Differential Equation
Differential Equation
Larry C. Andrews |
USED
Larry C. Andrews
University of Central Florida
Andrews, Larry C.
Ordinary differential equations, with applications.
ISBN: 0-673-15800-4
1234567 8-RRC-88 87 86 85 84 83 82 81
Preface, ix
Basic Concepts, 1
1.1 Introduction, 2
1.2 Classification of DEs, 2
1.2.1 Origin and Application of DEs, 4
1.3 Solutions of DEs, 6
1.3.1 General Solutions, 9
1.4 Initial and Boundary Value Problems, 14
References, 17
2.1 Introduction, 19
2.2 Separation of Variables, 20
2.3 Miscellaneous Techniques, 25
2.3.1 Exact Equations, 25
2.3.2 Integrating Factors, 28
2.3.3 Linear Equations, 30
2.3.4 Homogeneous Equations, 31
*2.4 The Theory of Linear First-Order Equations, 36
2.4.1 The Homogeneous Equation, 37
*2.5 The Nonhomogeneous Linear Equation, 41
2.5.1 Physical Interpretations, 45
2.5.2 Existence-Uniqueness Theorem, 47
*2.6 Bernoulli’s Equation, 52
References, 54
*Sections marked with an asterisk may easily be omitted for a shorter course.
vi CONTENTS
3.1 Introduction, 56
3.2 General Applications, 56
3.2.1 Orthogonal Trajectories, 56
3.2.2 Free-Falling Bodies, 59
3.2.3 Flow of Water Through an Orifice, 62
3.2.4 Curves of Linear Pursuit, 63
3.3. Applications Involving Linear Equations, 69
3.3.1 Growth and Decay Problems, 70
3.3.2 Motion of a Particle and Simple Electric Circuits, 73
3.3.3 Cooling and Mixing Problems, 76
References, 81
4.1 Introduction, 83
4.2 Linear Dependence and Independence, 84
4.2.1 Wronskians, 86
4.3 Constructing a Second Solution from a Known Solution, 91
>» 4.4 Homogeneous Second-Order Equations with Constant Coefficients, 94
4.4.1 Differential Operators, 98
4.5 Higher-Order Linear Equations, 101
4.5.1 Linear Independence, 102
4.5.2 Homogeneous Constant-Coefficient Equations, 104
4.6 The Nonhomogeneous Equation, 108
4.6.1 The Method of Undetermined Coefficients, 109
4.7 Variation of Parameters, 115
4.8 The Cauchy-Euler Equation, 121
References, 124
index, 335
This book is intended as a one-semester, one-quarter, or two-quarter introductory
course in ordinary differential equations and their applications. It is designed for
students in mathematics, engineering, or science who have already successfully
completed a basic course in calculus.
Applications in the physical sciences as well as some in the social and life
sciences are prominent throughout the text, but kept at a fairly elementary level so
that little background in the various sciences is required to understand them. And
although applications are discussed throughout the text, I have attempted to main-
tain a close relationship between mathematical theories and applications. In this
regard, I have tried to avoid the temptation of introducing a multitude of applica-
tions in many diverse areas of application. My experience is that exposing the
student to too many applications proves distracting, at least to the average student.
(More of something good is not always better.)
For the most part, the text contains the standard material that is usually found
in introductory one-term courses. There are, however, some distinctions in the text
that are perhaps noteworthy. For example, in addition to the integrating factor
technique for solving first-order linear equations, a separate discussion of the
general theory of these differential equations is presented (in Chapter 2) from the
same point of view that is used in Chapter 4 in developing the theory of higher-order
linear equations. Thus, not only is a more unified approach to the theory of all linear
equations possible, but the simpler theory for the first-order equation can be used
for motivation in developing the corresponding theory for higher-order equations.
Another feature of the text is the introduction of Green’s functions (Chapter 5)
for handling nonhomogeneous equations in a systematic and physically meaningful
fashion. I believe an early exposure to Green’s functions involving ordinary differ-
ential equations enhances the student’s understanding and use of these functions
when required in more advanced courses involving partial differential equations.
A brief discussion of the qualitative methods used in oscillation theory is pre-
sented in the chapter on initial value problems (Chapter 5). Because many students
take differential equations prior to a course in linear algebra or matrix theory, an
elementary (nonmatrix) treatment of linear systems of differential equations is
presented in the early sections of Chapter 7. Hence, using techniques already
familiar to the student, a natural transition from higher-order linear equations to
ix
PREFACE
linear systems of equations can be attained. A separate section on the use of matrix
methods in solving linear systems of differential equations is also included for those
students who are already versed in matrix techniques.
The text contains more than enough material for a one-semester course, and most
of the chapters after Chapter 4 are sufficiently independent of each other so that
various arrangements of topics can be made to suit individual needs. Also, sections
that can easily be omitted for a shorter course are marked [O]. Hence, the text
should be flexible enough to lend itself to several different types of courses. For
example, a course may consist of selected material from Chapters 1-5, 7, and 9,
or from Chapters 1-6, 8, and (possibly) 9. If applications are not stressed, a course
could be based upon Chapters 1, 2, 4, 6—9, and so forth.
With respect to the physical layout of the text, I have numbered sections and
subsections in decimal form. The subsections do not contain separate exercise sets.
Various equations, theorems, figures, and tables are numbered consecutively in
each chapter, but not according to section. For instance, Theorem 4.5 is the fifth
theorem in Chapter 4 without regards to any particular section. Since worked
examples usually play an important role in learning new material, I have included
a large number of them (over 170) of varying difficulty throughout the text. Each
example is generally indicative of typical problems to be found in the exercise sets
at the end of each section (which consist of more than 1100 problems). Further-
more, all examples are set apart from the textual discussion by the use of horizontal
lines in the hopes of making them easy for the student to find. The exercise sets
usually contain a blend of drill-like problems, some more difficult, and some that
extend the theory and applications beyond that discussed in the exposition. Prob-
lems that are considered to be more challenging and those used to extend the theory
are generally separated from the more routine or drill-like problems by placing them
either at the end of the exercise sets, or identifying them by a star (*), or both.
References are listed at the end of each chapter.
I am grateful to Gerald Bradley, William Fitzgibbons, John Gregory, Samuel
Rankin, Burton Rodin, and Chester Scott, who served as reviewers, and to Leslie
Rochlin for suggesting Scott, Foresman and Company to me. Also, I would like to
record my appreciation to the editorial and production staff of Scott, Foresman and
Company for their helpfulness and cooperation with this project. And finally, I wish
to thank Jack Pritchard, editorial vice-president, who became closely acquainted
with this project near its end, but who was very helpful in working out the final
details and seeing it through to completion.
L. C. Andrews
Differential equations play a fundamental role in engineering, mathematical sciences,
and the life sciences because they can be used in the formulation of many physical
laws and relations. The development of the theory of differential equations is closely
interlaced with the development of mathematics in general, and it is indeed difficult
to separate the two. In fact, most of the famous mathematicians from the time of
Newton and Leibniz had some part in the cultivation of this fascinating subject.
In developing the theory of differential equations systematically, it is helpful to
Classify the various types of differential equations, since equations in a particular
Class can often be solved by the same method. Therefore, in Section 1.2 we discuss
some of the various classification schemes, emphasizing order and linearity.
In Section 1.3 we carefully define what we mean by a solution of a differential
equation. Here we first discover that differential equations are peculiar in that they
generally possess many different solutions, and for this reason we usually seek a
specific function, called a general solution, with the property that all solutions can be
obtained from it.
We introduce the notions of initial value problems and boundary value problems
in the last section of the chapter. These are the names attached to those problems
occurring in applications wherein the solution of a differential equation must satisfy
certain additional auxiliary conditions. The study of these problems is so important in
applications that we have devoted a significant portion of the text to developing the
theory associated with them.
1.1 INTRODUCTION
The theory of differential equations (DEs) has played an important role in science
and engineering since the introduction of the calculus by Newton* and Leibniz.t
Problems in the physical sciences have long since been investigated primarily by
formulating them as DEs. Differential equations were first used in the early eigh-
teenth century to solve problems in mechanics, but more recently the theory of DEs
has found its way into the social and life sciences.
y’ =f). (1)
For instance, if f(x) = x’, the unknown function y is obtained through a simple
integration to yield
x?
where C is a constant of the integration which can assume any value. Other
F)
examp or VES
ir
LINENe. (3)
oe: ky Sis:
*Sir ISAAC NEWTON (1642-1727) was born on Christmas Day. Although mathematics began as a
recreation for Newton, he was soon known as a great mathematician after his invention of the calculus,
discovery of the law of universal gravitation, and experimental proof that white light is composed of
light of all colors. He accomplished all this before the age of 24! A controversy erupted between
Newton and Leibniz over the invention of the calculus, brought on primarily by friends of Newton who
accused Leibniz of plagiarism. Eventually both Newton and Leibniz themselves became occupied with
the controversy, even though both were convinced that they had reached their results independently.
Newton spent the last two years of his life in constant pain, and he finally died in his sleep on March
XO), WPT,
+GOTTFRIED WILHELM VON LEIBNIZ (1646-1716) is considered one of the great thinkers of modern
times. Known as both a philosopher and mathematician, he is credited with building a remarkable
system of modern philosophy and was greatly responsible for the development of the calculus. The last
years of his life were saddened by ill health, controversy, and neglect. A bitter controversy had been
started by friends of Newton over the legitimacy of Leibniz’s discovery of the calculus. Newton had
also discovered the calculus a few years earlier than Leibniz but failed to publish his results until after
Leibniz had published his. Leibniz died alone on November 14, 1716, and only his faithful servant
attended the funeral. Many members of the scientific community were not even aware of his death until
almost a year later.
{Derivative identities such as d/dx sinx = cosx are not included in our basic definition of DEs.
2
SEC. 1.2 / CLASSIFICATION OF DEs 3
When the ankeowa Fangtion pean upon moreethan onei neeencent vari-
able, the derivatives will be partial derivatives and the equation is called a partial
differential equation. Examples of ordinary DEs are given by (3) through (7) above,
while (8) through (10) are partial DEs.
Another classification is according to the order of the DE.
Definition 1.1 |
CHAP. 1 / BASIC CONCEPTS
eran (11)
Or
For example
Disciplines such as economics and the biological sciences are also now making use
of DEs to investigate problems in
7. interest rates,
8. population growth, and
9. the ecological balance of systems,
able is often the most critical part of a problem. Sometimes certain aspects of a
problem seem relatively unimportant and can be modified by assuming an approxi-
mate situation; sometimes an aspect of a problem may even be entirely eliminated.
The DE resulting from any such assumptions will actually be that of an idealized
situation.
Even after making a number of simplifying assumptions, the mathematical
formulation of a problem can still lead to a DE that is troublesome to solve. This
is particularly true when the resulting DE is nonlinear, since they are generally
difficult or impossible to solve exactly. Linear DEs are much easier to handle in
many ways, mostly because various properties of their solutions can be character-
ized in a general sort of way and standard solution techniques have been developed
for solving many linear DEs. For this reason, linear equations occupy a more
prominent place in the theory and applications of DEs.
Because of the difficulties encountered in solving most nonlinear DEs, they are
often approximated by linear DEs when the resulting linear equations can describe
certain fundamental characteristics of the nonlinear systems. For instance, the angle
6 that an oscillating pendulum of length 6 makes with the vertical direction
(Figure 1.1) is governed by
Ci)
then for small angular displacements 6, it may be reasonable to set sin = @, and
thus (13) can be replaced by the linear equation
4
a 0 (14)
(fe ae
In this particular example the linear equation (14) gives fairly accurate information
about the behavior of the pendulum when @ is small. There are a number of
applications, however, for which this approach is fruitless, since the phenomenon
being studied does not lend itself to any reasonable linear approximation. In such
cases we try to solve the nonlinear DE itself. Techniques for handling such DEs will
be discussed somewhat in Chapters 2 and 7.
EXERCISES 1.2
For each of the following DEs, state whether the equation is linear or nonlinear, and
give its order.
es Nhs |
SeNC 1.) —35+3=—+y=0
Sera a s 2. y' +x’ =0
wohpetbe rT eles : sei
3. (x7? + y’) dx = 2xydy rh 4. y' t+a@y =f(x)
[ IP
ee. View)
y ets Oy ea Vineex 6. saancs — bP?
A
Ss 4y x . \ f|
SN . Te mt
i 1) 8.F y ho .eS ||
Pee) 2 ls OOLUTIONSIORIDES
ous requ
Definition 1.3
EXAMPLE 1 Verify that y = e “is a solution of y’ + y = 0, and state the interval of validity.
Solution The function y = e * is continuous and has derivatives of all orders for all x.
Furthermore,
SEC. 1.3 / SOLUTIONS OF DEs 7
It should be pointed out that a given DE will usually possess many solutions
example, it is easily verified that)y== Cen rene: ihe DE i in Example | for any
value of the constant C. Thus we obtain the family of solutions shown in Figure 1.2.
Even the trivial solution y = 0, obtained by setting C = 0, is a solution of
y +y=0.
c=4 ‘
C=3
C=2
ae Cai
oo ee
;
ae
ee
C=-2
C=-3
C=-4
Figure 1.2 Solutions of y’ + y = 0.
EXAMPLE 2 Verify that y; = C, cos 2x and y) = C,sin 2x, where C, and C) are any constants,
are both solutions of
y" + 4y =0.
so that
y! + Ay, = —4C,cos2x + 4C,cos2x = 0.
and
y = C;sinxcos x
are also solutions of the DE, the verification of which is left to the reader.
EXAMPLE 3 Verify that g(x, y) = 0 is a solution of y’ = x/y, where g(x, 7y) =x? —y?— 1.
d d
—(x*)
ot — ee
—(y?) =0 or 2x — 2yy’
vA =0;
and solving for y’ yields
(y' = -1
clearly does not have a (real) solution. The question concerning the existence of a
solution is very important in applications where the DE is our mathematical model
of some physical situation. If our model does 1not HO aUE it is probaaly a
poor model, since meaningful physical problems should lution:
Another question Anat solutions concerns uniqueness. In the Saree we have
discussed, the solutions are not unique. Nonetheless, certain additional require-
ments can be imposed in most instances so that a unique solution can be singled out
for those problems having solutions. More will be said about this situation later.
The existence and uniqueness questions are generally difficult to answer. Usu-
ally it is preferable to discuss these questions with respect to rather narrow
classifications of equations, and this we do to some extent as we go along. A more
practical question that concerns us is, How do we find solutions? We will devote
most of our efforts in succeeding chapters to answering this question. Observe,
however, that once a solution has been found, we have answered the question of
existence. But if no solution is found, it is not clear whether this is because a
solution doesn’t exist or simply because we are unable to find it. It is embarrassing
SEC. 1.3 / SOLUTIONS OF DEs
m* —m — 12 =(m = 0.
— 4)(m + 3)
Thus, m = 4 or m = —3, and we find the two solutions
The reader should verify that y = C,e** + C,e~* is also a solution of the DE.
Solution Computing
p os y’ = 2x, it becomes clear that xy’
) — 2y) = 2. Also, we observe that
‘(y')? — y = 1. Hence, we conclude that given a particular solution function, we can
find several different DEs that it will solve.
From the examples considered thus far, it is apparent that DEs generally have more
than one solution—in fact, infinitely many solutions. For this reason we often seek
ee fev
Thati
is to say, be reduced to a fe
Solution The first solution, y = C,cos2x + C,sin2x, is a general solution, since no algebraic
manipulation will allow us to combine the two constants into a single constant. On the
other hand, we find that
10 CHAP. 1 / BASIC CONCEPTS
y = C\sinxcosx + C,sin2x
= 501sin2x + C,sin2x
= (50 a c,)
sin 2x
Z
=< C3 sin 2x
Solution In this case both functions do represent general solutions, since neither can be reduced
to a single-parameter solution. This shows that general solutions can assume different
forms for the same equation.
Using the definition of hyperbolic functions, we also note that
y = C,coshx + C,sinhx
ae Cle
l
He, Gt = ) Tek 31 Cle ¥, ren)
=<
illustrating that we can obtain one general solution from the other.
na
=] Sa
Since this situation arises only in connection with nonlinear DEs, and since most
of our analysis concerns linear DEs, such a distinction need not worry us for the
most part.
Another significant difference between linear and nonlinear DEs is that a linear
combination of two or more solutions is also a solution in the case of many linear
equations (specifically those for which F(x) = 0 in Definition 1.2). This property,
however, is generally not true for nonlinear equations, as illustrated in the follow-
ing example. This is one of the reasons why nonlinear DEs are usually much harder
to deal with than linear DEs.
Remark. ‘O or more s ; to ew S
the superpc n principle This principle, which HE an aancnan role
throughout our
o Frectision of linear DEs, is first introduced in Chapter 4.
EXAMPLE 8 Show that although y, = e * and y, = e* are solutions of both the linear equation
y" os y' i 2y =0
We a= 0,
the expression y = Cje * + Cye” is a solution of only the linear equation for
arbitrary nonzero values of C; and C).
Solution By direct substitution of y = Cje * + C,e~* into the linear DE, we see that
which is clearly not identically zero for arbitrary values of C, and C). In fact, it is zero
only when either C, or C; itself is zero.
Throughout the calculus we have become accustomed to the idea that the
“solution” of a problem is an exact, closed-form, analytical expression in terms of
elementary functions from which numerical values can be generated when needed.
In practice, however, this is rarely the case, especially when solving a DE. That
is to say, comparatively few DEs occur in applications for which simple exact
solutions can be found by known methods. Hence, in practice the “solution” often
consists of some type of approximation function or a set of numerical values
approximating the true numerical values of the solution over some interval. Several
such approximation methods are available for this purpose, and they are becoming
increasingly more common due to the complexities of the DEs arising in modern
science and engineering problems, and also due to the widespread availability of
modern high-speed computers.
In spite of what we have just said, most of the problems encountered in this text
will lead to exact solutions in much the same way as we have come to expect in
the calculus. The reason is that studying problems with exact simple solutions aids
the student in understanding the general theory and solution methods available; for
example, they provide an answer that can often be easily checked. Some approxi-
mation methods are briefly discussed in this text, but these techniques should be
taken up in more detail, such as in a course in numerical analysis.
EXERCISES 1.3
In problems 1-14, verify that the given function is a solution of the specified DE.
Assume C, and C, denote constants.
as + (1+ x?) dy = 0;
(1 + yo)de —1 = 0
xy ta ty
J Byd= xy, x =G
= Cys
6. Oy ==, Vy
= Ve +3, 2 >0
f—», aP ae®
N( 7.) —
dt = aP — bP?, P =~
(1 + be”) (a, b constants)
' LA s
Bs Mey ote) eae” y= pret + x #0
xX
Py " — 2y’ + Sy
= 0; y = Cie*cos2x
+ Cre*sin2x
12. y"+y’ — 2y = 2x — 40cos2x:
2|
%
0
Vv >
Each of problems 15-20 has a solution of the form y = e™ for some value of the
constant m. Determine for what values of m the following linear DEs have such a
solution.
& ne = ae
45) y'-2z=0 © 16. y"-y =0
Y)\ oe 18. y” + 2y’ =0
In problems 21-25, not all constants that appear in the given expression are
essential. Rewrite the expression using only essential constants.
In problems 28-35, find a first-order DE involving both y’ and y for which the given
function is a solution. sy! Rie
aA pe 1
j 28. 1p = 69" ate = A C9)oy =x-4 30. y= zxe"
vf vf ‘
wan K
| f
14 CHAP. 1 / BASIC CONCEPTS
Ye ee ee
> for any choice of the constants C, and Cp. Explain why this is not a general solution.
