0) = y.
(x,
au(t,t)
a2ax2,t)
(a) _ ko , x > o, t > o,
8u(x, t)
subject to u(0, t) = a; lim ax Q and u(x, 0) = ry.
(3)
D2u(x,t) - 9D2u(x,t)
t)
subject to = 0; lim au(x, t) = 0 and u(x, 0) = 0;
au(0,
8x 8x
Du(x,0)
at
a2u(x,t) _ 2a2u(x,t)
(4) -C , x > 0, t > 0,
ate axe
au(0, t) au(x, t) =
subject
J to = a lim
x-oo ax
and u x 0)) = 0
ax C
au(x, 0)
D2u(x, t) D2u(x, t)
() Dt2 Dx2
436 Chapter 11: Fourier Integrals and Transform Methods
a2u(x, t) 2 a2u(x, t
at2
=c ax2
x>0 t>0
,
au(0, t)
subject
J to =a; lim
x- +o0
u(x,t)= and u(x 0)= 0
ax
8u(x, 0)
8t = -vo4.
a2u(x, t) a2u(x, t)
VL" Vim"
8u(0, t)
subject to ax - 0; lim u(x, t) = 0 and u(x, 0) =
8u(x,0) _ 0
at
au(0, t) _a
subject to lim u(x, t) _ ,6 and u(x, 0) = f (x);
ax
8u(x, 0)
=
8t
11.3.9. Suppose you have a uniform tightly stretched string from x = 0 to x = L.
rx
Find the displacement of the string if the initial displacement is A sin L
and the initial velocity is 0.
a2u(x, t) a2u(x, t)
at2 - ax2
<x<2ir,t>0,
subject to
subject to
8u(0, t)
= 100 - u(0, t); lim u(x, t) = 0 and u(x, 0) = 0.
8x X-* o0
(2) Find the displacement of the string for any time t of the force is acting
negative y-direction where v c.
(3) Repeat part (1) and (2) where v = c.
(4) Suppose Fo = 1. Use mathematical software of your choice to graph the
solution when v = 1 c when v = c when v = 3 c.
2' 2
11.3.18. PROJECT: Find the steady state temperature distribution in a thin uni-
formly insulated sheet which corresponds with the upper half plane where the
temperature is maintained at
0 IxI>ir
and u(x, y) -* 0 as y -p oo. Hint: Use Laplace's equation. Also, there is not
an infinite set of eigenvalues, instead there is a continuous family of solutions.
11.3.19. PROJECT:
(1) Suppose you have a homogeneous uniform one-dimensional rod with per-
fect lateral insulation and its ends, at x = 0 and x = L, in 0 degree baths.
Further suppose that the initial temperature of the rod is ax, for a a real
constant. Show that the temperature distribution in the rod, u(x, t) may
be expressed both by a Fourier series and by using Laplace transforms.
Hint: The Laplace transform should be the infinite sum of the difference
of two error functions, erf.
(2) In a one page paper, discuss which solution you think will converge more
quickly for a short period of time and for a long period of time.
(3) Suppose your rod is made of cast iron and is 10 cm long, calculate the
temperature of the midpoint of the rod when
(a) a = 50, t = 10 seconds, t = 1 min, t = 10 min, and t = 50 min.
(b) a = -50, t = 10 sec, t = 1 min, t = 10 min, and t = 50 min.
In each case, determine if your suspected convergence in Part (2) is cor-
rect.
(4) Suppose your rod is made of cast iron and is 10 cm long, calculate the
temperature of the midpoint of the rod when the initial condition is given
as
x,0<x<5
u(x,0) _
10-x,5<x<10
Section 11.4: The Fourier Transform 439
11.3.20. PROJECT:'
(1) Find the solution of the semi-infinite cable equation for a neuron
av a2 v
- v, (11.86)
at - axe
subject to the clamped voltage IC
v(x, 0) = 0
f (x) =
Fourier integral form in Equation (11.32), which is restated here
1
2ir
f(x)=J00
eZexde J d,
'Adapted from James Keener and James Sneyd, Mathematical Physiology, ©1998 by Springer-
(11.90)
f f (S)e-ie£'"S
is a function of 0 and is known as the Fourier transform of 1(x). Thus, the Fourier
transform is stated as in the following definition:
Definition 61.
Since Equation (11.90) produces the Fourier integral of f (x), the inverse Fourier
transform is given by Definition (62)
Definition 62.
f(x)= (11.93)
0, Ix>5.
Then, the Fourier transform is
10
3
x
-S
x
-2
2
-3
F{f(x)}=F(O) =
5 e_iee s
£ t
-ied
-5S e -2B
where B # 0. For B = 0, F(9) = 10. Figure (11.14) is the graph of the function
f (x). Figure (11.15) is the graph of the Fourier transform, F(9), of the function.
In general, if
11, IxI<a
f(x) _ (11.95)
0, IxI>a,
then, the Fourier transform is
eia9 _ e-ia9 sin a8
(11.96)
.F{S(am)} = F(B) _ f d
f e-e-ie d
-a
a
f -a
e-(1+ie) d
a
_e-(1+ie)
1+2O -a
e(1+i9)a _ e-(1+i9)a
1 + iB
- e 2 i(e-i)°' _ 2511i(e(e 2)). (11.98)
i
442 Chapter 11: Fourier Integrals and Transform Methods
From the previous example, we see that the Fourier transform of a relatively
common function quickly becomes complicated. For your convenience, there is a
table of Fourier transforms in Appendix F. We now consider Fourier cosine and
sine transform pairs.
f(x) =
o
2f cos(ex) de f 0
cos(ee) ae, (11.99)
f() cos(O) d
J f() sin(0)
1(x) _ sin(9x) d8 (11.102)
J
Again, the integral
f() sin(Be) d
J0
is a function of 8, and it is known as the Fourier sine transform. Its usual form is
EXAMPLE 11.13. Find the Fourier cosine and sine transform of the function, as
shown in Figure (11.16),
1, 0<x<a
f (x) _ (11.105)
L 0, a < x.
2.5
2
1.5
0.5
x
4
-0.5
-1
d
1:1(e)
jcos(Oe)
a
d
a sin(a8)
(11.106)
0
e
Figure 11.17: The Fourier cosine Figure 11.18: The Fourier sine
transform, F(B) of 1(x). transform, F(9) of 1(x).
444 Chapter 11: Fourier Integrals and Transform Methods
CoS(ee) a 1 - CoS(aB)
(11.107)
B o B
Figure (11.16) is the graph of the function 1(x) for a = 4, Figure (11.17) is the
graph of the Fourier cosine transform of 1(x) for a = 4, and Figure (11.18) is the
graph of the Fourier sine transform of 1(x) for a = 4.
We now develop some necessary theorems about Fourier transforms, which we
need when solving PDEs.
Since integration is a linear operator, the right side of Equation (11.88) becomes
cl f (_i0 d + C2
g(e)e-Zed
d,
which equals
elf { f (x)} + c2J:'{9(x)}.
D
Fourier transforms like Laplace transforms are a linear operator. Since Fourier
transforms are going to be used to solve PDEs, it would seem reasonable to hope that
they transform a first, second, and nth derivatives similar to Laplace transforms.
We find this out in the next theorem.
Section 11.4: The Fourier Transform 445
Theorem 64. Suppose 1(x) is a continuous function on the x-axis and 1(x) - 0
as fxf -p oo. Also, suppose f'(x) is absolutely integrable on the x-axis. Then,
Proof. Since f'(x) is absolutely integrable, we know that a Fourier transform exists.
Therefore, we have
f .fF(e)e-ZB£d
=0.
Therefore,
(iO)J'{f(x)}. (11.111)
{f * g} = f (y)e-iey dy
g(s)e-Zes
ds.
i_:
Thus,
F{f*g} =J:{f}J:{g},
which completes the proof. D
Note: The method of convolution is used extensively when solving the heat
equation in an infinite spatial variable domain. If we let J {f} = 1(0) and J {g} =
g (B), then we have
j(9
(.f * 9)(x) - 1-00
(0)eiOX dB.
Theorem (64) is similar for Fourier cosine and sine transforms, which we state
here for future reference. For the Fourier cosine transform, we have
and
and
EXERCISES 11.4
11.4.1. Compute and graph the Fourier transforms of the following functions:
2, f0,
(1)
{
5<x.
e-2x1,
xf < 1
(2) 1(x) {
0, otherwise.
e-x, x>0
(3)
{
0, otherwise.
ex, x<0
(4) f(=) {
0, otherwise.
xe-2x, x > 0
(s)
{
0, otherwise.
-1, -1<x<0
1, 0<x<1
0, otherwise.
1
!]1 flml -
x2+25
x, 0<x<2
(8) f (x) = 2x-4, 2<x<4
0, otherwise.
448 Chapter 11: Fourier Integrals and Transform Methods
11.4.2. Compute and graph the Fourier cosine transforms for the following functions:
(1)
(2) f(x) =
3
(s)
x2 +9 .
x, 0<x<1
(6) J(=) { 0, otherwise.
x, 0<x<2
(7) 2x-4, 2<x<4
0, otherwise.
11.4.3. Compute and graph the Fourier sine transforms for the following functions:
(1)
0, otherwise.
(4) 1(x) = x.
Section 11.4: The Fourier Transform 449
x
(s)
x2 +9 .
x, 0<x<1
(6) f(x)= {
0, otherwise.
x, 0<x<2
(7) 2x-4, 2<x<4
0, otherwise.
-1, 0<x<1
(8) f(x)= 1, 1<x<2
0, otherwise.
11.4.4. This exercise develops the Fourier transform of a shift of a function on the
x-axis. Show that if 1(x) has a Fourier transform, then f (x - a) for some
a E 1R, and that
11.4.5. This exercise develops a shift on the 0-axis. Show that if f(0) is the Fourier
transform of 1(x), then f(0 - a) is the Fourier transform of eiax f(x)
11.4.6. This exercise develops the basic properties of convolutions. Prove the following
properties of convolutions:
(1) Commutativity of convolutions: f * g = g * f.
(2) Associativity of convolutions: f * (g * h) = (f * g) * h.
(3) Commutation with translation: Given a E R, let fa (x) = f (x - a). Show
that fa * g = f * ga = (f * g)a.
(1) Show that the above two definitions are equivalent to Definitions (61 and
62).
(2) Using Definition (11.119) find the Fourier transform of e-ac2 for a> 0.
Compare your answer with that in the table of Fourier transforms in
Appendix F.
(3) Using Definition (11.119) find the Fourier transform of
1, ff(x)
_
0, f
et sin x, IxI<ir
u(x, t) _
0, otherwise.
Section 11.5: Fourier Transform Solution Method of PDEs 451
sin
et
J_lr
sine- Ze
d = 2ie t e2 -
1
sin (ir0)
Thus u (9, t ) = 2iet 82-1 .
If we assume Theorem (64) holds true for the Fourier transform of a PDE, which
au(x, t) a2u(x, t) a2u(x, t)
it does, then for , ate , and axe where - oo <x < oo and t > 0
at
we have the Fourier transforms;
f8u(x,t)
at } (11.121)
. f82u(x,t)
ate }
__ d2
dt2
u(8, t), (11.122)
and
a2(, t)
, ' ax2 = (i9)21i(9, t), (11.123)
The method for solving PDEs by Fourier transforms is quite similar to that of
Laplace transforms. We will use the notation 11(0, t) introduced in this section to
simplify the equations. Thus, we have the following steps:
(1) Given a PDE with infinite boundaries and initial conditions, take the Fourier
transform of the infinite boundary value problem while holding t as a para-
meter. You will get an ODE in the variable t, with function 11(0, t). Note:
The ODE only depends on the variable t.
(2) Solve the ODE and find 11(8, t).
(3) Take the inverse Fourier transform of u(B, t) to find u(x, t).