&39.) Show that the piecewise-defined function
lee ieee O
tas Ms je = 0)
EN leet Sa 0
: VI-x, O<x<1
also a solution? Explain.
Verify that y; = 1 and y. = x° satisfy each of the DEs
xy” 3 y’ = 0 and 2yy" es (y')? as 0,
but that y = C; + C2x’ only satisfies the first equation for arbitrary values of C, and
C>. Explain.
42. Verify that y,; = 1 and y. = x'” satisfy each of the DEs
but that y = C, + C2x'” only satisfies the first equation for arbitrary values of C, and
C2. Explain.
Solution The general solution y = Ce* is a family of curves in the xy-plane (Figure 1.4). By
specifying y(O) = 2, we are seeking that one particular curve which passes through the
point (0, 2). Thus, by substituting x = 0 in the general solution, we obtain 2
= Ce°®,
or C = 2. The solution we seek is therefore
y = 2e’.
SEC. 1.4 / INITIAL AND BOUNDARY VALUE PROBLEMS 15
; g~ The auxiliary conditions require that the solution we seek pass through the point (0, 2)
of the xy-plane with zero slope at that point. Imposing these conditions on the general
-m>
SS) solution leads to _
y = 2cosx.
y = sinx.
When the auxiliary conditions are all specified at a single value of x, such as in
Examples 9 and 10, we refer to the problem as an initial value problem. Although
we could choose the single value of x as any value, it is customary in practice to
use x = O. If the auxiliary conditions are specified at more than one point on the
interval of interest, the resulting problem is called a boundary value problem (see
16 CHAP. 1 / BASIC CONCEPTS
Example 11). The boundary conditions are usually specified at only two points of
the interval, and we refer to these problems as two-point boundary value problems.
Such problems involve DEs that are at least second-order, since first-order DEs
have but a single auxiliary condition and as such are classified as initial value
problems.
In general, initial value problems are well behaved in that they almost always
lead to unique solutions. Unfortunately, the same is not true of boundary value
problems, and so the general theory of these problems is more complicated. The
following examples give some hint of the difficulties that arise in solving boundary
value problems.
EXAMPLE 12 Given the general solution y = C;cosx + C,sinx, solve the boundary value problem
y’+y =0, yO) =0, yo) = 2.
y Oe G a0
and
y(m) = Chsinaw = 2.
But since sin 7 = 0, we conclude heh no member in the tebe family of Seats
satisfies both boundary conditions. “Since the equation is li e know t the DE
praseeisemive any SireLAP SOLUTIONS: SO we must accept the factthatrans pou value
problem has no solution.
EXAMPLE 13 Solve the boundary value problem in Example 12 if the boundary conditions are
changed to y(0) = 0, y(7) = 0.
y(0) =C, =0
and
y(ar) = Ci sin7 = 0,
which are satisfied for C; = 0 and any value assigned to C). Hence, we obtain the
family of solutions
y = Cysinx,
where C, is an arbitrary constant.
EXERCISES 1.4
Solve the initial value problems 1-10, for which the general solution is specified.
-_ ; : iT
Find solutions (if they exist) for each of the boundary value problems 11-15. The
general solution is provided.
By 0;
y =Cre
y"-y =0, yO) =0, y')= +Cre
46. Verify that y = C,; + Cox? is a solution of the linear DE
xy Sy = Oe
REFERENCES
Boyce, W. E., and DiPrima, R. C. 1977. Elementary differential equations. 3rd ed. New
York: Wiley.
Eves, H. 1964. An introduction to the history of mathematics. 3rd ed. New York: Holt,
Rinehart & Winston.
Ross, S. L. 1980. Introduction to ordinary differential equations. 3rd ed. New York: Wiley.
Zill, D. G. 1979. A first course in differential equations with applications. Boston: Prindle,
Weber & Schmidt.
In this chapter we consider certain basic types of first-order DEs for which exact:
solutions may be obtained by clearly designated techniques. Proficiency in solving
such equations rests heavily on the ability to recognize the various types of DEs and
apply the corresponding method of solution, which often consists of certain
“devices” or “tricks.” This is particularly true when solving nonlinear DEs, since no
clear underlying theory can be applied to them.
In Section 2.2 we discuss equations that can be solved by the method of sepa-
ration of variables. This is the easiest technique to apply and comes up most often in
practice. Other equations that permit exact solutions in closed form are discussed in
Section 2.3, although some of them occur infrequently in applications. These include
exact equations, equations solvable by the use of integrating factors, linear equa-
tions, and homogeneous equations which are solved by reduction to variables
separable by a change of variable.
In Sections 2.4 and 2.5 we again discuss linear DEs of the first order, but in the
context of the general theory of linear equations of any order so as to provide a
more comprehensive approach. Thus we introduce the notions of a homogeneous
solution and a particular solution of a DE, the sum of which forms a general solution.
A physical interpretation of a general solution is given, and the basic existence-
uniqueness theorem concerning linear initial value problems of the first order is dis-
cussed.
A special nonlinear equation, called Bernoulli's equation, is solved in the final sec-
tion of the chapter by first reducing it to a linear equation through an appropriate
change of variable.
For a shorter course, Sections 2.4 through 2.6 can be omitted.
18
2.1 INTRODUCTION
d wa M(x, y) a
In applications the solutions of these DEs are usually required to satisfy an auxiliary
condition of the form
Theorem 2.1 If F and dF/dy are both continuous functions in some domain of the xy- plane
containing the point (xo, yo), then there exists a unique solution of the initial value
problem
y' = F(x, y), yo) = yo,
*For a proof of Theorem 2.1, the reader should consult E. I. Ince, Ordinary Differential Equations
(New York: Dover, 1956).
19
20 CHAP. 2 / SOLUTION TECHNIQUES FOR FIRST-ORDER EQUATIONS
n= (3) ’ ge)
yz =0
is another solution (called a singular solution). A unique solution does not exist
because
aF _ —a ( uy
ite 2/3
=-y
‘ay Oy 3
is not a continuous function in any region containing the x-axis (y = 0). If we
prescribe the initial condition at any other point (xo, yo) not on the x-axis, a unique
solution can then be found.
SEPARATION OF VARIABLES
(4)
or
(6)
dy) _\ ax
igh y hing See Ov4 Rte
Thus, integrating both sides, we get ~~ \i) X
In Example | we are using the symbol logx to denote the natural logarithm, also
commonly denoted by Inx. Since integrals leading to logarithmic terms are quite
prevalent when employing this solution technique, some care should be exercised
in evaluating these integrals. Recall from the calculus that
\
die
[= = gle +c, ago)
u
where the absolute value is usually retained unless we know in advance that u > 0.
As a general rule, however, we will retain the absolute value only in those cases
for which a logarithmic term remains in the final solution form of the DE.
dy (a a '
a?2 |
eet
x
and integrating yields
ieee | e
i + Z loge ,/\C >0.*
A |3 los Gy + |)= a
3y +1 = Ce
or
eiaGGs
= 1}
Although we technically have the general solution of the DE once the integration
has been completed, it is usually advantageous to simplify the algebra in this
solution when possible. The simplicity of the final solution in the next example
clearly illustrates the gain resulting from a little algebraic maneuvering.
ih AP Se
tan (arctan y + arctanx) = i
_ xy ;
WP Se BS
= tanC.
Leaety
Applying the initial condition y(0) = 1 reveals that tanC = 1, and hence
WS Vs
fx
Or
SEC. 2.2 / SEPARATION OF VARIABLES 23
ae, y y
In separating the variables, it was necessary to assume that x # 0 and y # 0.
Note, however, that y = 0 is also a solution of the DE, which cannot be obtained
from the general solution by selecting an appropriate value of C. It is a singular
solution of the problem.
EXERCISES 2.2
dV V oh aN
10. —S> = -—— =
4, a)a7 ae N = Nt
e
Lace
In problems 26-30, obtain the particular solution satisfying the prescribed auxil-
iary condition.
~ d
26. oa =g, Uv(to) = vo 27) y’ ' =—2xy, y(0)=yo
*31. Giveny’ = y? — 4:
(a) Derive the general solution
heed Us Cer)
die hay he ehegs
(b) Show that the initial condition y(0) = 2 cannot be satisfied by the solution in (a).
(Can you explain?)
(c) By inspection, obtain two singular solutions and verify that one of these satisfies
the initial condition in (b).
*32. Given the initial value problem y’ = xy'?, (0) = 0:
(a) Find two particular solutions of the DE satisfying the prescribed initial condition.
Hint: One is a singular solution.
(b) Verify that the piecewise-defined function
O; #2 <i
y= AND
ad x = 1
/ 16
42 vA is also a solution. Are there any others?
Y\ (*38.) Solve xy’ — y* + 1 = 0 subject to
a) yO CD) y (0) ade (c) y(O) = 2.
*34. Solve y’ — x*(y? + 1) = 0 when
@tk a =i (b) k = -1.
*35. Find all solutions of (vi) + yy? sata a ee:
2.3 MISCELLANEOUS TECHNIQUES
Although most of the first-order DEs that permit exact solutions are of the separable
variables type, several other types occasionally arise for which exact solutions can
be obtained.
Recall from the calculus that if f(x, y) is a function of two variables, then its total
differential is given by the expression
df=eeade a+ Ol:
ay. (7)
In this case it follows that the solution of (9) is given by the family of curves
Haar) Ce (10)
where C can assume any constant value.
For example, if
2Qxy3dx + 3x*y2dy =0, (11)
then we recognize the left-hand side of this expression as being the total differential
of the function f(x, y) = xy? (see Example 5). Hence, (11) has the solution
yr =C (12)
for any constant C.
Equations like (8) for which the left-hand side is an exact differential are called
exact differential equations. The difficulty in solving such equations is that it is not
always easy to determine by inspection whether a given differential expression is
exact, as we determined in the above example. Therefore, the following theorem
provides us with a test for deciding whether the expression M(x, y)dx + N(x, y) dy
is an exact differential, and the proof provides us with a scheme for finding the
function f(x, y) in those cases for which it is an exact differential.
25
26 CHAP. 2 / SOLUTION TECHNIQUES FOR FIRST-ORDER EQUATIONS
Theorem 2.2
Proof: To prove the necessity part of the theorem, we merely observe that if the
expression is exact, then
0
M(x, y)dx + Mx, y)dy= Las + od
and hence
Ofaas of
ee M, 7 N
Bol
ey CM ein ey
Oydx
0 <ey > oxoy mmow
from which we deduce 0M/dy = dN/dx.
To prove the sufficiency part, let us assume that 0M/dy = dN/0x and
ie i)
eee
If we formally integrate this latter expression with respect to x, we find
Ze
0
jp[Mes de + 8G) =Nw»).
The function g is then a solution of the first-order DE
Remark. In the sufficiency proof of Theorem 2.2, we could just as easily have
started with df/dy = N(x, y) and integrated to find
6)
h"(x) = Mx, y) — = |N(x, y)dy.
EXAMPLE 6 _ Show that the DE (4x — 2y + 5)dx — (2x — 2y)dy = 01s exact and find a gen-
eral solution. | ;
ia
Yu a a
=) | dy Ax
Hence the equation is exact. By Theorem 2.2 a function f(x, y) exists such that
=of
ay =4,-2y
a y +5 and
an —of
ay =2y 'y — 2x.
Be if
4 a\'=>— JU
It now follows that
0M oN
oy Bs
Thus,
OE a
== l
of
ax ey a daay xy sin Ms
For the sake of variety, let us integrate the second of these expressions this time to
get
f(x, y) = xy? + cosy + h(x),
from which we obtain
of
Neues ee = I
Hence, h'(x) = —1 or h(x) = —x, and our solution is
(15)
and thus the equation is not exact. However, if we multiply the DE by x, the new
equation
es x
y dx z nS a(®) (17)
y y
ydx + xdy = d(xy) (18)
ie OE (tos) (19)
xy y
pein
= alas
eyy= a{tun'*) \ (20)
Any y \
Solution The combination xdy — ydx suggests division by either y* or x*. Here division by
y*, accompanied by a division by x, leads to the expression
fs.
log|x}g |x| ;
+-=C
This section can be omitted if Sections 2.4 and 2.5 are discussed.
In Chapter 1 we defined the general form of a linear DE. For first-order equa-
tions, this form reduces to
d
MOXOLy’ + ao(x)y] = pyleooyl = LOY + feey (24)
M(x)ao(x)y = mw'(x)y,
or
b'(x)
a)
es)o(X) (25)
25
Mx) = exo|
|do(x) ax}, (26)
Of course, s(x) is not unique, since any constant multiple of it also does the job.
Returning now to (23) multiplied by (x), we write
d
Hoy] = p(x)f (x).
Therefore, it follows that
or
es
see - |M(x) Mix)f(x) dx
dx C Ce (27)
The function y given by (27) represents a general solution of Equation (23).
Remark. This solution technique has the disadvantage of not being adaptable to
higher-order linear DEs. A more encompassing technique is discussed in Section
BX
Let us now consider a class of DEs that can be reduced to the separable variables
type by a change of variable.
If the DE
M(x, y)dx + Mx, y)dy = 0 (28)
*The term homogeneous as used here should not be confused with the meaning of this term as used
in subsequent sections and chapters.
32 CHAP. 2 / SOLUTION TECHNIQUES FOR FIRST-ORDER EQUATIONS
N(x, y) = vn(1, AE
x
and hence (28) can be written in the form
dy _ _ MG, y) bil _ x"M(l, y/x)
dx N(x, y) x INCISED)
or, equivalently,
dy am A yCal
rae 30)
dx * (*) (
This last form of the DE suggests either the substitution y = vx or x = vy.
Setting y = vx, we get
Solution We note first that M and N are both homogeneous functions of degree 4. If we try
the substitution y =vx), there follows
(XX + x4v4) dx + 2x*v(vdx + xdv) = 0,
or, if x #0,
(1 + 20? + v*)dx + 2xvudv = 0.
Separating variables, we find
‘dx
eee
20 Be =
x Neeacehely ee
which yields the solution
SEC. 2.3 / MISCELLANEOUS TECHNIQUES 33
log
|x| =o" sh 1p t= Cc.
Replacing v with y/x now gives
Vere C
1 thatof dy, it 1s somewha
_x = vy rather than y = wx. Making this substitution, we obtain
dv
= ee ea
Il Se “Be y
which leads to
arctan v + log|y| = C
or
3
arctan (*)+ logly|=C, y #0.
EXERCISES 2.3
In problems 11-15, use the suggested integrating factor to solve the equation.
/
f
* 2, bilge. 2 ps 2 = {): =
1 :
142 Go + oxy = y*) de (ys 2x — % dy =0 Gane
1
15. (+ 2 ores
y°—x)dx —
—ydy == 0;1) pb =
Ona,
In problems 16-22, find an integrating factor by inspection and solve the equation.
In problems 23-32, verify that the given DE is linear and find its solution. ACheck
for linearity in both x and y.)
In problems 33—40, show that the equation has homogeneous coefficients and solve
the equation.
a Tae tee : aD >
+ @? — xy)dy=0
'338) Ge? + y*)de
~pi 3erga N34) xy’ =y He
eae
(u+v)du + (v —u)dv =0 36. xdx + (y — 2x)dy =0
xydx + (x? + y*)dy =0
*38. (s —1(4s + Dds + s(5s — thdt =0
‘39. ( — ylogy + ylogx)dx + x(logy — logx)dy = 0
44//x = 2y)dx + Ay — dy =0
“AS. xdy + 3ydx = (x — 2)dx (Solve two ways.)
46. (y — x)dx + dy =0 (Solve two ways.)
47. (xy? + x — 2y + 3)dx + x’ydy = 2x + y)dy, y(1) = 1
48./ x*y’ — y — xy = 0 (Solve two ways.)
Ue en ree ydy = y*(x? — y”)ady
50/ 2ny'=1+y%, yQ)=3
*31. 3ydx + 2xdy = xy*(xdy + ydx)
52. Sear es+ (y? — x*)dy =0
Prove that
p Pevaxyrrby
ad — bc #0
cx + dy”
is an exact DE if and only if b + c = 0, and find a general solution in this case.
*54. Prove that 4 = u(x, y) is an integrating factor for Mdx + Ndy = O if and only if u
satisfies the relation
it — yt 4 (2M _ aN)
dy Ox dy Ox
55. Using the result of problem 54, show that if u(x, y) = x’y’ is an integrating factor
of Mdx + Ndy = 0, then p and q must satisfy the relation
OM aN
pyN — qxM = »(Sy -S)
Ox
56. Use the result of problem 55 to obtain an integrating factor of the form w(x, y) = x”y4
and solve the equation.
Gey Cyaan idx cdye—() (OO) Cae) decd yaa)
(c) 2ydx + 3xdy = 3x~'dy
p57: Any DE of the form
y= xy fly)
is called an equation of Clairaut.
(a) Show that a general solution of this equation is the family of straight lines
y = mx + f(m)
where m is an arbitrary constant.
(b) Show that this DE also has the singular solution given by the parametric equa-
tions
Dale bet) ae eet Ata).
Hint: Differentiate both sides of the Clairaut equation with respect to x to obtain
Passa)ya 0:
58. Referring to problem 57, solve the following Clairaut equations:
The general theory of linear DEs is discussed in Chapter 4. However, one can
achieve a great amount of insight into this theory by first developing the corre-
sponding theory for first-order linear equations. Although we have presented a
solution technique for this type of DE in Section 2.3.3, it cannot be adapted to
higher-order equations and therefore cannot provide the kind of insight into the
general theory that we are seeking.
A first-order DE is said to be linear if it can be arranged in the form
Dy =y’. (35)
A more general example of a differential operator is
M =D + a(x), (36)
where a,(x) is any function. We interpret this operator such that
*The use of the term homogeneous as given here and in succeeding chapters is not the same as
implied
in Section 2.3.4.
36
SEC. 2.4 / THE THEORY OF LINEAR FIRST-ORDER EQUATIONS 37
(43)
*In general, a linear operator M is one for which M [Cif (x) + Cog (x) = Ci M[f(®)] + C.M[g (x)] (see
problem 27).
38 CHAP. 2 / SOLUTION TECHNIQUES FOR FIRST-ORDER EQUATIONS
The function y, represents the only nontrivial solution (to within a multiplicative
constant) of Equation (38). Thus (45) is truly a general solution ,since it contains
all solutions of (38), including the trivial solution y = 0 obtained by setting
C, = 0. Also observe that (45) reduces to (40) when a(x) = a.
In Section 2.5 it will be necessary to distinguish the solution function given by
(45) from another part of the general solution required in solving nonhomogeneous
equations. Therefore, the symbol y, will be used in that section to identify (45),
while here we will continue to use the symbol y alone since it is the only solution
that presently concerns us.
or, equivalently, at 3
sp C cee nC =e oN
Of course, we could have used (43) as a solution formula, which requires only
that we identify ao(x) = 2x. Completing the integration in (43) then leads to the
same result.
v= (pee
Solution Assuming x # 0, we divide the DE by x to get the normal form .
Thus, do(x) = 1/x, and substituting this expression for ao(x) into (43) gives us
a C
y = Cjexp(-[©) =e es 0.
XxX x
By imposing the initial condition y(1) = 3 on this solution function, we obtain the
- value
SEC. 2.4 / THE THEORY OF LINEAR FIRST-ORDER EQUATIONS 39
VU Os),
Hence, the solution we seek is
If we wish, we can use the general solution (45) to produce an explicit solution
formula for the initial value problem
y'’ ta(x~y =0, yx) = ko. (46)
That is, imposing the initial condition in (46) upon the general solution
y = C\y\(x), we find
k
C1. =—, (47)
yi(Xo)
and therefore the solution of (46) can be put in the form
k
y = OD) vil) # 0. (48)
yi(Xo)
If ao(x) is continuous throughout an interval containing the point x = Xo, then it
cannot happen that y,(%9) = 0 (see problem 26). In such a case, (48) is the unique
solution of (46). For those cases where ao(x) is not continuous, the initial value
problem (46) may still have a unique solution, but the problem may also have no
solution or more than one solution. Consider the following example.