452 Chapter 11: Fourier Integrals and Transform Methods
EXAMPLE 11.15. Given a uniform infinite rod with perfect lateral insulation,
thermal diffusivity of 1, and initial temperature distribution of solve the fol-
lowing problem using Fourier transforms.
First, we must state the mathematical formulation of the problem. We have
au(x, t) _ a2u(x, t)
<x < oo and t > 0, (11.125)
at axe ' - °O
subject to
Step 1. Take the Fourier transform of the entire problem, which is from Equations
(11.121 and 11.123) and Equation (11.91),
du(9, t) = -9211(9,t),
(11.127)
dt
= (11.128)
1(9),
where the Fourier transform of the IC was found in Appendix F. Although the
Fourier transform of the IC is 1 e-92'4 the preferred form is f(9). This form
becomes important when we must do the inverse Fourier transform.
Step 2. Equation (11.127) is a first-order ODE in t. Its solution is
u(B, 0) = (11.130)
u(8, t) = e-e2/4e-92t
= j(9)e_02t. (11.131)
4r
Section 11.5: Fourier Transform Solution Method of PDEs 453
Step 3. Find the inverse Fourier transform of Equation (11.131), which is, by
Equation (11.92),
1001 e-e2/4e-e2teiex d8
2ir 1
°°
1
f (e)eJ-e2t eZee d8. (11.132)
(f * g)(x) =
f j(O)(O)ei0X dO.
(.f * 9)(x) = f 00
f(y)9(x - y) dye (11.134)
./
{_x2/4t} = t o-92t
= t o
t
Replacing x with x - y in Equation (11.134), we get the final answer of
1too
f (y)e2' dy
u(x,0) = f(x),
our solution is
Our next example will make use of Equation (11.136) and the error function,
which was introduced in Section 11.3, Laplace transforms.
EXAMPLE 11.16. Consider a uniform infinite rod with perfect lateral insulation,
thermal diffusivity of iron, and initial temperature distribution of 30°C for x <1
and 0 otherwise. Solve the problem using Fourier transforms and the error function.
First, we must state the mathematical formulation of the problem. We have
au(x, t) _ a2u(x, t)
-. 22 axe - oo < x < oo and t > 0, (11.137)
subject to
(30, x<1 1
u(x,0) = f(x) = (11.138)
0, x>1.1.
Using Equation (11.136),
15 f1
e y (11.139)
22t
r.22t f_i
Section 11.5: Fourier Transform Solution Method of PDEs 455
u(x, t)
15 2x-1ti e_z2
-2 .22t dz
r.22t 2 vi x+1
x+1
30 2 vi 2
Ie_Z dz
2 .22t
x+1 x-1
30
o
2 .22t
e-z dz
2
- o
2 .22t
e-z2 dz
15 ( erf
x+1
erf
[x_1l\ (11.140)
L2vi
.22t 2 .22t
Figure (11.19) graphs the temperature distribution, u (x, t) for several values of t. U
t=( 2
r = 1.6
15
t=3.2
\ 10
y 5
subject to
1
' 1+x2
au(x, 0)
at
=0 (11.142)
Step 1. Find the Fourier transform of Equations (11.141 and 11.142). We have
d2u(B,t)
_ -92 (9,t), (11.143)
dt2
subject to
.F{u(x, o)} = u(e, o)
.l
and
J aunt, o 1 _ (9). (11.145)
1
Step 2. Solve the ODE in t subject to the ICs, f(9) and (9). Equation (11.143)
is a second-order ODE that has the solution
u(e, t) = a(e)cos(et) + B(e)sin(9t). (11.146)
Applying the ICs to Equation (11.146), we arrive at A(B) = f(9) and B(B) = 0.
Therefore, Equation (11.146) becomes
u(B, t) = f(9) cos(Bt) _ ire-0 cos(9t). (11.147)
= - Looe_0cos(0t00* (11.148)
EXERCISES 11.5
11.5.1. Solve the following heat problems using Fourier transform method:
(1)
t) _ 82ua( 2, t)
-Do <x < oo and t > 0, subject to
u(x,0) = f(x) =
(2)
t> _ t)
<x < oo and t > 0, subject to
-25, -1<x<0
u(x,0) = f(x) =
25, 0,x<1
0, 1<x.
au(x, t) _ a2u(x, t)
(3 ax2 ' -oo < x < oo and t > 0, subject to
at
0, x<-1
u(x,0)=f(x)= x, -1<x<1
0, 1<x.
au(x, t) _ 4a2u(x, t)
(9) -Do < x < oo and t> 0, subject to
at 8x2 '
u(x,0) = f(x) =
a2u(x,t) = a2u(x,t)
(2) - < <, >o,
sin(irx) au(x, 0)
subject to u(x, 0) = and =0.
subject to
11.5.4. Solve the diffusion equation with convection. That is, solve
u(x,0) = f(x).
au(x, t) a3u(x, t)
at C7x3
' - 00 <x < 00,
subject to
u(x,0) _
0, IxI>2.
Section 11.5: Fourier Transform Solution Method of PDEs 459
11.5.7. Consider
a2u(x, t) a4u(x, t
t2 - 00 <X <00,
8x4
subject to
1
u(x, 0) _
1+x2
(1) Find the Fourier transform of the boundary value problem.
(2) Solve the resulting ODE.
(3) Justify a reasonable assumption about the boundedness of the solution
of the ODE.
(4) Apply the IC.
(5) Give u(x, t) in the form of an inverse Fourier transform form.
11.5.8. Solve
82u(x, t) x,t) - oo <X <00
ax4
8t2 -a 8t82x -b
subject to
0)
u(x, 0) = f (x) and aunt' = g(x)
for u(x, t) and leave the answer in inverse Fourier transform form.
11.5.9. Given
au(x, t) _ a2u(x, t)
at
ko x2 - oo < x < oo and t > 0,
subject to
ul/x OJ ` _ fl/xJ ` _ e-amt
11.5.10. Solve the following variable-coefficient PDEs using the Fourier transform method:
8u(x, t) _ t8u(x, t)
(3) , subject to u(x, 0) =sin x2.
at ax
au(t,t) auax,t)
- Sin t , subject to u(x, 0) = sin x2.
au(x, t)
= cost
au(x t
lU/ , subject to u(x , 0) = cos x2 .
at ax
ki
au(x,t) - _tau(x,t)
at axe
subject to u(x,0) = 256.
11.5.11. PROJECT: Solve the Klein-Gordon equation in one dimension. That is,
solve
8 u x, t)
a 2 - 2a
ax2,t)
- m2u(x, t), - Do <x < oo,
subject to
0)
u(x, 0) = 0 and
au(x,
8t
Note: The Klein-Gordon equation is one of a number of equations which
model elementary particles; electrons, mesons, quarks, etc. where m is the
mass of the particle
a 2u(x, t) 2 x,t
(1) Find the Fourier transforms of u( x, t) , and the ICs.
ate '
Note: The Fourier transform of u(x, t) is 11(8, t).
sin tc2B2 + m2
(2) Solve the resulting ODE and apply the ICs. Hint: 11(9, t) _
c2B2 + m2
(3) Setup the inverse Fourier Transform.
(4) From the theory of cylindrical functions it is known that
sin r 1
Jo (r sin sin T) ear cos cos T
sin T dT
r 2 o
Section 11.5: Fourier Transform Solution Method of PDEs 461
(6) After working this part of the problem, you should arrive at the following
answer :
2
u(x,t) = 2c Jo m t2 - 2 for x < ct.
subject to
463
464 Appendix A: Summary of the Spatial Problem
Given
-Ao(x),
8x2
we have the following results:
(flit z
n=1, 2, 3,...
CA
(2n- 1)
2L (2n - 1) 7rx
cpn (x) =sin
2L
n = 1,2,3,...
ag(o) = o ((2n_1)ir
2L 2n- 1) 7rx
pn(x) = cos
2L
pSp(L) = 0 n=1,2,3,.
Ao = 0 p0(x) _ Co
( z
=
L)
0p(L)
(L)
Ox
- f0
= n = 1, 2, 3,. p(x) = cos
L
p0(x) _ Co
( p(-L)=cp(L)
(flit 2
cosnitx
L
&p(-L) &p(L)
L)
ax ax n=1,2,3,. pn(x) =sinnix
L
Appendix B
do
f f (x, t) dx
a = Ifa f att dx.
Since
a f (x'
at
t) is continuous, we know that g() --
t is continuous for c < t < d. Thus
465
466 Chapter 2: Proofs of Related Theorems
Therefore, by the integral interchange theorem, we may switch the order of integra-
tion in Equation (B.1). We have
d d b of (x, t) b cd
of (x, t)
g(t) dt = dx dt = dt dx
I c a at a at
b b b
[f(x,d)-f(x,c)] dx = f (x, d) dx - f (x, c) dx
a a a
= F(d) - F(c),
where F(d) and F(c) are integrations depending on a parameter. If we let d be a
variable t, we have
t
F(t) - F(c) = g(u) du. (B.2)
c
Equation (B.2) can now be differentiated with respect to t. Hence, by the funda-
mental theorem of integral calculus we have
fb af(x,t)
F'(t) = g(t) = dx.
Ia at
Note:
We want to show that the maximum-minimum value occurs on the sides of the
rectangle, R, given by
0<x<Land0<t<T
as shown in Figure B.1. Since k is a positive constant in Equation (B.6), we will
0 L
Figure B.1: u(x, t) in the rectangle 0 < x < L and 0 < t < T.
(XO, to), where 0 < xo <L and 0 < to, be a point where u(x, t) attains minimum in
R. Thus, u(xo, to) = m. Consider the function
(x,t) = u(x, t) - 4L2 (x - xo)2 . (B.8)
Note:
aµ(x, t) 8u(x, t
at - at
(B.9)
and
a2µ(x,t) _ a2u(x,t) _ E
.
axe axe 2L2
On the lower base and vertical sides of R,
x t >m+E- E =m+-4
(B.11)
82µ(xl,t
8x2
Hence, at (x1, t1),
aµ(xi, ti) a2µ(xi, ti) o.
at 8x2
However,
aµ(xi, ti) a2µ(xi, ti) aµ(x, t)
at - axe - at
au(x, t) a2µ(x, t)
a2u(x, t) E
= axe + 2L2 > 0.
Since
C1 (A 1 - Al 1) = C1 (\1 1 - \1 1)
by Definition (15), we have
c2(Ai-A2)=...=c(Ai-A2).
But each ai - A3 0. Hence, c2 = c3 = ... = cn = 0. Thus, Equation (B.12)
becomes
cl x1 = 0 ,
eigenvectors are linearly independent. We must show that they span W'. Let y
be any vector in W'. Then, the set of vectors {T' x2, x3, ... n , y } is linearly
dependent because each vector is a linear combination of any basis vectors of W'.
Thus, the constants c1, c2, c3,... , cn, and c exist where at least one is not 0, such
that
c1 x 1 -}- c2 x2 -}- c3 x3 -}- ... -}- cn n -}- c y = O. (B.13)
Since xi , x2, x3, ... , n are linearly independent, we must have c 0 in Equation
(B.13). Therefore,
c1xxi+c2x2+c3x3+...+cn n=-cy.
-4 -4 -4
Since y was an arbitrary vector of W', x1, x2 x3,... n span Rn . Hence,
2 22
, , , n is a basis of R.
Theorem 70. (Maximum Principle) Let 1 be a bounded set in l2. Let u(x, y) be a
harmonic function in 1, while u(x, y) must be continuous in the union of 1 and the
boundary of 1, denoted 91k. Then, the maximum and minimum values of u(x, y)
are attained on the aSZ, unless u(x, y) is identically equal to a constant.
Proof. Let E > 0 be given. Let µ(x, y) = u(x, y) - E (x2 + y2). Then,
a2µ(x, y)
ax2 +
a2µ(x, y)
ay2
_
_
a2u(x, y)
ax2 +
a2u(x, y)
aye
- 4E=0-rE<0.