Solution Dividing the equation by x, we see that a(x) = —4/x has a discontinuity atx = 0.
Nonetheless, we can still produce the general solution
VSS Cie
which is a well-behaved function for all x. Imposing the prescribed initial condition
on y, however, does not require C; to assume any particular value, and hence the
given problem has infinitely many solutions.
We should observe that if a nonzero value was prescribed for y(0), the problem
would have no solution.
40 CHAP. 2 / SOLUTION TECHNIQUES FOR FIRST-ORDER EQUATIONS
EXERCISES 2.4
(b) e (1/2) log x ae Vx, x>0 (d) e707 +2 logx) = >, x >0
x
a dy dy y
4. y' f + (tanx)y = =0 5! a
—_—_— = xy 6. ae
—>
= >
In problems 13-24, find a solution of the given initial value problem and state the
interyal for which the solution is valid.
(50)
ge tion 49). To see this, we will assume Y is any solution of (49)
and ieaney=ZY — yp. The direct substitution of y = Y — yp into (49) gives
Ye= ori)
and therefore belongs to the general family of solutions described by (50). Since
Y could be any solution of (49), we conclude that (50) is a general solution.
Summarizing, we have the following theorem.
‘EXAMPLE 16 Meaty. that yp = —2x — | and zp = e** — 2x — | are both particular solutions of
y’ — 2y= 4x, and find a general solution in each case.
and
that the pa
/any pa
which reduces to M[ yp] = u'y, since y, satisfies the homogeneous equation. Now
if also M[ yp] = f(x), then clearly the function u(x) must be chosen such that
u'(x)y(x) = f(x),
or
SEC. 2.5 / THE NONHOMOGENEOUS LINEAR EQUATION 43
u(x) =
F(x)
|——dx + C. 4
yi) Ce
Technically any constant C can appear in (52), and the resulting function will lead
to a particular solution, so it is customary to set C = 0 and write the particular
solution as
(53)
Combining yp with the homogeneous solution function y, leads to the formula
ye(ipen sea 2
Solution We first rewrite the equation in normal form,
I g
ote | en: # 0, }
and hence
1 e*
yi) = exp |
—[( ~~ 1)ar|Sti
be
which, added to the homogeneous solution yy = C\e*/x, gives
l C\e*
y= ae = aa x #0.
= —2cos*x + C,cosx,
where we have made use of the identity sin2x = 2sinxcosx. According to the
initial condition, we find
9) = C, = =),
When checking the linearity of a given DE, checking for linearity in just one
variable is not always enough, as our next example illustrates.
EXAMPLE 19 Find the solution of dy/dx = 1/(x + y’), which passes through the point (—3, 0)
of the xy-plane.
Solution The DE is not linear in y due to the presence of the term y*. However, if we invert ©
the equation, we get
y = x de Vw Ola, dx
2) eee Vie
dy
which is linear in x. Thus, interchanging the roles of x and y in (54), we first
calculate
and hence
To satisfy the prescribed condition x(0) = —3, we find that C; = —1, and so
Hence we write
y= Yeok Ya
3 FO 4,4 ku)
y =yi0) | SO ey >?
provided y,(%o) # 0.
Remark. It may be interesting to note that the solution formula (57) for the initial
value problem described by (55) depends only upon the solution function y,(x) of
the associated homogeneous DE and the input parameters ky and f(x). This is a
fundamental characteristic of all linear nonhomogeneous DEs and is the primary
reason why so much effort in succeeding chapters on linear DEs is directed at
finding solutions of homogeneous equations.
Not only have we derived the solution formula (57) for the initial value problem
(55), but our choice of yp and yy now leads to important physical interpretations.
For instance, since yp(x9)= 0, we can think of yp [defined by (56)|as the solution
of the initial value problem
or both limits
*It is customary to introduce a dummy variable of integration such as in (56) when one
of integration are variable.
46 CHAP. 2 / SOLUTION TECHNIQUES FOR FIRST-ORDER EQUATIONS
wherein the initial condition is homogeneous. The physical implication of yp, then,
is that it represents the response of a system at rest which at some time x = Xp is
subjected to the external disturbance f(x). Moreover, since yy must now satisfy the
auxiliary condition y;(xo) = ko, it is a solution of the initial value problem
[We should take note of the fact that if we were using (54) to solve this problem,
we could not impose the initial condition at this step of the solution process.|
Substituting y, into (56), we find
y = yp + yy = —2x — 1 + Ge.
pe ee
One of the difficulties that arises from time to time in solving initial value
problems is that the integrals that need to be evaluated are not elementary. Consider
the next example.
SEC. 2.5 / THE NONHOMOGENEOUS LINEAR EQUATION 47
eee ee ee ee ee ee ee Se Se
yo —2xy = 1, yi0)= 3.
y= eaiee 2 ci + 3e,
y= [sVeeton + Jer.
This last example illustrates that even simple-looking DEs may lead to non-
elementary solution functions such as the error function. A host of other functions
of this type falls under the category of special functions. Many of these other special
functions also arise in the process of solving a DE that does not possess solutions
expressible as elementary functions. As long as these new functions are tabulated,
we can treat them essentially as known functions. In computing numerical values
of functions, there is little difference between consulting a table to evaluate an
exponential or trigonometric function and consulting a different table to evaluate,
say, the error function.
.
2.5.2 Existence-Uniqueness Theorem
y = Cry;(x)
where
Since ao(x) is assumed to be continuous, its integral is also continuous, and there-
fore y; is always positive. Imposing the initial condition now yields
y(xo) = Ciyi%) = 0,
from which we deduce that C; = 0. Hence the only solution of the initial value
problem is y = 0.
We are now prepared to state and prove the following important existence-
uniqueness theorem for linear initial value problems of the first order.
Theorem 2.4 If M =D + a(x) is a normal differential operator on the interval x, <x < x
containing the point x = xo, and f(x) is continuous on this same interval, then the
initial value problem
“£0 4, koi) \
Yen de xo Yi(d) yi(Xo) ”
where
iz Te I xf koyi (x)
eRe ano aty sien
and so it follows that
= f().
SEC. 2.5 / THE NONHOMOGENEOUS LINEAR EQUATION 49
To show that y is a unique solution, let us assume that uw, and uw are both solutions
of the initial value problem and define z = u; — u2. The function z then satisfies
the homogeneous equation
Remark. Observe that the existence theorem in Section 2.1 (Theorem 2.1) also
applies to the linear equation discussed here. The advantage of Theorem 2.4 over
Theorem 2.1 for linear DEs is that it not only informs us of the existence of a
solution but also provides us with a solution formula. ~
Some comments about Theorem 2.4 should be made at this point. First, the
theorem not only guarantees the existence of a solution but also states that there is
only one solution, i.e., that the solution is unique. Furthermore, the solution is
differentiable and has a continuous derivative on the specified interval. This can be
observed by writing the equation in the form
y’ = —ax)y + f@),
where the right-hand side is composed of continuous functions. However, the
condition that ao(x) and f(x) be continuous on the interval of interest is not a
necessary condition, but a sufficient condition of Theorem 2.4. For instance, in
Example 17, the solution y = 5xe* + C,e*/x is not valid at x = 0. Yet if we
impose the initial condition y(0) = 0, then we can select C; = 0, and the solution
reduces to the well-behaved function y = 4xe* for all values of x. In other words,
the situation here is similar to that discussed in Section 2.4. That is, if ao(x) is not
continuous, then a unique solution may still exist, but it is also possible that either
no solution exists or more than one solution exist.
When the function f(x) is not continuous, there is also no guarantee that a unique
solution of the initial value problem exists. And in certain applications the input
function f(x) commonly exhibits a discontinuity at some point due to, say, a switch
being (instantaneously) turned on or off.
Pere er
FO et OF Os n= al
50 CHAP. 2 / SOLUTION TECHNIQUES FOR FIRST-ORDER EQUATIONS
Solution Here the input function has a jump discontinuity at x = 1, so we solve the problem
in two parts,
y ty =0, yO=H1, Osx a1
and
We yyoas lees
The solution of the first problem is found to be
y =e) Ose reel:
For x > 1, we have y = 1 + Cye*, and hence we write
Soe) eee
ys Vile (1 eres
Although this solution is continuous, it is not differentiable at x = 1 (see Figure
2a
EXERCISES 2.5
eeUD oa
ekey 2c Yui
ye= Sy
SEC. 2.5 / THE NONHOMOGENEOUS LINEAR EQUATION 51
dy
er = cscx — ycotx “10. udt + Bt — tu + 2)du =0
di
“ah L7. + Ri=E, i(0)=0 (L,R, E constants)
di
22. L—+Ri=Asi
FF 1 sinwt, El i(0) == 0
The nonlinear DE
*JAKOB BERNOULLI (1654-1705) was a member of the famous Bernoulli family, which produced eight
mathematicians in three generations. Besides studying the equation named in his honor, Jakob’s
research included probability theory, the integral and differential calculus, and the calculus of vari-
ations wherein he solved the isoperimetric problem and, along with several other prominent mathe-
maticians, the famous brachistochrone problem proposed by his younger brother Johann.
SEC. 2.6 / BERNOULLI'S EQUATION 53
N
II
<
then
ea) yey
and (60) can be reduced to the linear DE in z
z= ee ia rca + Cye*
=se*
it ee
3 + Ce?noo) I
By 2 =e + Ce.
— B ae Ges
a
Imposing the initial condition (0) = 1 /Po on z(t), we find C; = 1/Po — b/a, and
hence
CeaP»
OO bP + (a — bPaje*
We should also observe that P = a/b is a singular solution of the Bernoulli
equation, but it does not satisfy the prescribed initial condition.
EXERCISES 2.6
REFERENCES
Boyce, W. E., and DiPrima, R. C. 1977. Elementary differential equations. 3rd ed. New York:
Wiley.
Finney, R. L., and Ostberg, D. R. 1976. Elementary differential equations with linear algebra,
Reading, Mass.: Addison-Wesley.
Reiss, E. L., Callegari, A. J., and Ahluwalia, D. S. 1976. Ordinary differential equations with
applications. New York: Holt, Rinehart & Winston.
First-order DEs can be applied in many diverse areas of the physical and life
sciences. In spite of this, we often find the same DE evolving out of our
mathematical formulations of problems, regardless of the field of application. That
is, the DE together with an initial condition is simply a mathematical model that
someone has developed of a problem under study, and the same model can often
be used to describe different phenomena, such as the motion of a free-falling body
and_the growth rate of a certain type of bacteria. Once the problem has been
_ properly formulated mathematically, the origin of the problem makes little difference
until we need to interpret the solution.
In Section 3.2 we discuss some general elementary applications wherein the
governing DE may be either linear or nonlinear. These applications include finding
orthogonal trajectories of a given family of curves, finding the velocity of free-falling
bodies, applying Torricelli’s law to the flow of water through an orifice, and finding
curves of pursuit that describe the path of a pursuer tracking its prey.
We restrict the applications in Section 3.3 to those involving linear equations.
Here we discuss problems of growth and decay, free-falling bodies, electric circuits,
cooling bodies, and the mixing of two solutions. Again we point out that all of these
areas of application have basically the same governing DE.
oo
55
3.1 INTRODUCTION
The examples in this section are typical of applications that lead to first-order DEs.
It is well known in analytic geometry that the slopes of perpendicular lines are
negative reciprocals; i.e., m, = —1/mp. In a more general setting, we say that two
intersecting curves are orthogonal if and only if their tangent lines are perpendicular
at the point of intersection.
56
SEC. 3.2 / GENERAL APPLICATIONS 57
FX ye) = 0; (1)
where each member of the family corresponds to a particular value of the parameter
c. In certain applications it is important to be able to obtain a second family of
curves given by
4
=p
F(x, y).
(4)
The procedure is therefore to find a DE (3) for which family (1) is a general
solution, and then obtain the orthogonal trajectories as solutions of (4).
l curves (i.e.,
*For instance, in an electric field the lines of force are orthogonal to the equipotentia
curves of constant potential).
58 CHAP. 3 / APPLICATIONS OF FIRST-ORDER EQUATIONS
aE
eoie aa iby Set
te) 0,
or
2x + 2yy' = 0.
Solving for y’, we find
: x
a ss ae
> Xx
Figure 3.2
Solution If we rewrite the equation as.y*/x = 4p, then differentiation leads to the DE
yas2a
The orthogonal trajectories must therefore satisfy the DE
pees y ’
Figure 3.3
Consider the problem of a free-falling body that is acted upon only by the force
of gravity. If the body is close to the surface of the earth, the weight of the body
is essentially the constant mg, where g is the gravitational constant. It has been
determined experimentally that at sea level g = 32 ft/s* (English system) or
g = 980 cm/s? (metric system). Newton’s law in this case is simply
oD
ae = mg,
where our sign convention is such that the positive direction is downward.
If the body encounters air resistance as it falls, then the weight mg must be offset
somewhat by a resistive force Fr. In such a case the governing equation is
dv
me = mg — Fr. (5)
60 CHAP. 3 / APPLICATIONS OF FIRST-ORDER EQUATIONS
The amount of air resistance depends upon the velocity of the body, but a general
law expressing this dependency is not known. We often make the assumption that
Fr = cv or Fr = cv”, where c is some positive constant.
EXAMPLE 3 A body of mass m is dropped from a height of 5000 feet. If the air resistance is
described by Fr = mv’/40, find the velocity of the body for any time t.
Solution At time t = 0, the velocity of the body is assumed to be v = O. At any later time,
the velocity is described by the solution of the DE
dv mv ~ Moye
FeBerge, :
Separating the variables, we find
dv
5)
Ate
40g —v*
Ae
\40
! =
which leads to
1 " (ae) =a
fiemaG:
AV 10g : dV 10g —"6 40
Imposing the auxiliary condition v = 0 when t = 0, we see that C = 0. Hence we
rewrite this last expression as
2V 10g ata) = 9 Velo:
PASAY gb
and, solving for v, we get the desired result
Veilor —
e Ve/l0r + |
=e, i =p
1 —e Vegi0r
1 +e" Veil0
Observe that as t>, we find v2 V 10g. This is called the terminal velocity,
or limiting velocity, of the body.
At large radial distances r from the center of the earth (distances beyond sea
level), the weight of a body differs from the weight at sea level, since the acceler-
ation a is not equal to the constant g. According to Newton’s law of gravitation,
the acceleration of a body is inversely proportional to the square of the distance
from the center of the earth. In symbols we write
ae: ier.
SEC. 3.2 / GENERAL APPLICATIONS
61
where k is a proportionality constant. If the body is falling to earth, the constant kis
positive, since the velocity is increasing; if the body is leaving the earth, k is
negative. Suppose we denote the radius of the earth by R and consider a body
projected upward (Figure 3.4). Then a = —g whenr = R, and thus —g =k/R’,
or :
k = —gR°.
Earth’s
surface
Figure 3.4
dv _ dvdr dv
dt drdt dr’
and thus (6) becomes
ee ee (7)
If initially the body left the earth’s surface with velocity uv, then we have the
auxiliary condition v(R) = vo. Hence, a particle projected radially outward from
the earth‘s surface with velocity vp will have the velocity given by
2
otherwise the body will stop and then fall back to earth with negative velocity. The
critical initial velocityvp is such that v6 — 2gR = 0, or vy = V2gR. The minimum
value of v, is
Ve = V22gR, (9)
called the velocity of escape.
EXAMPLE 4 Given that R = 3960 miles and g = 32 ft/s’, determine the escape velocity of the
earth.
Solution Converting units, we have that g = 32 ft/s* corresponds roughly tog = 6 X 1073
mi/s*. Therefore,
Ve = V(2)(6 X 10-*)(3960) = 6.9 mi/s.
Of course, the escape velocity for other heavenly bodies, such as the moon or
Mars, will be different, since both g and R are different for these bodies.
v = 0.6V 2gh,
where g is the gravitational constant and /h is the instantaneous height of water
above the orifice (Figure 3.5). Hence, if V is the instantaneous volume of water in
the tank, it follows that
dV
a = Av = —(06A \V 2eh, (10)
Figure 3.5
dl
oe = —0.6AV2¢h. (11)
EXAMPLE 5 How long will it take to empty a cylindrical tank of radius } foot and vertical axis
2 feet if the tank is initially full of water and the orifice is a }-inch hole in the bottom
of the tank?
Solution The volume of the water in the tank at a height of h units is V = 7 (4)7h. Therefore
wv
dh 4”
and (11) becomes
1edh Bats
—a— =) 675 75 V 2(32)h.
2h fe end G.
ay oe aa
770°
The initial condition is h(0) = 2, so thatC = 2V2. Solving fort when h = 0 then
leads to the result
t = 270(2V2) = 764 seconds,
or approximately 12 minutes and 44 seconds to empty the tank.
=
NJ) /
iS j 3.2.4 Curves of Linear Pursuit
P(x, y)
0 (a, 0)
along the y-axis (see Figure 3.6). The pursuer is assumed to be at the point (a, 0)
when Q is at the origin. It is further assumed that the speeds of the pursuer and prey
are always in the same constant ratio.
To illustrate, let us imagine a large field in which a fox is located at the point
(a, 0). He spots a rabbit at (0, 0) running along the y-axis in the positive direction
with constant speed v. The fox immediately runs toward the rabbit with speed w,
and after t seconds the rabbit is now at Q(0, vt) and the fox at P(x, y). The line PO
is always tangent to the path of the fox and therefore has the slope
y= tt
Yet ie
: x (12)
In order to solve (12), we need to eliminate the parameter t. To do so, we observe
that the length of the path traveled by the fox at speed w can be computed by the
arc length formula
Solving (12) and (13) for ¢ and equating the resulting expressions, we find
(14)
Although (14) at first appears to be formidable, let us differentiate it with respect
to x to get (after simplification)
At time t = 0, the slope of the pursuit curve is zero, so that we write p = 0 when
= a; hence, C; = kloga. Using properties of logarithms and some algebra, we
can rewrite (15) in the form
r=v=al(e) -@)}
If we assume that the fox runs faster than the rabbit, i.e., k < 1, the integral of
this last expression yields the general solution
y= tafe ea),
1 x/
lee %
EEK
Pas
lek
leita), ala); ak
‘ ES t—-k | 1-# a
=> ———-—- — + ‘
The fox will catch the rabbit when x = 0, and this happens when y = ak/(1 — k’).
Those cases for which the two speeds are equal or the rabbit runs faster than the
fox are taken up in the exercises.
EXERCISES 3.2
In problems I-12, find the orthogonal trajectories of the given family of curves.
. tanw =r—,
where r and @ are polar coordinates. If two curves in polar coordinates are orthogonal,
show that (Figure b)
1
tan WN = ~ tanya’
A given family of curves is said to be self-orthogonal if it has the property that its
family of orthogonal trajectories is the same as that of the given family. Verify that
the family of parabolas y* = 2cx + c? is self-orthogonal.
Solve Example 3 when the resistive force is given by Fg = mv/40. Find the limiting
velocity of the body in this case.
Solve Example 3 when the resistive force is cv. What is the limiting velocity?
Suppose a parachutist falls from rest toward the earth. Assume the chutist weighs 160
pounds and his speed is 30 ft/s at the instant the chute opens (t = 0). If the air
resistance is Fr = cv?