But a2µ(x,y)
axe + a2µ(x'y) > 0 at an interior minimum point, by the second
- deriv_
a y2
ative test from calculus. Thus, µ(x, y) has no interior minimum in f Since µ(x, y)
is a continuous function it must have a minimum in SZ U as Let the minimum of
µ(x, y) be attained at (xO, yo) on %. Then, for all (x, y) in 1 we have
u(x, y) > µ(x, y) > µ(xo, yo) = u(xo, yo) + (x2 + y2) > min u + Er2,
asp
where r is the radius of the circle centered at the origin that contains SZ U of Since
this is true for all f > 0, we have
u (x, y) > min u
asp
If F(t) = 0, then the equation is homogeneous. For the homogeneous case, u(t) - 0
(where - means identically equal to) is the trivial solution. For a solution to exist,
it is important that S(t), K(t), and F(t) be continuous functions on some interval
a <t <b where the ODE is defined.
If S(t) is not equal to 0 anywhere on the interval a <t <b, then Equation (C.1)
may be written in the more familiar form as Equation (C.2),
Equation (C.2) is said to be linear since u'(t) and u(t) are linear. Therefore, let
us first consider Equation (C.2) where k(t) is a constant and then complete this
section by discussing Equation (C.2) where k(t) is a more interesting function.
473
474 Appendix C: Basics from Ordinary Differential Equations
where A is a constant. Note: In Equation (C.3), the function, 1(t), is the constant
0. Assuming u (t) 0 for any t, we can rewrite Equation (C.3) as
(c.4)
u(t) _ Mt.
ln[u(t)] _ At + a.
Now remembering some basic calculus, we arrive at the more compact and useful
solution to Equation (C.4), which is
u(t) =cert.
EXAMPLE C.2. Consider
u'(t) + au(t) = A. (C.5)
Note: In Equation (C.5), the function f (t) is the constant A, and a is a nonzero
multiplicative constant of u(t). We solve Equation (C.5) using the usual technique.
That is, we find an integrating factor. Therefore, we must determine a function g(t)
such that when you multiply Equation (C.5) by g(t), the left side is recognized as a
derivative of a product. Many times you can quickly discover the function g(t) by
remembering your calculus. In this case, if we choose g(t) = eat, we can form the
product eatu(t). When we then take the derivative of this product, we have
dt
[eatu(t)] - eatul(t) + ae°'tu(t) = eat [+ au(t)]. (C.6)
The left side of Equation (C.5) and the contents of the brackets on the right side of
Equation (C.6) are equal. This means that we have the correct integrating factor.
Thus, we can multiply both sides of Equation (C.5) by eat and get
resulting in
[u(t)eat]I = Aeat.
A
u(t) = ce-at
+ -. U
a
u(t)
In most applications, the solution u(t) is known for some value of t = to where
u (to) = no. This is generally called an initial condition (IC). The IC actually
describes a point (to, u0) through which the solution curve passes, resulting in a
specific solution. For example, if in u(t) = ce-at + - (Example C.2) we had A = 1
a
a = 1 and an IC of n(O) = 1 then our solution would become n(t) = -e 2t + 2.
This is graphed in Figure (C.2).
u(t)
2F
-4 2 4
-2
-4
-6
-8
-10
EXAMPLE C.3. Consider the following ODE subject to the stated IC:
u'(t) - 2tu(t) = t, u(0) = 1. (C.7)
There are three ways of solving this ODE. First, we could use separation of variables
and solve the related problem
u'(t)
1 + 2u(t) t
A second method is to treat the original problem as nonhomogeneous. In this case,
we solve the homogeneous equation uh (t) and then solve for a particular solution
up (t) . Adding uh (t) and up (t) yields the solution for u (t) . Although this method
works, I do not recommend it, because we solve the same problem twice and do
twice the amount of work.
The third method is the integrating factor method. In Equation (C.7), we have
k(t) = -2t, which suggests that an appropriate integrating factor is f (t) = e-t2.
To prove this, consider
d
dt
e-t2u(t) = e-t2u'(t) - 2te-t2u(t) = e-t2 [u'(t) - 2tu(t)].
which becomes
[e_t2u(t)] = to-t2.
u(t)e_t2 l _t2
Section Cl: .Some Solution Methods for First-Order ODES 477
the general solution. Using the given IC u(0) = 1, we find c = 2. Therefore, the
specific solution of Equation (C.7) is
3 t2 1
u(t) _ 2e -,
which is graphed in Figure (C.3).
u(t)
1.4
1.2
0.8
0.6
0.4
0.2
-
t
0.1 0.2 0.3 0.4 0.5
3
Figure C.3: The graph of u(t) = 2 et2 - 12 for 0 < t < 0.5. --
The ability to solve first-order ODES is an interesting mental exercise, but the
real benefit comes with applying the ability. Applications occur in engineering,
biology, chemistry, environmental studies, and many other disciplines. Later, we
give an example of an application in elementary mechanics.
In elementary mechanics, we assume motion of a rigid body is along a straight
line. Thus, Newton's law of motion applies, giving us "mass times acceleration
equals the sum of the external forces." In mathematical terms, this translates to
F=ma.
Knowing this fact, we proceed to our example.
EXAMPLE C.4. Suppose a body of mass 10 kgs is projected upward with an
initial velocity of 100 m/s. If we assume the gravitational attraction of the earth
is constant, and we neglect all other forces acting on the body, we can find the
maximum height attained by the body, the time at which the maximum height is
reached, and the time it takes the body to return to its starting point.
Solution: We begin by assuming the positive direction is upward and the origin is
the surface of the earth. We have, from Newton's law, F = ma, (force equals the
478 Appendix C: Basics from Ordinary Differential Equations
mass of the object times acceleration due to gravity), (lOkg)(-9.8 m/s2). Force
dv
also equals the mass of the object times acceleration, which is (10 k9) . Thus we
dt
must solve
(-9'8 /S2)t2
x(t) _ + 100t m/s.
2
Note: Since initial distance is at the surface of the earth, implying x(0) = 0, the
constant of integration c equals 0. Applying t = 100 s' Yields
g.8
1002
xmax = 2(9.8) m.
To answer the third question, the time it takes the body to return to its starting
point, we multiply the time it takes to reach maximum height by two, giving us
200 s
2t= 9.8
As the above example shows, using basic laws to acquire an answer to a problem
often results in solving a first-order ODE. This usually means that the problem is de-
scribed by a first-order ODE. For instance, consider problems describing radioactive
decay. Here the ODE is Q'(t) = kQ(t), where k is the constant of proportionality.
Other examples of problems that are described by first-order ODEs are determining
Section C.2: Some Solution Methods for Second-Order ODES 479
Solutions for Equation (C.9) depend on the nature of k(t), h(t), and f (t); I will
limit this section to solutions of second-order linear-constant coefficient (k(t) and
h(t) are constants) ODEs,
we assume that the solution is of the form u (t) = &'t. Finding both the first and
second derivatives of u(t) and plugging them into Equation (C.11) yields
r2+br+c=0. (C.13)
480 Appendix C: Basics from Ordinary Differential Equations
-b+(b2-4(1)(c))2 -b+(b2-4c)2
rl =
2(1) 2
and
The discriminant (b2 - 4c) determines the solution type. We have three possible
solution types:
real and unequal roots: b2 - 4c > 0;
real and equal roots: b2 - 4c = 0; and
complex roots: b2 - 4c < 0.
However, there are other ways of doing the linear combination. For instance,
consider the form of rl and r2 wherein
rl _ -b + (b2 - 4c) 2 _ _b + (b2_4c) 2
2 2 2
and
r2= 2 =2- 2
or
and
ear - e-ate
sinh ax =
2
called the hyperbolic sine and hyperbolic cosine functions. Doing so allows us to
form the new linear combination
t (b2_4c)12
u(t) = e 2 a cosh yt + /3 sinh 7t]; = 2 ,
in which a and /3 are arbitrary constants. We use this form of the solution to
Equation (C.11) with real roots, because it is much more advantageous in our study
of PDEs. The advantage is discussed in both Chapters 4 and 5. Note: If the term
b2 - 4c, is a perfect square, then we will not have the hyperbolic sine and hyperbolic
cosine as part of the solution.
r2+3r+1 =0.
The quadratic formula yields the following roots:
-3+(9-4) 2 -3+v"
rl 2 2
and
Letting rj = 2 and = -3
2 , r1 = + rj and r2 = - rj. This means that the
solution is
or
Note: This is the form of the solution only for ODEs with constant coefficients.
EXAMPLE C.6. Solve
u" (t) + 4u' (t) + 4u (t) = 0.
r2 +4r+4=0,
which factors into
Remember, our second solution depends on using the reduction of order method
and the ODE having constant coefficients. Combining the two solutions yields
u(t) = ae_2t + /3te_2t.
Complex Roots
For complex roots, both r1 and r2 are complex numbers. That is, each root has the
form A + Bi, where A is the real and Bi is the complex part. Note: A and B are
real constants, thus, the roots to the characteristic equation, Equation (C.12) are
= -b -24c 2
Ib2
-b + Ib2 - 4c 2 i
r1 = 2 2+ i
Section C.2: Some Solution Methods for Second-Order ODES 483
and
Thus, we know the two solutions of Equation (C.11) with complex roots are
_7it Ib2
_ 2t 7it _ 2t - 4C) 2
u1(t) = e e and u2 (t) = e e where 'y = 2
_ bt Ib2 - 4C) 2
u(t) = cie 2 e7it + c2e - 2t e -7it , where 'y = 2 (C.16)
However, there are other ways of doing the linear combination. For instance, you
could do the linear combination as
bt d1
e7it + d2 e-lit I b2 - 4cl 2
u (t) = e- 2 2 where 'y = 2
or
bt
[mie7it - m2 e-lit I b2 - 4cI 2
u(t) = e- 2 where 'y =
2i 2
and
eiax - e-iax
sin ax =
2i
- bt b2 - 4cI 2
u(t) = e 2 [a cos'yt + /3 sin'yt] ; 'y =
2 '
in which a and /3 are arbitrary constants. We will use this form of the solution to
Equation (C.11) with complex roots, because it is much more advantageous in our
study of PDEs. Again, the advantage is described in Chapters 4 and 5.
EXAMPLE C.7. Solve
u"(t) + 6u'(t) + lou(t) = 0.
Assuming u(t) = ert, the characteristic equation is
r2 + 6r + 10 = 0,
484 Appendix C: Basics from Ordinary Differential Equations
r2 = -3-i.
These roots yield the general solution
u(t) = e-3t [a cos (t) + 3 sin (t)]. U
We have covered all three possible general solutions to Equation (C.11). For a
specific solution, we must determine the coefficients. Since the basic idea behind
the solution to a linear homogeneous second-order ODE is two integrations, you
can see that we have two constants of integration to solve for. These constants
of integration have become, through various manipulations, the coefficients of our
general solution (a and /3 in the previous). Thus, a well-defined solution needs two
initial conditions. If the ODE has time, t, as an independent variable, then the
first IC is usually the initial position of the physical body you are modeling. The
second IC is generally the initial velocity of the body and, as such, is the derivative
of the function describing the position of the body. For instance, u (to) = a and
du(to)
= b. Notice that a and b are constants and not functions.
dt
Another way of describing the beginning state of a body with a second-order
ODE without using initial conditions is to describe its boundaries. In this case, the
problem is called a boundary value problem (BVP). Boundary conditions (BCs) are
generally given when the equation is restricted to an interval and the endpoints are
known. Like initial conditions, we need two BCs. Examples of BCs for 0 <t < L
are the following:
u(0) = a, u(L) = b;
u' (0) = a, u' (L) = b;
u(0) = a, u'(L) = b;
or
u'(O) = a, u(L) = b.