(a) determine the Pca v at any later time t.
(b) What is the skydiver’s limiting velocity as t>%?
(c) Calculate the limiting velocity if the air resistance is cv instead of cv?.
A parachutist weighing 160 pounds falls from rest toward the earth. Before the
parachute opens, the air resistance is equal to $v. The chute opens 5 s later, ane the
air resistance changes to 5v7/8. Find the velocity of the skydiver
(a) before the parachute opens.
(b) after the parachute opens.
(c) If the parachute never opened, what would be the limiting velocity of the
skydiver? Compare this value with the value obtained after the chute opens.
*26. A man and his parachute together weigh 192 pounds. Assume a safe landing
velocity
SEC. 3.2 / GENERAL APPLICATIONS 67
is 16 ft/s and that the air resistance is known to be proportional to the square of the
velocity, equaling 5 pound for each square foot of cross-sectional area of the parachute
when it is moving 20 ft/s. What is the cross-sectional area of the parachute necessary
for the chutist to make a safe landing?
27. If it takes time T for a ball thrown upward to reach its highest point (Fr = 0), show
that the return time is also 7. What is the velocity of the ball upon return if the initial
velocity is Uo?
28. Determine the escape velocity from the moon given that the moon’s radius is roughly
1080 miles and the acceleration of gravity is 0.165g, where g is the acceleration of
gravity on the earth’s surface.
29. Determine the escape velocity from Mars given that its radius is 2100 miles and the
acceleration of gravity is 0.38g.
30. Given that the force of gravity on Venus is about 85% of the earth’s gravity and the
radius of Venus is roughly 3800 miles, determine the escape velocity.
31. At 200 miles above the earth’s surface, the atmosphere offers almost no resistance.
What velocity should a rocket have at this altitude in order to reach a height of 4000
miles if all its fuel is exhausted at this point?
*32. If a body is shot straight up from the earth’s surface with an initial velocity vp and no
air resistance is assumed, show that the rising time t as a function of the distance r
of the body from the center of the earth is given by
39. Referring to the fox-rabbit problem in Section 3.2.4, find the fox’s path when v > w,
or k > 1. Compute the distance between the fox and the rabbit in terms of the vari-
able x.
CHAP. 3 / APPLICATIONS OF FIRST-ORDER EQUATIONS
*40. A man standing at O holds a rope of length a to which a weight is attached, initially
at the point (a, 0). The man walks along the positive y-axis, dragging the weight after
him (see figure).
(a, 0)
Problem 40
(a) Show that the slope of the path along which the weight moves is
where k = w/v.
medida |
43. In problem 42, if the wind speed and plane speed are equal, show that the
path is that
of a parabola. Will the pilot ever reach city C?
44. Find the equation of a curve that passes through the point (4,
1) and has slope
—y/(x — 3) at any point (x, y) on the curve.
45. Find the shape of a curved mirror such that light from a distant source
will be reflected
to the origin.
Hint: The slope of such a curve in the xy-plane is given by
y=
) SN EEE RE aie
y
SEC. 3.3 / APPLICATIONS INVOLVING LINEAR EQUATIONS 69
*46. Ona winter day it began snowing early in the morning, and the snow continued falling
at a constant rate. The speed at which a snowplow can clear a road is inversely
proportional to the height of the accumulated snow. The snowplow started at 11:00
A.M. and had cleared 4 miles of road by 2 P.M. Another 2 miles was cleared by 5 P.M.
At what time did the snow begin?
“47. Acable of constant density hanging from two pegs (such as a telephone line) assumes
a shape determined by the DE
ytkVl+o'y¥=0
where k is a constant. Show that the cable assumes the shape of a hyperbolic cosine,
called a catenary.
*48. (Brachistochrone problem) One of the most famous problems in mechanics is called
the brachistochrone problem. The problem is to determine the curve along which a
particle will slide (without friction) from point O to point P in the shortest time where
gravity is the only acting force. Point P is below O, but not directly beneath it. The
curve that solves the problem is a solution of the nonlinear DE (derived through
principles of the calculus of variations)
[1 + (y')]y = 2a,
where a is a constant.
(a) Solve the DE for y’ using the negative square root. (Why use the negative square
root?)
(b) Introduce a new variable ¢ through the relation
y = —2asin’t
and solve the resulting DE.
(c) By writing 0 = 2t, show that the solution of the original DE can be expressed
in the parametric form
ve alee sin 0),; y = — a(1 = cos @);
which satisfies the auxiliary condition y(0) = 0. These last equations are the
well-known parametric equations of a cycloid.
Jeg
Problem 48
First order linear DEs occur in a wide variety of applications, some of which are
discussed in the following subsections.
70 CHAP. 3 / APPLICATIONS OF FIRST-ORDER EQUATIONS
eles (17)
arises in numerous physical theories concerning either growth or decay of some
entity. For instance, Equation (17) might describe the rate at which a radioactive
substance decomposes or the rate at which temperature changes in a cooling body.
This same equation might be used to predict the population growth of certain small
animals over short intervals of time or to describe the growth rate of certain
bacteria.
The graphs of solutions of (17) are shown in Figure 3.7 for cases when k > 0
and k < 0. Since dy/dt represents the slope of y, the sign of k gives an indication
of whether the function y is increasing (k > 0) or decreasing (k < 0).
Solution If y denotes the amount of undecayed matter at any time f, then the experimental
evidence is mathematically described by the initial value problem
dy
It
hh ky, > y(0)=y,
= -) k>O0
=
where the negative sign indicates that y is a decreasing function. This is a homoge-
neous equation with general solution
y a Ges
a
y(t) = woe ™.
Se a mre ce ee
where P(t) denotes the population at any time and a is a positive constant (since P
is increasing). Let us assume the population is Po at time t = O, although we don’t
know its numerical value. The solution of this initial value problem is therefore
P(t) = Poe”.
Now when t = 5 years, it is known that P(5) = 2P), which leads to
a = =log2 = 0.1386.
Thus
72 CHAP. 3 / APPLICATIONS OF FIRST-ORDER EQUATIONS
P@ = Pye? B®
To find the time at which the population has tripled, we solve
3P) = Poe? 186
for t. Taking logarithms of each side of this expression, it follows that
0.1386r = log3
or
=] .9 yeats:
pete ie. Ot ee ee eee ee ee eee ee
dP ‘
i aPe—iDPe, (19)
where both a and b are positive constants. This equation is now known as the
logistic equation, and its solutions are referred to as logistic curves.
Although (19) still generally does not provide a very accurate model for human
population growth, it has proven quite effective in predicting the growth patterns
(in a limited space) of fruit flies, for example, and certain types of bacteria.
The solution of (19) subject to the initial condition P(0) = Pp was given in
Example 24 in Section 2.6,7 viz.,
aPo
re
OT FRG bP jee ey)
*This solution can also be expressed in the equivalent form P(t) = Po2°, which more clearly reveals
that P(t) doubles every 5 years.
+Equation (19) is also solvable by separation of variables.
SEC. 3.3 / APPLICATIONS INVOLVING LINEAR EQUATIONS
73
The general shape of (20) is illustrated in Figure 3.8 for the cases when
0 < Po < a/2b, a/2b < Py < alb, and Py > a/b. In the first two cases, the value
P = a/2b cuts the graph at a point of inflection, as can be verified by examining
the second derivative P”(t). Regardless of the value of Py, we observe that
When the initial population Py exceeds the value a/b, then P(t) approaches this
value asymptotically from above rather than from below. Finally, if the initial value
of the population is a/b, it has this population value for all t.
F = ma, (21)
where the mass m is the constant of proportionality. But since acceleration is the
rate of change of velocity, Equation (21) can also be expressed as a first-order DE,
dv
Le rept 22
(22)
In those problems for which the force is due entirely to gravity, then F = mg, and
(6) becomes simply
m— = meg, (23)
Fp
Mass
mg
Figure 3.9 Free-falling body.
liquid into which a ball is dropped. Such a resistive force is frequently proportional
to the velocity of the moving mass, i.e., Fr = cv, so that the governing equation
is then modified to
m aS OTC eee a
dt 2 ; ‘
or
dvaee
ae (24)
dimeaetit 2
The value of the positive constant c is determined by the nature of the resistive
force.
EXAMPLE 8 (Particle motion problem) A parachutist weighing 150 pounds opens his chute when
_ his downward velocity is 100 ft/s.
(a) If the force of air resistance is 25v, find the velocity of the parachutist at any later
time (prior to hitting the ground).
(b) What is the limiting velocity of the parachutist?
(c) If distance s is related to the velocity by ds/dt = —v and the parachutist opens
his parachute at 2000 feet, how close to the ground will he be after 5 minutes?
Vx. = 6 ft/s.
(c) Since the position s of the parachutist is determined from the initial value
problem
7 = 2,t (0)
Cpe: 0) =
= 2000 20 ,
where qg denotes the electric charge on the capacitor and is related to the current i
by i = dq/dt. Thus, the governing equation is
*From a practical point of view, the velocity approaches the limiting velocity in a relatively short
interval of time and remains at this value from that point on.
+Named in honor of the German physicist, GUSTAV R. KIRCHHOFF (1824-1887).
76 CHAP. 3 / APPLICATIONS OF FIRST-ORDER EQUATIONS
q
Ri i+ C—=E(p),
(t)
dag wast
—+--—qc4 = E(t).
Ro (t) (25)
For the special case when E(t) = Eo, a constant, we find that (25) has the general
solution
: Car, e de
i(tt)= - ae eI 1g Coe (27)
Solution . Substituting R = 200, C = 5 Xx 10~°, Ey = 100, and iy) = 10? into (27) we get
Newton’s law of cooling states that “the rate of change of temperature u in a cooling
body is proportional to the difference between u and the temperature Ty of the
surrounding medium.” In symbols, this law reads
d
$ =e k(us— 15); (28)
EXAMPLE 10 (Cooling problem) A metal ball is heated to a temperature of 100°C and then
immersed in water at temperature 30°C at time t = 0. After 3 minutes the tem-
perature of the ball is reduced to 70°C. Find its temperature at all later times.
Solution The governing equation is (28) with Ty) = 30°C. (It is assumed that the water can
be maintained at 30°C even with the metal ball immersed).* The solution of (28)
satisfying the initial condition u(0) = 100 is
u(t) = 30 + 70e™.
The additional information u(3) = 70 allows us to determine a value for k, 1.e.,
70 = 30 + 70e *,
or
1 t/
k = 3 log (7)
— | = 0.1865.
Time (minutes)
Figure 3.11
The mixing of two solutions can also give rise to a first-order linear DE. In the
next example we will consider the mixing of pure water with a salt brine solution.
EXAMPLE 11 (Mixing problem) A certain tank contains 100 gallons of a solution of dissolved salt
and water, the mixture being kept uniform by stirring. If pure water is now allowed
to flow into the tank at the rate of 4 gal/min, and the mixture flows out at the rate
of 3 gal/min (Figure 3.12), how much salt will remain in the tank after ¢ minutes
if 15 pounds of salt is initially in the mixture?
4 gal/min
3 gal/min
Figure 3.12
Solution Let us denote the amount of salt present in the tank at any one time by x(t). The net
rate at which x(¢) changes is given by
dx ;
¥ = rate of salt in — rate of salt out.
Since pure water is coming in, the rate of salt entering the tank is zero. The rate
at which salt is leaving the tank is the product of the amount of salt per gallon and
the number of gallons per minute leaving the tank, i.e.,
EXERCISES 3.3
c) The half-life of a certain radioactive substance is 1620 years. If 10 grams are initially
present in a given sample, how much will be left after 162 years?
2: A certain breeder reactor converts uranium 238 into the isotope plutonium 239. After
15 years, 0.043% of the initial amount of plutonium has decayed. What is the half-life
of this isotope?
3) By comparing the amount of carbon 14 found in a fossil with the constant ratio found
in the atmosphere, the age of the fossil can be estimated. Assuming the half-life of
carbon 14 is 5600 years, determine the approximate age of a fossil that is found to
contain 0.1% of the original amount of carbon 14.
A certain chemical is converted into another substance at a rate proportional to the
square of the amount of unconverted chemical. Starting at time ¢ = 0 with an amount
yo of unconverted chemical, determine the amount of unconverted chemical for all
P= 0.
The population of a certain country was | million in the year 1950, and the instanta-
neous growth rate since that time is observed to be 3% of the current population.
Assuming this trend continues, what population is expected by the year 2000?
Suppose that a certain population has grown to 10,000 after 3 years and that after 4
years the population will be double the original amount. What was the original
population, and what will be the population after 10 years if it continues to grow at
a rate proportional to the number of people present at any time?
Tf If we allow a population to change by either immigration into or emigration out of,
then the governing equation is modified to
—dP
LC =aP + mattef(0,
where f(t) is the rate that members of the population are being added or subtracted
from outside the system. Determine the population growth if a = —3 and the immi-
gration is governed by the periodic function
f(t) = 100001 + 5b sin 2).
Assume the initial population is Po, and discuss separately the cases |b| <1 and
|b| > 1; i.e., which case represents immigration and which emigration? Finally,
determine the steady-state population by considering the limit of P(t) as t>~.
The infusion of glucose into the bloodstream is an important medical technique. Let
us imagine glucose is infused into the bloodstream at the constant rate of b grams per
minute. At the same time, the glucose is converted and removed from the bloodstream
at a rate proportional to the amount of glucose present.
(a) Show that the amount of glucose G(t) present at any time is governed by the DE
0.06\“
M(t) = Mo| 1+ pF ;
where k is the frequency (number of times each year) at which the interest is ©
compounded.
(b) Determine the amount of money in the bank after 10 years ifk = 1, k = 4, and
k = 365.
(c) If the interest is compounded continuously, write the DE for M (?) illustrating this
growth of investment, and show that its solution is the same as obtained by
allowing k—> in the formula given in (a). Determine the amount of money in
the bank at the end of 10 years as predicted by this model.
In problem 9, how long will it take to double the original investment if the interest
is compounded continuously? If k = 1?
A stone weighing 4 pound falls from rest from a tall building. If the air resistance is
known to be v/160, where v is the velocity of the stone, determine v at any time.
If the building in problem 11 is 200 feet high, what is the velocity of the stone when
it hits the ground? (Approximate your answer.)
An object is thrown upward with an initial velocity vo, and the air resistance is cv.
(a) Find the time required for the object to reach its maximum height.
(b) Find the maximum height if c = 7, mg = 4, and vo = 8.
In the RC circuit shown in Figure 3.10, how long will it take the current i(f) to
decrease to one-half its original value if the voltage source is the constant Eo?
Find the steady-state current in an RC circuit when the voltage source is
E(t) = Eosin wt. Hint: Let t—~© in the solution.
“16. A variable resistance R = (5 + t)”' ohm and a capacitance of 5 x 10~° farad are
connected in series with a voltage source of 100 volts. If the initial charge g is zero,
what is the charge on the capacitor after 60 seconds?
17: The current i(¢) in a circuit containing only a resistance R and an inductance L in series
with a voltage source E(t) is governed by
di . ;
tes + Ri =E(), iO) = io,
where L and R are known constants (see figure). Solve this initial value problem when
(a) E(t)= Eo (constant).
000
Problem 17
SEC . 3.3 / APPLICATIONS INVOLVING LINEAR EQUATIONS 81
REFERENCES
Boyce, W. E., and DiPrima, R. C. 1977. Elementary differential equations. 3rd ed. New
York: Wiley.
Coddington, E. A. 1961. An introduction to ordinary differential equations. Englewood
Cliffs, N.J.: Prentice-Hall.
Derrick, W. R., and Grossman, S. I. 1981. Elementary differential equations with applica-
tions. 2nd ed. Reading, Mass.: Addison-Wesley.
Rainville, E. D., and Bedient, P. E. 1980. Elementary differential equations. 6th ed. New
York: Macmillan.
Ross, S. L. 1980. Introduction to ordinary differential equations. 3rd ed. New York: Wiley.
The study of linear DEs is of both theoretical and practical importance. In Chapter 2
we investigated the general theory concerning first-order linear equations and
presented various applications of these equations in Chapter 3. Here we wish to
build upon that theory and in some instances use it for motivational purposes for
developing the higher-order linear theory. Applications of higher-order linear DEs will
be taken up in subsequent chapters.
The notion of linear independence, which is so critical in the development of a
general solution, is introduced in Section 4.2 for second-order equations. The ques-
tion of linear independence is later reduced to evaluating a special determinant
called the Wronskian, the nonvanishing of which is established as both a necessary.
and sufficient condition for linear independence of a set of solutions of a homo-
geneous linear DE. In Section 4.3 a method is discussed for producing a second
linearly independent solution of a second-order equation given that one solution is
known. Although this technique is often considered to be mostly of theoretical
importance, it is used in practice from time to time.
The solution of second-order, homogeneous, constant-coefficient equations is dis-
cussed in Section 4.4. The significant feature here is that such equations can be
solved entirely by algebraic methods. In Section 4.5 we generalize the theory and
solution techniques to equations of order greater than 2.
We turn our attention to nonhomogeneous DEs in Sections 4.6 and 4.7. The
method of undetermined coefficients, which requires us to guess at the structural
form of the particular solution, is introduced first. Although this method is fairly easy
to apply, its application is restricted to a narrow class of DEs. A more general
method called variation of parameters is then introduced, which theoretically is
applicable to all linear equations.
In the last section we extend the solution techniques for constant-coefficient
equations to a special variable-coefficient equation called the Cauchy-Euler
equation.
82
41 INTRODUCTION
Most nonlinear equations of order greater than | are very difficult to solve. This is
because there is no general theory concerning the solution of these equations, and
even the techniques discussed in Chapter 2 are not generally applicable. We will
therefore restrict our attention in the remainder of the text almost exclusively to
linear equations.
Linear DEs of order greater than | can be applied in numerous areas. Of these
equations, the most prominent are those of the second order, and so we will direct
most of our theory development and applications at them. For example, the forced
and free oscillations of a spring-mass system are governed by the second-order
linear DE
Theorem 4.1 If y = y,(x) is a solution of the homogeneous DE (5), then so is y = Cy,(x) for
any constant C;. More generally, if yi, y2, . - -» Ye are all solutions of (5) on some
specified interval, then
83
CHAP. 4 / LINEAR EQUATIONS OF HIGHER ORDER
Proof: For notational simplicity we will prove only the case when n = k = 2.
Let y, and y, be two solutions of
= Ci: 0 37C.20
= Ole)
In the development of the theory of linear DEs that follows, we will generally
present the theory for second-order equations first and later generalize the results
for equations of order n. Unfortunately, even for second-order linear DEs no
general solution techniques are available, unlike the case for first-order linear
equations. Only for certain narrow classes of equations can formulas for general
solutions be found.
In order to SENCLED S
some oo the theory for second-order DEs, it is helpful to put
(6)
(7)
A,(x)/A2(X), do(x) =
AsG)Aa(x), and FQ) = FOD/As(,
dn Sa 1
(9)
SEC. 4.2 / LINEAR DEPENDENCE AND INDEPENDENCE
EXAMPLE 1 Show oe yi = x and y)= 5x are linearly dependent on any interval while y, and
y3= x? are linearly independent on any interval.
y3(x)
=x # constant,
yi (x)
which implies that y, and y; are linearly independent.
EXAMPLE 2 Discuss the linear independence of y,; = x and y, = |x| over the intervals
Q=x <™~ and —~ < x < & (see Figure 4.1).
Vi 25 V2 = |x|
Figure 4.1
Thus, since the ratio y2/y, remains constant over the interval 0 = x < %, these
functions are clearly linearly dependent on this interval. However, over the larger
interval —0%© < x < ©, the ratio y2/y, changes values (from —1 to +1), which
implies linear independence on this interval.