EXAMPLE C.8. Consider a weight of 16 lbs attached to a steel spring that has
a natural length of 2 feet. The mass stretches the spring 0.1 ft. Suppose the system
is started in motion by stretching the spring an additional 0.25 ft downward; then
it's released. Determine and solve the resulting equation of motion, neglecting air
resistance.
Solution: Our basic equation is derived from Newton's law of motion, F = ma.
If we let u(t) represent the displacement of the mass from equilibrium, we have
d2u weight 16 lbs 2
ma = m where m = = = 0.5 lbs-s /ft. Though we have no
dt2 ' gravity 32 (ft/s2)
external driving force, we do have a force due to the spring. This force we call FS ,
and it always acts to restore the spring to equilibrium. The force FS is proportional
to the displacement of the spring, u; this is called Hooke's law. FS = ku, where k is
the spring constant. We determine the "spring modulus" k by dividing the weight
16 lbs = 160-j.s
by
y the distance the weight stretches the spring. This gives k =
0.1 ft
Putting it all together results in
d2u
0.5 -160u
dt2
lbs-s2/ft lbs/ft
or
d2u
dt2 + 320u/s2 = 0,
which has the solution
where uh (t) refers to the homogeneous part of the nonhomogeneous ODE. We dis-
covered in the previous section that the solution to the homogeneous equation is
where uhl (t) and uh2 (t) are two linearly independent solutions. The particular
part of Equation (C.17) really refers to any solution, up (t), of the nonhomogeneous
equation, Equation (C.17), which is different from uhl (t) and uh2 (t). Thus, the
complete solution to Equation (C.17) is
where uh (t) refers to the homogeneous solution. Assuming uh (t) = ert, we form the
characteristic equation
r2 + 3r + 2 = 0,
which factors into
Equation (C.20) has real and unequal roots of r = -2 and r = -1. Therefore, we
know that uhl (t) = e-2t and uh2 (t) = e-t. Thus,
(Note: Equation (C.21) is not in the form of hyperbolic sine and hyperbolic cosine
functions because b2 - 4c is a perfect square).
We must now determine the particular solution, up (t) . First, we notice that 1(t) is
a polynomial in the independent variable t, 1(t) = 3t2 + 2t - 1. This suggests that
a polynomial may be the form of the particular solution, up(t). Thus, we let
We then find the first and second derivatives of Equation (C.22), which are
and
Then, we replace u' (t) with up (t) and u" (t) with u , (t) in Equation (C.18), resulting
in
2A + 3B + 2C = -1, (C.26)
6A + 2B = 2 (C.27)
and
2A = 3. (C.28)
3
From Equation 2 Thus, by replacing A by 2 in Equation
(C.28), we know A = 3.
C = 13. Therefore
4 '
u(t)=t2_t+
7 13
4 (C.29)
and
Replacing u" (t) with u , (t), u' (t) with up (t), and u(t) with up (t) in Equation (C.30)
yields
We solve for the constants A and B in Equation (C.33). By matching the coefficients
of similar terms of the function sin t and cos t on both sides of the equation, we have
-B-3A+2B=B-3A=1 (C.34)
and
-A + 3B + 2A = A + 3B =0. (C.35)
-3
Solving Equations (C.35 and C.34) simultaneously yields A = and B = 1 .
10 10
Therefore, the solution to Equation (C.30) is
u(t) =u(t)
p +uh(t) =-3cost + 1
sint+c1e-2t+c2e-t.
10 10
Note: Using up(t) = A sin t would not have given you the correct answer for u(t).
This completes our review of ODEs. For an extensive review of this topic, I
suggest Elementary Differential Equations, by Boyce and DiPrima.
Appendix D
Mathematical Notation
Throughout this text, we use mathematical notation that may be new to you,
such as IIg to represent the real number line, commonly called the Reals. Also, in
your study of mathematics, you may have thought mathematicians are notation
happy. However, this is not the case. Notation, and knowing the correct definition
of notation, is very important. Another example that shows the importance of
notation is
f
1
x dx.
In English, this notation says integrate the function x on the interval 0 to 1 with
respect to x. I think the notation is much clearer than the English version and
certainly easier to write. I hope you agree. Thus, you easily see it is important
in your mathematical education to learn some basic notation, particularly notation
found in many text books, such as names of spaces and mapping notation.
The real number line is known by the symbol IIg and is called one-space. The
natural numbers, a subset of the real numbers, use the symbol N. The xy-coordinate
plane is commonly known as IIg2 and is called two-space. Three-dimensional space,
the xyz-coordinate system, is known as IIg3 and is called three-space. Both
and IIg3 are easily identified. They are formed by taking copies of the real number
line, then intersecting them at right angles to each other. Since we live in three
dimensional space, it is hard to imagine what 1R5, or R where n E N, meaning
n is any natural number, look like. But mathematically they exist, and they are
quite useful. Also, the spaces that are formed by copies of the real number line,
such as IIg2, IIg3 or R, are known as finite-dimensional spaces. One reason they are
considered finite is the number in the superscript; it indicates a finite number of
copies of the real number line. The reason for the superscript (hence, why they are
finite dimensional) will be addressed in the next section.
The other notation that is important to know is mapping notation. Mapping
notation is very important. It is closely tied to functions, and is easily understood.
For instance, when you add the number 2 to the number 3, 2 + 3, you are actually
dealing with a function. The function is addition. It would be silly to write mapping
489
490 Appendix D: Mathematical Notation
notation for the addition of 2 and 3. However, if you wanted to indicate addition
of all real numbers to each other, then mapping notation is much easier and more
compact. The mapping notation indicating that you can add all real numbers to
each other is
+:IIgxIIg-*II.
It says that you are adding a real number from one copy of IIg to a real number from
another copy of Ilk, and you expect the answer to be a third copy of JIB.
Generally, we do not use mapping notation for addition. We reserve it for other
things. For example, suppose you wanted to discuss, in general terms, all the
functions that are in Mapping notation is a great help. The notation would
simply be
f : IIg -* III.