Theorem 4.2 | If y, and y, are linearly independent ons of the homogeneous linear equation ©
86 CHAP. 4 / LINEAR EQUATIONS OF HIGHER ORDER
Solution Note that y, and y2 are linearly independent on any interval, since
ya(x) =
e ** # constant,
yi (x)
and hence the general solution is
y = Cie* + Coe”,
where C, and C, are arbitrary constants.
4.2.1 Wronskians
When more than two functions are involved, our definitions of linear dependence
and independence do not apply. Thus we seek a more general definition of these
terms which can be applied to situations involving more than two functions.
Observe that if constants C; and C, (not both zero) can be found such that
C
yi) C220):
which shows that y, and y2 are proportional and hence linearly dependent.
Now suppose we consider the expression
*For a proof of Theorem 4.2, see E. L. Ince, Ordinary Differential Equations (New York: Dover,
1956).
SEC. 4.2 / LINEAR DEPENDENCE AND INDEPENDENCE 87
If we think of C, and C, as the unknowns in (11a) and (11b), then using Cramer’s
rule we see that nonzero values for C, and C; are possible only when the coefficient
determinant of these two equations is zero, i.e., when
yitx) ya(x)
yi(x) y2(x) Vash//s-/
I STK CICC &.
(12)
This coefficient determinant is called the Wronskian , named after the Polish math-
ematician J. Wronski* and denoted by
yix) yo(x)
| W(y,
Yis 2 y)@)= ae
|", :
yi(x) = yilx)y,(x)y2(x)
y2 (x) — yi
yi(x)y2(x).
(x)a(x) 13
(13)
‘a on \Y\>
et
( Theorem 4.3 If y, and y possess first derivatives on some interval J, and
8
“——— (a) if W(y1, y2) (x) # 0 for at least one point in /, then y, and y> are linearly
independent.
(b) ify, and y, are linearly dependent on /, then W(y,, y2) (x) = 0 for all x in J.
EXAMPLE 4 Show that y, = cosx and y,= sinx are linearly independent.
EXAMPLE 5 Show that y, = e”" and y. = e”™ are linearly independent if m, # mp.
We should be aware that the conditions listed in corte =oaare only ene
{Dhara That iiS to) Say,
not may Ca! yen when —
in
*JozeF M. H. WRONSKI (1778-1853) studied mathematics in Germany but lived most of his life
France. His sole lasting contribution to mathematics appears to be the Wronskian determinant.
88 CHAP. 4 / LINEAR EQUATIONS OF HIGHER ORDER
EXAMPLE 6 Show that the functions y, = x? and y, = |x|’ are linearly independent on the
interval —%© < x < ©, but that the Wronskian of y, and yp is identically zero (see
Figure 4.2).
peo eS
W(y1, y2) (x) i Le 3x2 <a 0,
Because our interest is confined to only those sets of functions that are solutions
of linear DEs, an additional restriction can be imposed on any given set of func-
tions, which, together with the nonvanishing of their Wronskian, imply linear
independence. The only requirement is that the functions involved be solutions of
a homogeneous linear DE in order that the nonvanishing of the Wronskian becomes
both necessary and sufficient for the functions to be linearly independent. The proof
of this result relies on the following important lemma.
SEC. 4.2 / LINEAR DEPENDENCE AND INDEPENDENCE 89
Lemma 4.1 |
£
oy Worn)
Visey2) ie=iyLonrys
Yi¥2 —— Wiyi )ya) == Viy2
yh —— Yiyfyr,
and since ty = —a,(x)y’ — ao(x)y, we can rewrite this expression in the form
w'+ acowe =n wy
with general solution (5 fuction af,Horo agn)
By ey ary a0)
to within a multiplicative constant.
y' +/-y_
” 1 '
+y = 0.
x
Using Abel’s formula, we then have
*NeILS H. ABEL (1802-1829) was one of six children born into a poor Norwegian family. One of his
early accomplishments was proving that the general fifth-degree algebraic equation has no radical
solution. Stricken with tuberculosis in 1827, he died two years later at the age of 26.
CHAP. 4 / LINEAR EQUATIONS OF HIGHER ORDER
dx C
W(y1, Ya) (x) = cexp(-[©) Sane
Proof: First, if W(y1, y2) (x) # 0 for every x in J, it follows from Theorem 4.3
that y, and y2 are linearly independent. Conversely, if y, and y2 are known to be
linearly independent, then from Lemma 4.1 their Wronskian is given by
From Theorem 4.4 it is clear that when y, and y, are solutions of a second-order
homogeneous linear DE on some interval J, either their Wronskian is identically
zero or is never zero on J. Moreover, if the Wronskian is identically zero on J, it
now follows that the functions y, and y, are linearly dependent.
EXERCISES 4.2
In problems 1-5 , determine whether the given functions are linearly dependent or
independent .
\fi) Vumale ype 2. yp =x", yo =x"!
6. Show that the functions y, = x* and y2 = x|x| are linearly dependent on any interval
for which either x > 0 or x < 0, but are linearly independent on —2%2 < x < ~,
&@) Verify that y; = sin3x and y2 = cos 3x are solutions of
y eye= 0
and show that their Wronskian satisfies Abel’s formula (Lemma 4.1).
In problems 8-11 , use Abel’s formula (Lemma 4.1) to determine the Wronskian to
within a multiplicative constant of the solutions of the given DEs.
8. y"—4y'
4y+=0 @) y"-3y' +2y =0
10... Cl = x*)y" = 2xy" + nin Dy 0 41) XH" + xy’ + (x? — ny =0
SEC. 4.3 / CONSTRUCTING A SECOND SOLUTION FROM A KNOWN SOLUTION
91
: . eT ee _— Ce ae
It is a curious fact that given one nontrivial sol f a d-order me
linear DE,
a second linearly inc
y’ + ay’ + aay = 0,
then
exp (—[asada]
yo = yi) av err
ae ipa, 4 Ria
f dx) S ee1)Sa
re
© (y-e™
=O, e%4Ca pom (Ca Cee ao*
EXAMPLE 9 _ Given that y, = x“! is a solution of
Boy A Bx i ya Oe
find a general solution valid for x > 0.
y
tee) res.
sey ae 20 = 0.
The utility of Theorem 4.5 is clearly limited to those situations wherein one
solution of the DE is known. In most situations it is just as difficult to produce one
solution of a DE as it is to produce both, and so the theorem is of little help.
However, occasionally one solution of a DE is obtained by “inspection,” a series
method, or some other means, and the theorem can then be very useful in the
construction of a general solution.
EXERCISES 4.3
which is simply the first n + 1 terms of the Maclaurin series for e*.
“17. By assuming y2 = u(x)yi(x) is a second solution of
y" + alx)y’ + ao(x)y = 0,
given that y, is a known solution,
(a) show that the function u satisfies
(x
te ae |)(C8) oF 2H | =0
yi(x)
(b) Let v = w’ and solve the resulting first-order DE in v to obtain the result
CHAP. 4 / LINEAR EQUATIONS OF HIGHER ORDER
v(x) = age
Ae |-
-{ ao ar|.
(c) From (b), obtain an expression for u(x) and verify that y2 = u(x)yi(@) is the same
solution as given in Theorem 4.5.
Differential equations with constant coefficients are the easiest class of linear
equations to solve by any general technique. This particular fact coupled with the
fact that they occur in such a wide variety of physical applications accounts for the
special place that these equations occupy in the general theory of linear equations.
We have previously found that
thefirst-orconstant-coefficient
der linear DE
y! + ay = 0)
has the exponential solution y = Cje~“ valid on the interval —-* <x < &. Be-
cause of the special property associated with derivatives of the exponential func-
tion, it may seem natural to inquire as to whether higher-order linear DEs with
constant coefficients also exhibit exponential solutions.
For example, suppose we consider the second-order equation
ay" + by' + cy = 0, (14)
where a, b, and c are constants. Let us assume that (14) has an exponential solution
of the form
ier ere
for some value or values of the parameter m. Direct substitution of this function into
the left-hand side of (14) leads to
Solution This time the auxiliary equation is m? — 4m + 4 = 0 with the double root m = 2.
Hence,
where C; and C, are any constants. We usually relabel the constants as C; and C,
once again and write
y =e (C,cosqxoog + Cysingx) (21)
as a general solution. That y; = e” cos qx and y, = e”* sin qx are linearly indepen-_
dent solutions of (14) will be verified in the exercises.
y=i(3i— 2x)e™
SEC. 4.4 / HOMOGENEOUS SECOND-ORDER EQUATIONS WITH CONSTANT COEFFICIENTS
97
'
oe SS eg ee a
EXAMPLE 14 Find general solutions of
y"+k’y =0 and y"—k’y =0.
Solution The first DE has the auxiliary equation m? + k? = 0 with pure imaginary roots
m, = ik and m, = —ik. Thus the general solution is
y = C,coskx + Cysinkx.
In the second case the auxiliary equation is m? — k? = 0, with m, = k and
m2 = —k the roots. Its general solution is therefore
Ye Cer air Gea.
and these are linearly independent functions (see problem 5, Exercises 4.2), it
follows that ;
y = C3coshkx + C,sinhkx
However, using a device similar to that used in Case III for obtaining a general
solution in terms of trigonometric functions, we can write the general solution of
this problem in the alternate form
1
ines eat? E cosh (55) + Cysinh (5v5:)].
the verification of which is left to the reader. (See problem 20 in this section.)
98 CHAP. 4 / LINEAR EQUATIONS OF HIGHER ORDER
< In Chapter 2 we introduced the notion of an operator for the first time, the
simplest one of which is D, where Dy = y’. We now consider the second-order
differential operator
P(D) = aD? + bD +c (22)
associated with the constant-coefficient equation ay” + by’ + cy = 0. The inter-
esting feature of the operator P(D) is that it can be manipulated according to the
basic rules of algebra applied to any polynomial. To see why, let us apply the
operator (22) to the exponential function y = e”™, which leads to the relation
P(D)[e™] = (am? + bm + c)e™ = P(m)e™
Hence P(D) and P(m) have the same polynomial form and therefore factor exactly
the same.
The above remarks suggest that the constant-coefficient DE
(2D? —"D —3)y =0
can also be written in either of the equivalent forms
oeet ied — 3)y =0
or a Pk c= 51)
= (2D? — D — 3)y
and
EXERCISES 4.4
For problems 1-6 , show that the auxiliary equation has distinct real roots and find
the general solution.
@ 3y"—y' =0 2). y" — 4y =0
For problems 7-12 , show that the auxiliary equation has a repeated root and find
the general solution.
For problems 13-18, show that the auxiliary equation has complex roots and find
the general solution.
5)
2 8
cos
9 = —= and sin@d
= ——.
V 13
Hint: See problem 21.
In problems 31-35, find the general solution. Assume that x is the independent
variable.
G6) y"-4y'
+13y =0, yO=-1, y'@=2
(37) y"- 2y'—3y=0, y)=0, y'@=-
(38) y"+ ky =0, yO =yo, y'O=
*39,, y"— 3y'+ 2y=0, y(1I)=0, yQ)=
\40. (0?7-D-6y
=0, yO=0, yo) =e?
Although second-order linear DEs are far more prevalent in practice than equations
of higher order, there are applications wherein the mathematical model demands a
DE of order greater than 2. For example, in studying the small deflections of a beam
supporting a distributed load, we find that the governing DE is fourth order, as is
the DE describing the buckling modes of a long, slender column under an axial
compressive force.
In this section we wish to extend the theory of second-order linear equations
developed in previous sections to higher-order linear equations. Much of the theory
is simply a natural generalization of that developed for second-order DEs, so we
will mostly state appropriate theorems without providing their proofs. In discussing
the results for these higher-order equations, we find it convenient once again to put
the DE in normal form, but this time we will also introduce the concept of a
differential operator. Assuming A,,(x) does not vanish anywhere on the interval of
interest J, let us divide the equation
Extending our notions of linear dependence and independence from Section 4.2.1,
we get the following definition.
Definition 4.1
ioe ys (29)
EXAMPLE 17 Show that y; = 3x* — 8x, y. = x”, and y; = 4x are linearly dependent on any
interval.
Solution We see that y, = 3y, — 2y; for all x, and thus it follows that the functions are
linearly dependent.
Theorem 4.6 | If M is a normal differential operator of order n on an interval J, and if y,, y»,.
-» Yn Constitute a set of linearly independent solutions of MLy] = 0 on J, then
the general solution of this homogeneous equation on J is
EXAMPLE 18 Discuss the linear dependence of y; = 3x? — 8x, y. = x, and y; = 4x, using the
Wronskian.
0
Nonetheless, ¥
2 = i I ( ident.
Example 17, however, we did establish that these functions are indeed linearly
dependent. )
Abel’s formula (Lemma 4.1) can also be extended to nth-order equations, which
in turn can be used to prove Theorem 4.8 below.
Lemma 4.2 | (Abel’s formula) If y,, y2, . . . , y, are linearly independent solutions of M[y] = 0
on the interval J where M is the normal linear operator
; ra D'+ ano tee e + a,(x)D as a(x),
Theorem 4.8 . » +Yn are solutions of M[y]= 0 on the interval J where M is normal
a
TE TTOGERATTE of order n, then the set of solutions is linearly independent
on J if
and only if.
Wor.) y)Q) + 0
| for
everyx in J.
which is the auxiliary equation associated with (31). We obtain (32) by putting
y = e”™ into (31) and simplifying. Because of the many possible combinations of
roots of (32) depending upon 2, it is difficult to identify all cases when n > 2.
However, the following generalizations are fairly easy to establish:
eee
se ea eee
(m — 1)°(m + 2)°(3m — 2) = 0,
with roots m = 1, 1, 1, —2, —2, 3, obtained by inspection. Hence the six solutions
corresponding to these values of m are
MigesGes aye = Ae sys =e, =e ys = xe“. yem en
providing us with the general solution
y = (Ci + Cox + Cax*)e* + (Cy + Cexle~™ + Cre?*?,
acre e eee 1 i eg th ee ae
EXAMPLE 20 Find the general solution of the eighth-order DE
— 2m + 5)? =0
m*(m?
with roots m = 0, 0, 0, 0, 1 + 2i, 1 + 2i, leading to the solutions
De Cy a0 1 Cox. 4 Cax e?[ (Cs Cox) cos 2x24 (C, + Cex) sin 2x].
When the DE does not have the operator in factored form (which is the normal
situation in practice) as in Examples 19 and 20, finding the roots of the auxiliary
equation can be the most difficult part of solving the DE. According to the general
theory of polynomials, if (32) has a rational real root of the form m, = p/q, where
p and q are integers, then p must be a factor of dp and q a factor of a,. Of course,
once we have found one root m,, we can divide (32) (either directly or by using
synthetic division) by the factor (m — mj) to obtain a polynomial of one less degree
from which to determine the remaining roots.
which shows that the remaining roots are m, = m3 = 5. Thus, the general solution
is :
EXERCISES 4.5
In problems I—5, determine whether the given functions are linearly dependent or
independent.
In problems 9-29, find the general solution. Assume that x is the independent
variable.
This situation can be generalized to linear equations of any order n as the next
theorem states.
Proof: For simplicity, we will present the proof only for the case when n = 2.
Suppose that yp is a particular solution of the second-order equation
*The function yx is frequently called the complementary solution and is also denoted by
the symbol
yc in many texts.
SEC. 4.6 / THE NONHOMOGENEOUS EQUATION 109
Y= yp = Cy) 'Coy2@):
Transposing, we obtain
To illustrate the gist of the method, suppose we wish to find a particular solution
of
y" + y = 3e*. (33)
QA + Axe = 3x7,
which cannot be satisfied for any choice of the constant A. The problem is that the
derivatives of Ax* produce new functions that are linearly independent of it. To get
an idea of what to do, we observe that (34) can be transformed into a homogeneous
DE by taking three derivatives of each side. That is, (34) becomes
yO + "= 0 (35)
since d°/dx?(3x*) = 0. Now the auxiliary equation associated with (35) is
m> + m? = 0 with roots m = 0, 0, 0, +i, so that its general solution can be
expressed as
y =C, + Cox + Cax? + a
Qe ee
Cycosx + Cssinx. (36)
yp Yu
It can be argued that every solution of (34) is also a solution of (35), and since
yy = Cycosx + Cs sinx
is the homogeneous solution of (34), it follows that the particular solution of (34)
gives
yp = Ax? + Bx +C
for some choice of the constants A, B, and C. The substitution of this yp into (34)
gives
DAS TrAL@
IBY + 1Ce—sKs
or
(QAs-2 OC) Bx Ax base
By equating like coefficients in this last identity, we have
2A +C=0
yp = 3x? — 6,
The genera] rule illustrated here is that yp should have the basic structure of the
nonhomogeneous term f(x), plus all linearly independent derivatives of f.
Another difficulty in the method arises when f(x) is composed of a function
that
occurs in the homogeneous solution. For example, suppose we wish to solve
SEC. 4.6 /THE NONHOMOGENEOUS EQUATION 111
yp = x(Acosx + Bsinx),
Yp Xx .
The method of solution is summarized in Table 4.1. Using the table consists of
the following four steps.
My] = fix),
My] = f-().
Step 3: For each f;(x), assume the form of yp given in conditions (a) or (b),
depending upon whether the specified value of m is a root of the auxiliary equation.
Step 4: Add all yp’s found in Step 3 together to form the proper yp for the
original DE.
Table 4.1 Method of Undetermined Coefficients
a. m#0 = 0, k times
FQ) = be™
b. m =c,k times
ye = Ax*e™
a. mF#p
= ig b. m =p + ig, k times
ee ee eee
Solution The roots of the auxiliary equation are m = —1, —1. In order to determine yp, we
first note that if m = —1 were nota root of the auxiliary equation, we would select
yp = (Ax? + Bx + Che™
as suggested in Category III in Table 4.1. But since m = —1 is a double root of
the auxiliary equation, the proper form to assume for yp is
yp = x*(Ax? + Bx + C)e™ = (Ax* + Bx? + C*x%)e™*
Solution By inspection, we see that the auxiliary equation has roots m = 0, 0, 1. The
homogeneous solution is therefore
a ae ge Cote Gg. Cae.
yp = Acosx + Bsinx.
Corresponding to the term —Se*, however, we need to write
Vian (Ce),
since m = 1 is a root of the auxiliary equation (see Category II). Thus, we write
the complete particular solution as
yp = Acosx 2 Bsinx + Cxe*.
Vo yest Ya
=cosx + sinx — Sxe* + C, + Cox + Ce".
Solution The roots of the auxiliary equation are m = +i, and thus
Yo = C,cosx + Cysinx.
Since m = +i are roots (one time), the proper form to assume for our particular
solution is
yp = x[(Ax + B)cosx + (Cx + D)sinx]
= (Ax? + Bx)cosx + (Cx? + Dx) sinx.
The substitution of yp into the DE yields
yp + yp = —(Ax? + Bx)cosx — (44x + 2B)sinx + 2Acosx
— (Cx? + Dx)sinx + (4Cx + 2D)cosx + 2Csinx
+ (Ax? + Bx)cosx + (Cx? + Dx) sinx
= 4Cxcosx + 2(A + D)cosx — 4Axsinx + 2(C — B)sinx
= xcOSx — COSx.
Equating like coefficients gives
4C =1
24 +2D = -1
—4A =0
2G.— 2B = 0;
from which we deduce A = 0, B = j, C = 4, D = —4. The general solution is
therefore
YS yp tye
=a 1
= 4x’sinx + 4xcosx — 4xsinx + C;cosx + C) sin x.
. . .
EXERCISES 4.6
y” + 4y = 2sinxcosx
Q)
— 7 10. y" —3y’ — 4y = 30e*
0 219)
y" + ay" + 4y' = xe™ “18. 2y” — 3y” — 3y’ + 2y = 4cosh?x
l6y — y = 6e*? 20. y® — 4y" = 8e-* + 3e7 -— x +8
In problems 21-24, set up the correct form for yp but do not solve for the
coefficients.
e4) y” + 4y = 3 + le*cosx
When f(x) is not of the form assumed in Section 4.6.1, or when the DE has variable
coefficients, a more general method of constructing yp is required. This more
+
116 CHAP. 4 / LINEAR EQUATIONS OF HIGHER ORDER
is known to be
u'(x)y(x) + v'(x)y2(x) = 0.
Thus,
= f(x).
Therefore,
yilx) yo)
yi) y2(x)
does not vanish. But since this determinant is precisely the Wronskian of the
linearly independent solutions y, and y2, it can never be zero on the interval
a =x <b, and so the simultaneous solution of (44) and (45) yields
yO ney) | 47
-10) |Fo, pent * 1 Wor ynen ee
In certain instances these integrals cannot be evaluated explicitly, and so we must
leave yp in the integral form suggested by (47).
1 Pee
y” + 9y = 5 esc 3x. fe aon e
1
u'(x) = -: sin 3x csc 3x = We
and
ere toe _ COS 3%
v'(x) = 6 cos 3x csc 3x = Aaa
cy ih
—
6 608 3x
a +
ie log |sin 3x| sin 3x,
s
and
c.. E
u(x) ee
Fearee
hr =
log(1 + e~*)ail =x
and
an e a e
0) = [aaa = |(6 ae
or
EXERCISES 4.7
Ue 25 yl ay =x
3. y" + 9y = sin3x 4. y"+y =cscx
Bx y" + y =secx 6. y"+y =sec*x
\7/ y" +y = sec*x 8. y"+y =tanx
(9) y+ y = cotx 10. y"’+y =tan’x
14"y" + y =secxescex 12. y"+y =cscxcotx
yaa 3x
f ) iD , a2 x * 1B fs ’ =
13, y”" — 4y' + 4y = + De 14 ry Bye Fey. crraee
"3 2x
19.) by 2y y= e 7 logx,, x20 *16. Ve area
Vax
*29,. (1 —x?)y” — 2xy’ = 2x; yw = Ci + Clog she
yp = i,sin(x — s)f(s)
ds
yp = iksinh(x — s)f(s)
ds
Thus far we have solved only constant-coefficient equations although most of the
theory developed in the early sections of this chapter is applicable to general linear
equations with variable coefficients. Unfortunately, we usually cannot solve these
general linear equations as easily as constant-coefficient equations. That is, to solve
variable- coefficient equations one must ass resort to some sort of
and
(49)
121
122 CHAP. 4 / LINEAR EQUATIONS OF HIGHER ORDER
Solution This time the DE is nonhomogeneous, and so we must find a particular solution yp
(by variation of parameters) as well as the homogeneous solution y,. The homoge-
neous DE x*y”" — 3xy’ + 3y = 0 transforms into
d*y
Se dy
ae - =
dt? dt eae
by making the change of variable x = e'. Thus the roots of the auxiliary equation
are m = 1, 3, giving us
W Gx =
and thus
SEC. 4.8 / THE CAUCHY-EULER EQUATION 123
: x3 2x2e% 2x? x
EXERCISES 4.8
In problems 1-7, find the general solution of each homogeneous DE valid for
hee Ue
11. — 4xy' + 6y = 4x — 6
x?y" 12) xy ty ty = 4xlosx
130 x2y xy > y= x 14. — 2xy’ + 2y = x°e*
x2y"”
(15 x2y" — 2xy’ + 2y = x? logx?
In problems 21-23, solve the equation by assuming a solution of the form y = x",
where m must be determined.
Gr) 2x*y"”
+ 3xy'’ -— y =0 22; x2y"
— 2xy' + 2y =0
23. x2y" + Txy’ + Sy =0
*24. Show that under the transformtion x = e’,
a) meDD SAND ey,
xty = D(D: = 1KD-— 2)D —-3)y:
where D = d/dt, and deduce that in general
x"y™ = D(D — 1(D —2)---(D-—n + Dy.
In problems 25-30, use the result of problem 24 to find the general solution of each
DE Vor. 0:
REFERENCES
Boyce, W. E., and DiPrima, R. C. 1977. Elementary differential equations .3rd ed. New
York: Wiley.
Coddington, E. A. 1961. An introduction to ordinary differential equations .Englewood
Cliffs, N.J.: Prentice-Hall.
Ince, E. L. 1956. Ordinary differential equations. New York: Dover.
Rabenstein, A. L. 1972. Introduction to ordinary differential equations .2nd ed. New York:
Academic Press.
Rainville, E. D., and Bedient, P. E. 1974. Elementary differential equations .5th ed.
New
York: Macmillan.
REFERENCES
125
Ross, S. L. 1980. Introduction to ordinary differential equations. 3rd ed. New York:
Wiley.
Zill, D. G. 1979. A first course in differential equations with applications .Boston, Mass.:
Prindle, Weber & Schmidt.
In the present chapter we consider applications of DEs to problems involving me-
chanical vibrations and electric circuits. Such applications naturally lead to initial
value problems, since time is normally the independent variable and the auxiliary
conditions are all prescribed at a particular instant of time.
Newton’s law of motion is used in Section 5.2 to derive the DE governing the
small vertical movements of a spring-mass system. Free motions are discussed here
for cases of both undamped and damped systems. The latter type of motion results
when frictional or other types of resistive forces in the system are taken into account.
The three cases of overdamped, underdamped, and critically damped motions are
carefully distinguished, and typical solution curves are illustrated for each case. In
Section 5.3 we examine certain kinds of forced motions resulting from an external
stimulus applied to the system. In particular, we consider the case of an impressed
periodic force whose frequency is at or near that of the natural frequency of the sys-
tem, leading to a state of resonance. Analogous systems involving simple electric
circuits are briefly discussed in Section 5.4.
In Section 5.5 we introduce the notion of a Green’s function, which permits us to
develop general solution formulas for initial value problems similar to the formula de-
rived in Section 2.5.1 in connection with first-order equations. The concept of an im-
pulse function, which is so useful in circuit analysis, is presented in Section 5.6, and
this in turn helps to describe a physical interpretation of the Green's function.
In Section 5.7 we look at equations with variable coefficients, which arise in appli-
cations where some of the system parameters vary over time. Since solution tech-
niques for such DEs have not yet been developed (see Chapter 9), we limit our
study to the qualitative behavior of the solutions. We are particularly interested in the
oscillatory characteristics of the solutions, which are discussed in terms of the fa-
mous Sturm separation and comparison theorems.
Fora shorter course, Sections 5.5, 5.6, and 5.7 can be omitted.
126
5.1 INTRODUCTION
Initial value problems arise in the study of particle motion, population dynamics,
and electric circuits, as well as several other areas of application. The general
problem is to solve the linear equation
Remark. Since the independent variable in initial value problems is usually time,
it is customary to use f rather than x.
Theorem 5.1 If f(t) is continuous on the interval t = f and M is a normal differential operator
of order n on this same interval, then for any choice of the constants ko, ki, . . . ,
k,-1, there exists a unique solution of the initial value problem
Corollary 5.1 If M is anormal differential operator of order n on the interval t > f, then the initial
value problem
M[y]=0, t>h,
127
128 CHAP. 5 / APPLICATIONS INVOLVING INITIAL VALUE PROBLEMS
When different weights are attached to an elastic spring suspended from a fixed
support, the spring will stretch by an amount that varies with the weight. Hooke’s
law* states that the spring will exert an upward restoring force proportional to the
amount of stretch s (within reason); i.e., F = ks. The constant of proportionality,
denoted by k, depends upon the “stiffness” of the spring and thus is different for
each spring. For example, if a 10-pound weight stretches a spring 6 inches (3 foot),
then 10 = k(3), or k = 20 lb/ft, whereas if the weight only stretches the spring 2
inches, we find k = 60 lb/ft.
The two most common systems of units and their abbreviations are given in
Table 5.1.
Suppose the natural length of a spring is b units and a weight W = mg is attached
to the spring. The weight, which is also referred to as the “mass,” will then attain
a position of equilibrium at y = 0, which is s units from the equilibrium position
of the spring itself (see Figure 5.1). The upward restoring force is ks, which is offset
by the weight mg. If the system is now subjected to an external force (downward)
of magnitude F(1), the weight will move in the vertical direction. Let us assume
such motions are “small” so that Hooke’s law will remain valid.
In addition to an external force, there frequently exists a retarding force caused
by resistance of the medium in which the motion takes place or possibly by friction.
For example, the weight could be suspended in a viscous medium, connected to a
(a) (b)
Note: To convert from one system to another, we use the following relations:
1 N = 1 kg-m/s? = 0.2248 Ib 1 kg = 0.0685 slug
1 m = 3.2808 ft 1 Ib = 1 slug-ft/s? = 4.4482 N
dashpot damping device, etc. In practice, many such retarding forces are approxi-
mately proportional to the velocity y’. Hence, we will assume the resistive force
is cy’, where c is a positive constant, and this force acts in a direction opposing the
motion. Now, taking into account all forces acting on the system, we deduce
im REGFY Nes a WS— am is a a 9) [ ;
Remark. Note that the terms on the left-hand side of (4) represent system forces
such as restoring and damping forces, while the function F (¢) on the right-hand side
represents an external force to the system. In this sense, the function F (ft) is referred
to as a forcing function or input function, which gives rise to additional motions of
130 CHAP. 5 / APPLICATIONS INVOLVING INITIAL VALUE PROBLEMS
the system superimposed onto the free motion that would result in the absence
of F(t).
EXAMPLE 2 Derive the initial value problem for a 2-kg mass suspended by a spring with spring
constant k = 32 N/m. A force of 0.5 sin 3t is applied to the mass, and a dashpot
damping mechanism is such that c = 5 kg/s. The mass is released from rest 3 cm
below the equilibrium position.
For the case when c is sufficiently small compared with mk and the time span is
short, it may be acceptable to neglect the damping term cy’. (All systems have a
certain amount of damping, no matter how small the motions or how short the
period of time.) If this is done and if no external force acts on the mass, then the
initial value problem whose solution describes the motion is
my + ky= 0,=-t> 0, yO) =x, y'(0)i= wp. (6)
The general solution of the DE above is
y = C\ COs wot + Cr sin wot,
where we write w) = Vk/m for convenience. If we subject this solution function
to the prescribed initial conditions, we see that
Uo.
Y = YoCOS @pt + — sin Wot. (7)
Wo
A = 4/98 + ey (9)
Wo
is called the amplitude of the motion. It gives the maximum (positive) displacement
of the mass from its-equilibrium position. The angle ¢ is referred to as the phase
angle and is chosen in such a way that
cos d =7
tang = te ; (10)
sing =—~ 4
WA
Any motion described by a single sinusoidal function as in (8) is called simple
harmonic motion. Such motion is clearly periodic, since the mass will oscillate
between y = —A and y = A. The time between successive maxims, or the length
of time required to complete one cycle of the motion, is the period of the motion
and is given by
20
Wo
Yo A
——_> t
*The unit cycles per second (cps) is now called hertz (Hz).
132 CHAP. 5 / APPLICATIONS INVOLVING INITIAL VALUE PROBLEMS
cosine curve will appear to shift a certain amount along the t-axis and possibly
change amplitude.
EXAMPLE 3 Find the natural period of a spring-mass system for which the spring is stretched 4
in. by a 6-lb weight.
6=k--,|
3
or k = 18 lb/ft. The mass m = W/g= % slug, and hence
m\ 1/2 6 ve = aV6
T= 2n() = an aim Fie ch
EXAMPLE 4 Suppose a 16-lb weight is attached to the spring in Example 3 and released 3 in.
above the equilibrium point of the spring-mass system with an initial velocity of 2
ft/sec directed upward.* Describe the subsequent motion of the mass.
Solution In Example 3 we found that k = 18 lb/ft, and the 16-lb weight corresponds to a
mass of 3 slug. Thus the motion of the mass is described by the solution of the initial
value problem
Solving, we get
1 1
y=- cos 66 — 3 Sin 6r.
Putting this solution in the form of Equation (8), we find the amplitude of motion
given by
whereas the phase angle satisfies cos@ = —3, sind = —!, or d = 4.07 rad
(233°). Hence,
2
Va
D cos (6¢ — 4.07)
*The positive y-axis has been chosen downward so that, for example, y(0)
= —Yo if the mass is pushed
yo units above the equilibrium position. The same is true of the velocity.
SEC. 5.2 / SMALL MOTIONS OF A SPRING-MASS SYSTEM 133
Sometimes it is useful to know the values of time for which the graph of y(t)
crosses the positive f-axis. This corresponds physically to the mass passing through
its equilibrium position. Writing the solution in the form of Equation (8) is very
helpful for such calculations. For instance, using the solution in Example 4, we
observe that cos (6t — 4.07) = 0 when
(27271) as
6t — 4.07 = ee
Z
where n is an integer. The first positive value of ¢that satisfies this relation is found
to be tf; = 0.42 rad (n = 0), whereas the next value is t, = 0.94 rad (n = 1), and
so on (see Figure 5.3).
Figure 5.3
A mass m is suspended from the end of a rod of constant length b whose weight
is negligible (see Figure 5.4). We wish to determine the equation of motion of the
mass in terms of the angle of displacement @.
Summing forces makes it clear that the weight component mg cos 6, acting in the
normal direction to the path, is offset by the force of restraint in the rod. Therefore,
Figure 5.4
134 CHAP. 5 / APPLICATIONS INVOLVING INITIAL VALUE PROBLEMS
the only weight component contributing to the motion is mg sin 6, which acts in the
direction of the tangent to the path. If we denote the arc length of the path by s,
then Newton’s second law of motion reads
Uke
us kon meg sin 0 ;
dt? °
where the minus sign illustrates that the tangential force component opposes the
motion for s increasing. Now the arc length s of a circle of radius b is related to
the central angle 6 through the formula s = b@, and so the equation of motion (after
simplification) transforms to
2
d‘0
bt 8 sing = 0. (125
The equation of motion (12) is nonlinear due to the term sin 0. To solve this
equation exactly would necessitate introducing a special function called the Ja-
cobian elliptic function ,* since the equation has no solution that can be expressed
in terms of elementary functions. However, if we restrict the motion so that 6 is
always a “small angle,” then we might use the approximation sin 9 = @ and replace
(12) with the linear DE
d*6
b—>; + 2g =0. 13
IY ty ee ky =O, = (14)
The auxiliary equation from which the solutions of (14) are determin
ed is
mA* + ct +k = 0,
with roots
—c + (c? — 4mk)!”
XA, a0) = a
(13)
Case I—Overdamping: The damping is large compared with the spring con-
stant. Both roots of the auxiliary equation are real and distinct, leading to the
solution formula
: t
y= corny exp Gc ~ ami)? |+ C,exp |-« _ amy? || (16)
Case II—Critical damping: For this case the roots of the auxiliary equation are
equal, so the solution takes the form
wasyen UC hrr.Cof): (17)
The motions here are similar to those of the overdamped case, as shown in Figure
oO:
Vo = initial velocity
yo Vv) > O
> <x
ae
Case III—Underdamping: This case is the most interesting of all three cases.
The roots of the auxiliary equation are complex, and so the solution can be
expressed as
y a3 (CAN OF cos pt + C sin Lt),
(18)
In all three cases, the solution of the homogeneous DE (14) contains the multi-
plicative factor e~’””", which tends to zero after a sufficiently long period of time.
136 CHAP. 5 / APPLICATIONS INVOLVING INITIAL VALUE PROBLEMS
Taw
Hence, regardless of the prescribed initial conditions, the solution of (14) must
approach zero as time goes on. This situation confirms what we intuitively expect—
without damping (friction), the motion of the system continues indefinitely; but
with damping, the motion eventually dies out.
a es eee eee
EXAMPLE 5 A spring-mass system involves a mass of 4 kg, a spring with k = 64 N/m, anda
dashpot with c = 32 kg/s. The mass is lowered 1 m from its equilibrium position
and released from rest. Determine the subsequent motion.
Solution The initial value problem describing the motion of the Spring-mass system is
Ay" + 32y + 64y = 0, yO) =1, yO) = 0.
Dividing the equation by 4, we find the auxiliary equation A2 + 8A + 16 = 0 with
solutions A = —4, —4. Thus the general solution of the DE is
y= (Ce Collen:
and imposing the initial conditions leads to C; = 1 and C; = 4, resulting in
y=(1 + 4pe™.
The system is therefore critically damped. The mass will never pass through
its
equilibrium position y = 0, but will slowly approach it as time goes on.
i 2 —1/2
aot nan ~ | : (19)
which can also be expressed in terms of T [given by (11)]
b)
ae ce \ate oe
TFS (1 ets | ein ee
W, ( a r(1 ) f Cu
SEC. 5.2 / SMALL MOTIONS OF A SPRING-MASS SYSTEM 137
where wo = (k/m)'*. Here we see that when damping is small, i.e., when
c7/4km<<1, the quasiperiod T is approximately equal to T.
EXERCISES 5.2
A Find the frequency of oscillation of a spring-mass system if the mass is 4 kg and the
spring constant is 100 N/m. What is its natural period?
re Calculate the time necessary for a 0.03-kg mass hanging from a spring with spring
constant 0.5 N/m to undergo one full oscillation. What is the natural frequency of the
system?
A spring-mass system is stretched 6 in. by a 12-lb weight. If the weight attached to
the spring is pulled downward 4 in. below the equilibrium position and started upward
with a velocity of 2 ft/s, show that the subsequent motion is described by the function
y = 4cos 8r — fsin 81.
Show that the solution in problem 3 can be expressed in the form
y = Acos(8t — d),
and find the first two positive values of time for which y = 0.
A 10-lb weight stretches a steel spring 2 in.
(a) Determine the natural period of the spring-mass system.
(b) If the spring is stretched an additional 2 in. and then released, describe the
_ subsequent motion of the mass.
4 ew aus RANKKOAM ;
The period ‘of free oscillations of a mass on a string is 7/2 s. What is the numerical
value of the length of the string in feet? Hint: See Equation (13).
A clock has a pendulum | m long. The clock ticks once for each time the pendulum
makes a complete swing, returning to its original position. How many ticks will the
clock make in | min?
Hint: Use Equation (13).
A 24-lb weight stretches a spring 4 in. Determine the equation of motion if the weight
is released from a point 3 in. above the equilibrium position with a downward velocity
of 2 ft/s.
The period of free, undamped oscillations of a spring-mass system is observed to be
m/4 s. If the spring constant is given by 16 lb/ft, what is the numerical value of the
weight in pounds?
10. Find the solution of the DE
my" + ky =0
in the form A cos (wot — ) when the initial conditions are prescribed by
12. Prove that the maximum value of the speed of a mass undergoing simple harmonic
motion occurs when y = 0.
13. Determine the natural period of oscillation of the pendulum in problem 7 if the
pendulum is 2 m long.
14. Solve Equation (13) of the pendulum problem when the weight is 8 Ib and the rod is
1 ft long. Assume the weight is released from an angle of 3 rad with a positive velocity
of 3 rad/s.
15. At what time does the pendulum in problem 14 first pass through the angle 6 = 0?
What is its velocity at this time?
16. An 8-lb weight, attached to the end of a vertical spring, is pulled yo ft below its
equilibrium position and released at time t = 0 with a downward velocity of 3 ft/s.