Notice the mapping has two single copies of III. Adding their superscripts to-
gether indicates the graphs of the functions are living in two-space. What would
the mapping look like for functions in II? One might naturally imagine it is
f: li x li -* IIg.
This indicates functions that take a point from the xy-plane and map it to a
value on the z-axis giving us graphs in the normal IIg3 space. But we could have a
mapping which looks like
This mapping indicates functions which take a value on the z-axis to a point on
the xy-plane. It is another copy of but with a slightly different twist. The
important thing is knowing where the mapping originates and where it terminates.
Appendix E
Summary of Thermal
Diffusivity of Common
Materials
In this appendix, a summary of thermal diffusivity, k, for common materials is
given. Remember, thermal diffusivity is given by
k = K°
cp
where K° is the thermal conductivity, c is the specific heat, and p is the mass density
of the given material.
Thermal conductivity, K°, is considered temperature independent provided the
temperature variation is not too great. If the temperature variation is great enough,
then the formula for calculating the thermal conductivity is given by
1 Ta
Ki°
=T _T K°dx,
2 1 Tl
where Ko is the new thermal conductivity. However, for most problems in this text,
we will assume that thermal conductivity is temperature independent.
491
492 Appendix E: Summary of Thermal Diffusivity of Common Materials
Metals:
Non-metals:
493
494 Appendix F: Tables of Fourier and Laplace Transforms
1, x< aa
1 -1 sin a9
(
9
)
0, a < Ix
Ire-a9
1
2 x2 + a2 , a > a
ex, a>O,x>0 1
3. J
0, otherwise a + iO
ex, a>0,x<0 1
4.
0, otherwise a - i9
1
5. a-ax 2 , a > 0 e_02/4a
ra
2a
6. a-aIxI, a> 0
a2 + 02
ire-a0
7. x ,a>0
x2 -}- a2
2a
8.
x2 + a2
, a>0 e -alel
Section F.1: Table of Fourier Transforms 495
7r, 9<a
sin(ax) 0
9
x
a> ?r 8= a
0, >a
?r 82 ?r
11. sin(ax2 ), a > 0 cos + iJ
a 4a
1
1, 0<x<a sin(a8)
J
' e
lo, a<x
ire-a6
a
2. a2a>
x2 -}- 2a
a
3. a-ax, a > 0
a2 + 82
496 Appendix F: Tables of Fourier and Laplace Transforms
a2 - e2
4. xe-ax, a> 0
(a2
+ e2)2
7. -ax a> 0
cos(ax)e, ae2 + 2a3
4a4 -}.., e4
8. - a>0
sin(ax)eax, 2a3 - a612
4a4 -}.., e4
sin ax
ir/2, e<a
10. ,a>0 7r/4, e=a
x 0, e > a
e2
12. cos(ax2 ), a>0 cos e2 + sin
8a 4a 4a
Section F.1: Table of Fourier Transforms 497
1, 0<x<a 1 - cos(ae)
1.
e
0, a < x
x 7I'e-ae
2 x2 a> 2
+a 2,
a-ax , a > 0 e
3. a2 + e2
tae
4. xe-ax, a> 0 (a2
+ e2 )2
xe-ax2 ee_ea/4a
5. a 0
4a3/2
' >
e3
7. cos(ax)e-aX, a> 0
4a4 -}.., e4
2a2e
8. sin(ax)e -a x , a > 0
4a4 -}.., e4
sin ax
10. a> 0 1 In f B+ a 1
x 2 9-a]
0, e < a
cos(ax)
11. ,a>O ir/4, e=a
x ir/2, B>a
x
12.
ex
,a>0 -
tan1
\/a
1. 1 -s ,s>0
1
1
2. t 2
s
3. t,nEN n!n,s>0
S
1 1
4.
art
Section F.2: Table of Laplace Transforms 499
1
5. a«t
s-a , s > a
a
6. sin(at) ,s>0
s2 a2
F(p+l)
p sp+1
S
8. cos(at) ,s>0
s2 a2
a
9. sinh(at) s2 - a2 , s > a)
s
10. cosh(at) s2 - a2 , s > Ia
s- a
12. a«t cos(,3t) , s> a
(s-a)2 +132
n.
13. the«t, n E N , s> a
(s - a)n+1
e-«S
14. u(t - a), unit step function , s > Ox
s
t
18. f (t - u)g(u) du, Convolution Integral F(s)G(s)
0
1 -a2 e-a
19. a 4t
art
20.
a
e 4t
2
e-a
2 't3
21. er f c
2 a
, Complementary Error Function
e
s
t -a2 a
22. 2
VT_
a- a er f c
24.
2
-eabeb t {erfc b -}-
a
+ er f c
(a) be-a1
2) J
e-as
25. H (t - a)
S
Bibliography
1. Asmar, Nakhle H. Partial Differential Equations and Boundary Value Prob-
lems. Englewood Cliffs: Prentice-Hall, 2000.
2. Bartle, Robert G. The Elements of Real Analysis, 2nd ed. New York: Wiley,
1976.
501
502 Bibliography
15. Hayt, William H. Jr. Engineering Electromagnetics, 5th ed. New York,
McGraw-Hill, 1989.
16. Keener, James and Sneyd, James Mathematical Physiology. New York: Springer-
Verlag, 1998.
35. Walker, P. L. The Theory of Fourier Series and Integrals. New York: Wiley-
Interscience, 1986.
36. Wolfram, Stephen The Mathematica Book, 3rd ed. Champaign, IL.: Wolfram
Media and Cambridge University Press, 1996.
37. Wylie, Ray C. and Barrett, Louis C. Advanced Engineering Mathematics, 6th
ed. New York: McGraw-Hill, 1995.
38. Zachmanoglou, E. C. and Thoe, Dale W. Introduction to Partial Differential
Equations with Applications. New York: Dover, 1986.
39. Zill, Dennis G. and Cullen, Michael R. Advanced Engineering Mathematics.
Sudbury: Jones and Barlett, 2000.
Index
505
506 Index