Determine the spring constant k and the initial displacement yo if the amplitude of the
resulting motion is known to be V5 and the period is 77/2.
mae Show that underdamped free motion has the following characteristics:
(a) The characteristic angular frequency pw is independent of the initial conditions
but decreases as c increases.
(b) The natural logarithm of the ratio of two consecutive maximum amplitudes is the
constant 6 = mc/mp. The number 6 is called the logarithmic decrement of the
oscillation.
(c) Find 6in the case y = e ‘cost, and determine which values of t correspond to
maximum and minimum displacements.
18. Show that overdamped free motion has the following characteristics:
(a) The mass cannot pass through y = 0 more than once.
(b) If the initial conditions are such that the constants C; and C>2 in the general
solution have the same sign, the mass never passes through y = 0.
19. Under what conditions on yo and vo, where yo = y(O) and vo = y'(0), will critically
damped free motion have a maximum or a minimum for t > 0?
20. A spring is stretched 6 in. by a 3-lb weight, which is started from its equilibrium
position with an upward velocity of 12 ft/s. If a retarding force equal in magnitude
to 0.03v exists, find the resulting motion.
A certain straight line motion is described by the initial value problem \
If an external force F(t) is also present, the initial value problem describing the
possible forced motions of a spring-mass system reads
In order to investigate undamped motions due entirely to the input function F(#),
it is convenient to assume the system is initially at rest. Thus we prescribe the
conditions
where w) = Vk/m is the natural (angular) frequency of the system. Applying the
initial conditions (23), we deduce
In this case the mass oscillates at its natural frequency between the points y = 0 and
y = 2P/k with a period of 277/a s.
If a sinusoidal force F(t) = P cos wt is applied, we assume a particular solution
exists of the form (see Section 4.6)
*A viscous damping term arises if the spring-mass system is suspended in a fluid like oil or water, or
if air resistance cannot reasonably be neglected.
140 CHAP. 5 / APPLICATIONS INVOLVING INITIAL VALUE PROBLEMS
yp = Acoswt + Bsinat,
where the constants A and B are to be determined. The substitution of y, into the
DE leads to
(k — mw’)(Acos wt + Bsin wt) = Pcos at.
Hence, equating like coefficients of coswt and sinwt, we obtain A =
P/(k — mw) = P/m(ws — w*) and B = 0, so our general solution is
Since |@ — @o| is small, the period of the sine wave sin[(w — wp) t/2] is large
compared with the period of sin[(@ + @p) t/2]. The motion described by (26) can
then be visualized as a rapid oscillation with angular frequency (@ + wp)/2, but
with a slowly varying sinusoidal amplitude (see Figure 5.8). Motion of this type,
possessing a periodic variation of amplitude, exhibits what is called a beat. The
al 9
27
i | or oy
Figure 5.8 Phenomenon of beats.
SEC. 5.3 / FORCED MOTIONS 141
It is clear that the amplitude of motion will become unbounded in (27) as t>~,
and thus we have the phenomenon of resonance (see Figure 5.9). Of course, we
recognize that in a physical problem the amplitude cannot become unbounded. A
certain amount of damping, however small, is always present, and this has the
effect of limiting the amplitude. Also, if the amplitude should become large
enough, the system is likely to fail. This situation has actually caused certain
bridges to collapse, such as the Tacoma Narrows bridge at Puget Sound in the state
of Washington. On November 7, 1940, only four months after its grand opening,
a huge portion of the bridge collapsed into the water below. From the very begin-
ning the bridge had experienced large undulations, which were later attributed to
the wind blowing across the superstructure. Therefore, in designing such struc-
tures, it is very important to make the natural frequency of the structure different
(if possible) from the frequency of any probable forcing function.
cals
Remark. It should be noted that the reason soldiers do not march in step across
bridges is to avoid any possibility of resonance occurring between the natural
frequency of the bridge and the frequency of the uniformly stomping feet.
When damping effects are important, we can think of the solution of (21) as
composed of two parts, i.e., y = yp + yy. Since it contains the multiplicative
os
factor e~/2™, the solution function y, contributes only initial effects to the motion
and it is called the transient solution. The function yp dominates the response of the
system after initial effects diminish and is therefore referred to as the steady-state
solution of the system.*
Let us assume the forcing function is F(t) = Pcos wt. Then it seems natural to
assume that the steady-state solution has the general form
yp = Acoswt + Bsinat, (28)
where A and B must be determined. Observe that we are assuming that the fre-
quency of the steady-state motion is the same as the frequency of the forcing term
producing the motion. When (28) is substituted into the DE, we obtain
[((k — mw*)A + wcB|cos wt + [(k — mw?)B — wcA]|sin wt = P cos at.
Equating like coefficients of cos wt and sin wt results in the simultaneous solution
of A and B
P(k — mw’) Pwc
Ae Ge
(k — mw’) + wc”
re Bg ee
(k — mw’) + wc?
The particular solution, or steady-state solution, is then
P(k — mw’) Wwe :
ee = Goma ame ¢ Sa
kona.
ot: (29)
Since the amplitude R of the oscillation is of some importance, let us rewrite (29)
in the equivalent form
yp = Rcos(at — ), (30)
where (@) = Vk/m)
P
Kes [m%(we — w2)? + wc?” (31)
tangd =——~
k— mw
a
m(w — w*)
(32)
Unlike the case of undamped motion, here we see that the maximum value of R
does not occur when w = wo. The maximum amplitude of the motion can be found
by setting dR/dw = 0, which occurs when (see problem 19 in this section)
Spee (33)
However, for sufficiently large damping such that c? > 2m?w4, there is no value
of w satisfying (33), and hence no maximum amplitude. For damping coefficients
*By steady state, we mean only that portion of yp that does not go to 0 as t>,
SEC. 5.3 / FORCED MOTIONS 143
The amplitude R given by (31) is plotted in Figure 5.10 as a function of the input
frequency w. It thus becomes clear that large amplitudes due to resonance can be
avoided by a sufficient amount of damping.
c= Mwy
p Ee
mw
C= MW
c= 2m = c< V2 may
(critical damping) —] TSS - Seinen a
-=<—
Upon differentiation and substitution into the nonhomogeneous DE, we find that
the coefficients A and B must satisfy the equations
—8A + 168 =0, 16A + 8B = 10.
Calculating, we have A = 3 and B = j, so that
1 Lay
yp = gos ott qoin2t.
Gn 1 AG 2 Se2 eye’
Peer sine
&
y= ge a 5I 008 2t a 4Lisesin2t.
(ee
transient solution steady-state solution
After a short period of time (¢ > 3), the dominant part of the solution is the
steady-state term, which represents simple harmonic motion with maximum am-
plitude
Oia:
Ne Tye
k= = Wert = feree Os
and period 7.
EXERCISES 5.3
A 2-lb weight stretches a spring 6 in. An impressed force 16 sin 87 is acting upon the
spring, and the weight is pulled down 3 in. below the equilibrium position and
released. Determine the equation of motion.
If an impressed force }cos 41 is imposed upon the system in problem 3 of Exercises
5.2, determine the subsequent motion.
*3. A spring with spring constant k = 0.75 lb/ft has a weight of 6 Ib attached, which is
at rest in the equilibrium position. A 1}-lb force is applied to the weight in the
downward direction for 4 s and then removed. Discuss the subsequent motion.
A spring stretches 6 in. when a 4-lb weight is attached. If the weight is started from
the equilibrium position with an upward velocity of 4 ft/s and has an impressed force
of 5cos 8t acting on the weight, determine the position of the weight for all time. What
is the position when ft = 2 s?
A 2-kg mass is attached to a spring with k = 32 N/m. A force of 0.1 sin 4t is applied
to the mass, which is at rest. Neglecting damping, calculate the time required for
failure to occur if the spring breaks when the amplitude of oscillation exceeds 0.5 m.
Show that the solution of
y’ + 25y = 10cos7t, y(O)=0, y'(0) =0,
is given by y = 2sin¢sin6t. How many seconds are there between beats?
A 20-N weight is suspended by a frictionless spring for which k = 98 N/m. An
external force of 2cos7t acts on the weight. Find the frequency of the beat, and
determine the maximum amplitude of the motion that starts from rest.
How many seconds are there between the beats in problem 7?
Verify Equation (27) by solving directly the initial value problem
16. A mass of 3 slug is attached to a spring with spring constant k = 6 Ib/ft. A damping
force numerically equal to twice the instantaneous velocity acts on the system.
(a) Find the steady-state response of the system due to an external driving force
F(t) = 40sin2t.
(b) Will Riax occur?
(c) What is the amplitude in this case?
17. Find the steady-state response of the system in problem 16 if the driving force is
constant; i.e., F(t) = P.
“18. The ratio of successive maximum amplitudes of a particular underdamped spring-
mass system is found to be 1.25 when the system undergoes free motion. If k = 100
N/m, m = 4 kg, and a driving force of F(t) = 10cos 4t is imposed on the system,
determine the amplitude of the steady-state motion.
19. Show that dR/dw = 0, where R is defined by (31), occurs when the frequency w
satisfies w? = we — c7/2m?.
20. Derive Equation (34).
R E(t)
1. The sum of the currents into (or away from) any point is zero.
SEC. 5.4 / SIMPLE ELECTRICAL CIRCUITS 147
. The first Kirchhoff law indicates that the current is the same throughout the
circuit. To apply the second law, we must know the voltage drop across each of the
idealized elements in the RLC circuit. From experimental observations, we have
voltage drop across a resistor = Ri, f
di
voltage drop across an inductor = Le
t
where q denotes the electric charge on the capacitor and is related to the current i
by i = dq/dt. The impressed electromotive force E(t) contributes to a voltage gain.
Applying the second Kirchhoff law to the circuit shown leads to the differential
equation
z
ts +Ri t+ Cg =E(). (35)
In terms of the charge q, we find
] La’ +Rq' + Cg =EQ@, t>0/ (36)
We recognize that (36) is the same equation in form as that which governs the
damped motions of a spring-mass system. In fact, we notice the following analogies
between mechanical and electrical systems.
Such analogies between mechanical and electrical systems prove very useful in
practice. For example, in studying a certain mechanical system that is either too
complicated or too expensive to build, the electrical counterpart is often constructed
instead for the purpose of analysis. Interestingly, the phenomenon of resonance
also occurs in electrical systems, but it does not have the undesired side effects of
mechanical resonance. Quite the contrary, it is primarily because of electrical
resonance that we can tune a radio to the frequency of the transmitting radio station
in order to obtain reception.
The units most commonly used are listed in Table 5.2
148 CHAP. 5 / APPLICATIONS INVOLVING INITIAL VALUE PROBLEMS
ee ee eee ee ee
EXAMPLE 7 The circuit shown in Figure 5.11 contains the components L = 1H, R = 10000,
andC = 4 x 10°°F. Attimet = 0, both current and charge are zero, and a battery
supplying a constant voltage of 24 V is instantaneously switched on. Find the
charge q(t) on the capacitor and the current i(t) for any later time.
EXERCISES 5.4
If the resistance R is not included in the RLC circuit of Figure 5.11, we have what is
called an LC circuit. Show that the current i(f) of the LC circuit satisfies the relation
5 E(O) (0)
es a: LC # 0,
where q(Q) is the charge in the capacitor at time t = 0.
Using problem 1, find the current in the LC circuit when i(0) = 0, g(0) = 0, and
(a) L =1H,C =0.25 F, E(t) = 30sint V.
(b) L=10H,C =0.1 F, E() = 10t V.
A steady-state current in the RLC circuit results after a sufficient length of time
(t—). Find the steady-state current where
(a) R=40,L =1H,C =2~x 10%F,
E(@) = 220 V.
(b) R=100,L =2H,C =0.5 F, E(t) = 10.9cos2r V.
Show that the current i(f) in the RLC circuit satisfies
IN eeE(O) _ ed
R q(0)
ee
i'(O) L Oo) LC” LC
# 0,
i(t) = Eo aor = SE
vA i
where X = Lw — 1/Cwand Z = VX? + R’. The quantityX is called the reac-
tance of the circuit, and Z is the impedance of the circuit.
(b) Show that when
1
o-=— =,
VEC
the amplitude of the steady-state current is a maximum. Electrical resonance is
said to occur for this value of w.
*GEORGE GREEN (1793-1841) gained recognition for his important works concerning the reflection and
refraction of sound and light waves. He also extended the work of Poisson in the theory of electricity
and magnetism.
SEC. 5.5 / THE METHOD OF GREEN'S FUNCTION 151
yilto) y3(to)
Remark. The explicit representation for the constants C, and C; as given by (42)
is mostly of theoretical importance. In practice, it is usually just as convenient to
solve for these constants directly, as illustrated in Example 8 below.
Notice that when ky = k,; = 0, we get C; = C, = 0, so that necessarily yy = 0.
The physical implication of this result is that a system which is initially at rest and
not subject to any external disturbance must remain at rest.
*Theorem 5.2 is a repeat of Corollary 5.1 for the case of second-order DEs.
152 CHAP. 5 / APPLICATIONS INVOLVING INITIAL VALUE PROBLEMS
In order to solve the problem described by (40), we will start by assuming yp has
the form
yp = u(t)y\(t) + v(y2(t) (43)
for some functions u(t) and v(t). Using the method of variation of parameters as
discussed in Section 4.7, we find that
yf (0) yilOf(0)
u(t) = —
TENG ee lao
+. 7 ; = pe LEN TRA OFM aS
where W(y;, Y2) = yiy3 — yiy2 is the Wronskian function. If we choose u(t) and
v(t) as any indefinite integrals as indicated above, the resulting solution function
yp is unlikely to satisfy the prescribed homogeneous initial conditions. A proper
choice turns out to be
t ya T)f(7) t y(t
)f(7)
Meet
=-
Cah a
ORR eo kicckse
Te
olaile
+
Aa
which we choose to write as
yi(t) y2(T)
yi(t)ya(t) — yidy2(7) y(t) y(t)
Oa 46
cee W(y1, y2)(7) y(t) y(7) “
yi(t) ya(t)
To show that (45) satisfies the homogeneous initial conditions in (40), first
observe that when t = fo, we have
ang t OF
al
aes F
Gonliai
b
igs6 =
[ es5p m2 t) ae oe AC ae
we find
19 02
yp (to) = Ba (to, TIF(T) dt + gi(to, tof (to),
t9
The function g,(t, 7) is called the one-sided Green’s function for the initial value
problem described by (37). Its construction depends only upon knowledge of the
homogeneous solutions y(t) and y2(f); i.e., it is independent of to and the prescribed
initial conditions and is completely determined by the operator M = D? +
a,(t)D ar a(t).
Remark. It is important to observe that the particular solution (45) was derived
under the assumption that the DE was in normal form. Strict adherence to this form
is necessary for proper identification of the forcing function f(t).
Remark. Although (47) represents a general solution formula for the given initial
value problem, it does not always represent the simplest approach to finding the
solution. The Green’s function method is important for developing and under-
standing some of the general theory, and it can be useful in those situations where
the same DE must be solved a number of times with various input functions. It is
the responsibility of the practitioner to determine those occasions for which such
general formulas are useful.
Solution Linearly independent solutions of M[y] = 0 are simply y\(t) = cos ¢ and y2(t) =
sin t. The Wronskian is
COS.t ssinft)P
W(y1, y2)(t) = |
—sint cost
?
EXAMPLE10_ Using the one-sided Green’s function, solve the initial value problem
154 CHAP. 5 / APPLICATIONS INVOLVING INITIAL VALUE PROBLEMS
yp i)sin (4 — 7) sin 7 dt
0
t t A
li sin | cos T sin T dt — cos ¢ | sin? t dt
0 0
= eee bot
2 2 :
yy = 0? —-t.
The Wronskian of y, = ft and y) = f° is
raat
W(t, ?) = = 273,
338
and therefore the one-sided Green’s function is given by
3
gilt, T) =
ihe a Mee earn is
Noms Pot 27° ;
Dividing the DE by the leading coefficient r* puts the equation in normal form, from
which we identify f(t) = 2r’e'. Therefore,
yp = |ets \2s%e)das
1
*Note that the lower limit of integration is | since the initial conditions are prescribed there.
SEC. 5.5 / THE METHOD OF GREEN'S FUNCTION 155
ty t
| et dr ~ 1| te’ dt
1 1
(t — Pe + 2t(t — Ne’.
Finally, combining solutions, we have
It should be observed that one of the distinct features of the Green’s function
technique is that the nonhomogeneous initial conditions are imposed only upon the
solution of the associated homogeneous DE. This is in sharp contrast with the
methods employed in Chapter 4, wherein it was first necessary to find a general
solution of the nonhomogeneous DE before applying the prescribed initial condi-
tions. The reason for this situation, of course, is that we are selecting a particular
solution yp that always satisfies homogeneous initial conditions.
A listing of one-sided Green’s functions for some of the more common differential
operators is provided in Table 5.3 for easy reference.
Le 1B Peg
(t — 7)"|
Ga = i
1 5
2
8. Ds 2
tp p sin b(t — 7)
4. D? — b? +sinh b(t — 7)
I alm). b(t—7)
5: (D
=
—a)(D
=
—b),a#b eh se e
6. (D — a) (<7) cone
?D?+tD —b?
156 CHAP. 5 / APPLICATIONS INVOLVING INITIAL VALUE PROBLEMS
EXERCISES 5.5
In problems 1-8, find the solution satisfying the prescribed initial conditions.
In problems 19-24, use the one-sided Green’s function to solve the given initial
value problem.
25. Show that the one-sided Green’s function associated with the undamped spring-mass
system described by
my" + ky = F(t), y(O) =yo, y'(O) = vo,
feo kK
gi(t, T) =— sin[wol(t — 7)], wo= ve
Wo m
26. Using the one-sided Green’s function given in problem 25, find the response of the
spring-mass system described there when the system is initially at rest and subject to
the forcing function
(a) F(t) =P (constant).
SEC. 5.5 / THE METHOD OF GREEN'S FUNCTION
157
(b) F(t)=Pcos
wt, w # a.
(c) F(t)
=P cos wot.
wai Show that the one-sided Green’s function associated with the damped spring-mass
system described by
KM = (dk
=62)am :
ae (c? — 4mk)'?
2m ;
*28. Using the one-sided Green’s function given in problem 27(a), (b), and (c), find the
steady-state response of the system in all three cases when the driving force is
31. t’y"—6y
”
=logt, y)=—2,l y()=
'
areI
32. t?y’ + ty’ + 4y =sin(logt), y(1)=1, y'(1)=0
33. Show that for all 7 = fo,
(a) gi(t, 7) = 0.
(b) gi(t, 7), along with its first and second derivatives with respect to ft, is con-
tinuous.
34. Show that for a fixed value of 7, the function @(t) = g,(t, 7) is a solution of the initial
value problem
y’ + a(t)y’ + a(t)y = 0, (7) =0, O'(7) = 1.
Hint: Use the result of problem 33.
where y:, y2, and y3 are linearly independent solutions of the associated
homogeneous DE.
(b) Establish the solution formula
In problems 36-41, find the one-sided Green’s function for the given operator
using the result of problem 35.
Solve the initial value problems 44 and 45 using the one-sided Green’ sfunction (see
problem 35).
*44, y”+y m
=te’, yd)=0, y')=0, y")=1
*45. yr ye tay oe dye 1 meyO) c= Ons ye Okan Umm (Olea
46. Prove that M = D? + a,(t)D + ao(t) is a linear operator; 1.e., that
M[C\y; + C2y2] = CiM[yi] + CoM[y2]
for any constants C,; and C).
Now let us idealize the function d,(t) by requiring it to act over shorter and shorter
intervals of time by allowing e—0. Although the interval about ¢ = a is shrinking
to zero, we still want J = 1; thus it follows that
We can use the results of this limit process to define an idealized unit impulse
function, 6(t — a), which has the property of imparting a unit impulse to the
system at time ¢ = a but being zero for all other values of t. The defining properties
of this function are therefore
Of — a) =O tas
Solution Because of the homogeneous initial conditions, we only need consider the particu-
lar solution; i.e., y = yp. The one-sided Green’s function for the operator
M = D* + I has previously been shown to be (see Example 9)
*Named after PAUL A. M. DirAc (1902-), who was awarded the Nobel Prize (with E. SCHRODINGER)
in 1933 for his work in quantum mechanics.
SEC. 5.6 / IMPULSE FUNCTIONS
161
Figure 5.13
a ee
The unit impulse function is also useful for providing us with a physical inter-
pretation of the one-sided Green’s function. To see this, let us consider the initial
value problem
where g(t, 7) is the one-sided Green’s function associated with the operator M.
Thus the function g(t, a) must represent the response of the system described by
(53) for t > a, which was formerly at rest and then subjected to a unit disturbance
(impulse) at time ¢ = a.
Based on the above interpretation of g)(t, 7), let us introduce the more general
function
0, i St Se
co Una he
gi(t, as 7 Sy Se,
or equivalently,
*The influence function in some textbooks is defined as a constant multiple of g(t, tT) rather than g(t, +)
itself.
162 CHAP. 5 / APPLICATIONS INVOLVING INITIAL VALUE PROBLEMS
For a fixed value of r, we see that g(t, 7) must necessarily satisfy the differential
equation M[g] = 5(t — 7), where M is the differential operator associated with the
construction of g\(t, 7). The Green’s function must also satisfy the homogeneous
initial conditions g(t, T) = (dg/dt)(t, T) = O for any % less than or equal to the
fixed value of 7. It is a continuous function for all values of t since g\(7, T) = 0
and g\(t, 7) is itself a continuous function of t by definition. However, there is a
jump discontinuity in the first derivative of g(t, 7) att = Tof unit magnitude. That
is, for t < 7, the derivative is clearly zero, while
0g _ yilt)y2(7) = yiryal7) _ ,
Ot |r=r+ W(y1,¥2)(T)
The jump discontinuity in the derivative turns out to be an essential feature of the
Green’s function.
In summary, we have the following definition.
Definition 5.1 The Green’s function g(t, T) associated with the nonhomogeneous initial value
problem
M[yl=f(t), t >t, y(t) =k, y'(o) =k,
is a function satisfying the following conditions:
EXERCISES 5.6
In problems 1-5, find the solution of the given initial value problem.
[ fe) dx = Ob ~ a)
b
a [dingo dt eet
lim = lim 52 is ery
f(t) dt ke= lim
ye 5- l+
east f(€) >2¢,
whereas
€1— 6 <daiet €.
Show formally that dh(t)/dt = 6(t), where h(t) is the Heaviside unit function defined
by
Ona)
nia)= | be,
Hint: Use integration by parts to show that
carat
a CFO) dt = FO).
10. Show that 6(¢ — x) = 6( — f).
ah Show formally that f(t) 6(¢ — a) = f(a) 6(t — a), and use this result to deduce that
t d(t) = 0.
Hint: Show that thy e(t)if(t) 6(t — a)] dt = iL e(t)if(a) d(t — a)] dt.
Wp (t) =
ppg ea 1
lim | (dt = 1.
164 CHAP. 5 / APPLICATIONS INVOLVING INITIAL VALUE PROBLEMS
V,
n
2
t
Leo
n =e
Problem 14
(b) More generally, if f(t) is any function continuous at t = 0 and bounded, show
that
(a) Y(t)= ee
LD PF Waal eu
aw(1 + n°t?)’
We now turn our attention to equations more general than the constant-coefficient
DEs arising in most spring-mass systems or electric circuits. In particular, we want
to discuss the DE
A2(t)y” + Ai(t)y’ + A,(t)y = 0, (57)
in which certain parameters of the system vary over time. Although in general we
cannot obtain explicit solutions of (57), we can study the qualitative behavior of
these solutions by directly analyzing the equation itself. For example, information
concerning the existence and relative positions of the zero-crossings can be ob-
tained without formal solutions. In the theorems that follow, we will treat the
coefficients of (57) as continuous functions on the interval of interest and assume
A,(t) # 0 on this interval.
Theorem 5.3 (Sturm separation theorem) If y, and y, are linearly independent solutions of
the second-order equation
on some interval I, then y, has precisely one zero between any two consecutive
zeros of y2 on the interval /.
Figure 5.14
Put another way, Theorem 5.3 states that the zeros of y, and y, occur alternately,
and that on any finite interval the number of zeros of y, and y, can differ by at most
one. For example, the solutions cost and sint of y” + y = 0 have zeros that
alternate on the entire real axis. A somewhat less obvious consequence is that any
two functions of the form
y, = C,cost + Cysint, y, = C3cost + C7 sin t
have alternating zeros whenever C\C, # C>C;, as all such pairs of functions are
also linearly independent.
The Sturm separation theorem does not imply the existence of any zeros of the
solutions y, and y). Furthermore, it does not allow for a comparison of the number
of zeros (rate of oscillation) of solutions of two distinct DEs. To pursue these
matters it becomes convenient to put the differential equation in what we call the
standard form.
166 CHAP. 5 / APPLICATIONS INVOLVING INITIAL VALUE PROBLEMS
ONC M te Perel 1
sen ea Stent loz 4y2
=
and Bessel’s equation reduces to
t+ (14259) y= 08
We can now show that if Q(t) is negative on a given interval, the solutions of
(61) do not oscillate on that interval. To better understand this situation, consider
the solutions of y” — y = 0.
*Although we don’t need it, we see that v(t) = ¢~"/? in this case.
SEC. 5.7 / ELEMENTARY OSCILLATION THEORY
167
Theorem 5.4 If Q(t) < 0 on the interval J and uw is a nontrivial solution of u” + Q(t)u = O, then
u has at most one zero on the interval /.
Theorem 5.5 (Sturm comparison theorem) Let u, and u2 represent nontrivial solutions of
Proof. Let t = aandt = b be consecutive zeros of uz, and assume that u, does
not vanish on the interval a < t < b. We will further assume that both u, and uw,
are positive on the interval a < t < b. Since the zeros of u, are the same as —u,
this assumption is justified. Following the argument presented in the proof of
Theorem 5.3, we have
d d ; '
qu u2)(t)] = Gy i — Uj Up)
= WU — UjUy
—UjQouz + U,Q {Uy
u,U(Q; — Qo),
and since Q,(t) > Q(t), we conclude that the Wronskian is an increasing function
on a <t <b. Also, since we have assumed that u, is positive on this interval, it
follows that u5(a) = O and w3(b) S 0, and hence
The solution (sin 22) of the second equation oscillates more rapidly than the solution
(sin?) of the first equation. That is, the zeros of sin 2f are located at t = 0, 7/2)
a, 37/2, . . . , whereas the zeros of sint are located att = 0, 7, 27, . . . (see
Figure 5.15).
>
\ y=sint
\
N\
—IF
EXAMPLE 14 Show that every solution of y” + t*y = 0 has infinitely many zeros on the interval
seca
Solution The equation y” + y = O has one solution sin with infinitely many zeros at tr= 0
and t =n, n = 1, 2, 3,.... Since t? > 1, it follows from the Sturm com-
parison theorem that one nontrivial solution of y" + t?y = 0 has at least one zero
between nar and (n + 1) for n = 1, 2,3, ... , and hence has infinitely many
zeros. By comparison with cost, it can be established that a second linearly
independent solution of y" + ¢*y = 0 has infinitely many zeros between the zeros
of cos ft. Since all other solutions of this DE are linear combinations of these two
linearly independent solutions, we have our intended result.
The zeros of Bessel’s equation are of both practical and theoretical importance,
and are found to satisfy the conditions of the following theorem by comparison with
the zeros of solutions of y” + y = 0.
Theorem 5.6 (Zeros of Bessel’s equation) Every nontrivial solution of Bessel’s equation, which
has the standard form
eee ( eee = 0,
4t-
has infinitely many zeros on the positive t-axis. Moreover, the distance between
suCCesSive Zeros is
SEC. 5.7 / ELEMENTARY OSCILLATION THEORY 169
EXERCISES 5.7
13. Putthe DE my” + cy’ + ky = 0 instandard form, and show that it does not possess
oscillatory solutions unless c* — 4mk <0.
14. Prove Theorem 5.6.
REFERENCES
Birkhoff, G., and Rota, G.-C. 1978. Ordinary differential equations. 3rd ed. New
York: Wiley.
Boyce, W. E., and DiPrima, R. C. 1977. Elementary differential equations and bound-
ary value problems. 3rd ed. New York: Wiley.
Bradbury, T. C. 1968. Theoretical mechanics. New York: Wiley.
Finney, R. L., and Ostberg, D. R. 1976. Elementary differential equations with linear
algebra, Reading, Mass.: Addison-Wesley.
Potter, M. C. 1978. Mathematical methods in the physical sciences. Englewood Cliffs,
N.J.: Prentice-Hall.
Ross, S. L. 1980. Introduction to ordinary differential equations. 3rd ed. New York:
Wiley.
rm isan e ficient method for s near, stant-coefficient -
ns with scribed auxiliary conditions, usually in the form of initial conditions.
It offers the advantage of solving the problem directly without first producing the
general solution or even having to solve separately for the homogeneous and
particular solutions.
The transform method is formally equivalent to the operational calculus devised
by Oliver Heaviside (1850-1925) for the solution of transient problems in physics
and electrical engineering. It is an especially useful tool for solving problems in
circuit analysis, where the nonhomogeneous terms are frequently of a discontinuous
or impulse nature. More conventional methods tend to be clumsy when piecing
together solutions valid in different intervals, whereas the Laplace transform has the
effect of “smoothing” the problem in the transform domain and making it more
tractable.
In Section 6.2 we calculate the transforms of some elementary functions directly
from the integral definition. We follow this in Section 6.3 with some of the operational
properties of the transform, which permit methods of calculating certain transforms
other than straight evaluations of the defining integral.
We discuss methods of computing inverse Laplace transforms in Section 6.4,
which rely mostly on operational properties and sometimes partial fraction
expansions.
in Section
Solutions of initial value problems by the transform method are studied
g some of the
6.5 and the vibrating spring-mass system is revisited, illustratin
benefits of a transform analysis.
impulse func-
The use of the Laplace transform in dealing with discontinuous and
d in Section 6.6, and in Section 6.7 the convoluti on theorem and its
tions is discusse
last section contains a
relation to the one-sided Green’s function are examined. The
short table of Laplace transforms.
171
6.1 INTRODUCTION
K(s, 0)
a ia
(2)
*Named in honor of the French mathematician PIERRE SIMON DE LAPLACE (1749-1827). Laplace made
use of this particular integral transform in his work in probability theory, although it is believed that
the integral was really discovered by Euler.
172
SEC. 6.2 / THE LAPLACE TRANSFORM OF SOME ELEMENTARY FUNCTIONS 173
Lif} = I e “sint dt
0
ey
=
= st
CmaCOSt — s| en costdt
0 0
ae CT COSdia Ss 1 0:
0
PS KSCAN ADE
and solving for S{f(t)}, we find
Not all functions of t have a Laplace canst even if they are continuous,
since the defining infegral is1 imprOPee The basic requirement for the existence of
i
transform is that f be of «
the 1
Definition 6.1
a:see Figure 6.1). For instance, ihe functions 1, e”, and sin tareall Bf eteoucn
order, whereas the function f(t)= e” is not since its graph grows faster than any
linear power of e for t > fo (see Figure 6.2).
} S Me“
>t
Definition 6.2 | A function f(t) is said to be piecewise continuous in a given interval provided
(a) f(t) is defined and continuous at all but a finite number of points in the
interval, and
(b) the left-hand and right-hand limits exist at each point in the interval.
and even when it is, the functional value assigned at these points really doesn’t
matter. Also, the interval of interest may be open or closed, or open at one end and
closed at the other (see Figure 6.3).
Theorem 6.1 | ecewise continuous and of exponential order, then it has a Laplace ~
Proof: Consider
+| era
[e“feoa = [Pepa
‘CO
to
The first integral on the right exists, since f is assumed piecewise continuous. Since
fis also of exponential order, the second integral on the right satisfies the inequality
For s > c, this last expression vanishes in the limit as fy so that we say the
integral is absolutely convergent. [a]
Most functions met in practice satisfy the conditions of Theorem 6.1. However,
these conditions are sufficient rather than necessary to ensure that a function has a
Laplace transform. For example, both t-'! and t'? have infinite discontinuities at
t = 0, and while the integral
Aft} = [etetas
0
Hence t~'’? has a Laplace transform, but t~' does not. Finally, we remark that if
the transform of a-function exists, it is unique, since the definite integral of a
function is uniquely determined.
EXAMPLE 4 Find the Laplace transform of the piecewise continuous function f(t) where
tes t AO eee
FOES Sh nip ele
bitten “4 : id
20
1 1 =
Se 6 te Oa Ore ee
5 RY S RY
which simplifies to
Hitt (Dip
{elabirencore aa5e Te
Figen
Remark. Observe that the original function f(t) in Example 4 had a discontinuity
at t = 2, whereas the transformed function F(s) is continuous for all s > 0. This
phenomenon is known as the “smoothing effect” of the transform.
EXERCISES 6.2
In problems 1-12, evaluate the Laplace transform of each given function directly
from the defining integral.
1) f@=t 20 FG)
On ue ue ee 4. f(t) = sinkt
SEC. 6.2 / THE LAPLACE TRANSFORM OF SOME ELEMENTARY FUNCTIONS 177
In problems 13-16, evaluate the Laplace transform of each function, recalling the
identities coshx = (e* + e *)/2 and sinhx = (e* — e*)/2.
a
{aw
\43. f() = cosh?kt 14. f(t) = sinh*kt
15. f() = e“coshkt 16. f(t) = e“sinhkt
In problems 17-20, evaluate the Laplace transform of each function using integra-
tion by parts.
*27. Using the results of problem 26 and the special value (3) = V7, verify the follow-
ing identities.
178 CHAP. 6 / THE LAPLACE TRANSFORM
/2
where
Operational properties often offer an attractive alternative for evaluating the trans-
form of a function to tedious and often cumbersome evaluations of the integral that
defines the transform. For example, it becomes fairly easy to evaluate the trans-
forms L{e~* cos 2t}, L{t? sin St}, and L{r°e*} with these operational properties
once we know the transforms of L{cos 27}, f{sin 5t}, and L{t°}, respectively. First
among such properties is the linearity property, which is a simple consequence of
integrals.
Theorem 6.2 (Linearity) If L{f(p} and L{g(H} are the Laplace transforms of f(t) and g(#),
respectively, then
= CAFO} + CAz(o}. O
SEC. 6.3 / OPERATIONAL PROPERTIES 179
indSoe verte
St
Sin
es eal
where we have used the results of Examples 1, 2, and 3.
Solution While both of these transforms can readily be found through routine integration
methods, another approach using Theorem 6.2 is available to us. Setting a = ik in
L{e"} = 1/(s — a), we get
1 S
Tier deed teege
From Euler’s formula, e’® = cos @ + isin 6, and Theorem 6.2, it follows that
Solution From Example 6, we know that S{cos 37} = s(s? + 9)~', and hence, through the
shifting property, it follows that
SAEZ
Die =2i
ta cos3ty =
G +2
ee
+9"
are Se
The real merit of the Laplace transform is revealed by its effect on derivatives.
Suppose that f(1) is continuous with a piecewise continuous derivative f’(1). Then
by definition
where we have performed an integration by parts. If the function f(‘) has a Laplace
transform F(s), then lime “f(t) = 0, so that
EXAMPLE 9 Use the result L{sin kt} = k/(s? + k”) and Equation (6) to evaluate L{cos kt}.
Solution Let f(t) = sinkt and then f'(t) = kcoskt. From (6) it follows that
Similarly, if f(¢) and f'(t) are continuous and f(t) is piecewise continuous, and
all three have transforms, it follows that
Theorem 6.4 | (Differentiation) If f(t), f'(),..., f" ?@ are continuous for t = 0 and of
exponential order, and if f(t) is piecewise continuous and of exponential order,
then
Dike S
The function f(t) = ¢” and all its derivatives are continuous and of exponential
order. Also we observe that
RY
and therefore
and formally differentiate both sides with respect to s. This action leads to
Solution From the known relation £{sint} = (s* + 1)~', simple differentiation with respect
ee tee a
to s yields
d 2s
{t sin t} ras [(s Dal G+ De
(0) =(DFM).
Theorem 6.6 If F(s) = L{f(}, then forn = 1, 2,3,...
The integration of both sides of (8) produces yet another interesting result.
af
heorem 6.7 If F(s) = L{f(o}, then
ee = [ Fwdu,
F(s) = i“ef(0)dt
184 CHAP. 6 / THE LAPLACE TRANSFORM
with respect to s from s to , and interchanging the order of integration on the right,
we find
[F(u)du = [ =| f()dt
ioe) 2]! e -—st |]
s Fh tage! aah
= ea Ka dt,
| te ~costdt.
0
Letting s = 2, we find
% 3
I e te ~costdt
cos = —.
5
EXERCISES 6.3
In problems 1-12, evaluate the Laplace transform of the given function using
known results and any of the operational properties.
“igs fin
=2 —< 10. f= [waa + en")du
0
‘ep eae and f(t), where f(t) = t + 1,0 <t S2, and f(t) = 3, t > 2.
(a) Find PL}.
(b) Find S{f'(t} in two ways.
14. If L{f()} = F(s), show that L{f(at} = (1/a)F(s/a), a > 0.
15. Using problem 14 and L{cos t} = s(s* + 1)", find L{cos 44}.
| 46. If n denotes a positive integer, obtain L{t"e“} from L{e“} = 1/(s — a) using The-
orem 6.6.
17. Using Theorems 6.5 and 6.7, determine the Laplace transform of the sine integral
sity = [du
a
18, Using transforms, evaluate the following integrals.
og Se St? = ea CE!
(a) | ee ea (b) | Saeed
10) 0
(b) Using the relation 5’ e “= (1 —e ~*~ verify that the Laplace transform of
n=0
(poy = LOE
2
*21. Use the result of problem 20 to establish the following Laplace transform relations:
= i, OS
rp Se
FO = ors ee.
L c 2c 3c 4c 5c
=
Problem 21(a)
186 CHAP. 6 / THE LAPLACE TRANSFORM
OSS
fO = Ne Sie SiS We
Problem 21(b)
Pa
|
2a
At |
Problem 21(c)
e'd
n
*22. The Laguerre polynomials are defined by L,(t) = nian! é),. m=O; 172° ee
Show that
Lait} _1fs—1\n
= 4(— 4)",
Hint: First find L{e 'L, (b}.
*23. The error function (erf) and complementary error function (erfc) are defined, re-
spectively, by
Die e gre Es
erf ) =| ear, erfe(1) = =e) ede
Veal (=a) eta
(a) Show that erf() = 1.
Hint: Use the gamma function relation TG) _ Vox, given in problem 29 of Exercises
(9),7%,
(b) Show that erfc(t) = 1 — erf(a).
*24. Using problem 23, establish the following Laplace transform relations.
(a) L{e~V") = Ware” erfe(s).
Hint: Write jt? + st = (4t + s)? — s? and make the change of variable y = 4¢ + s.
SEC. 6.4 / INVERSE LAPLACE TRANSFORMS 187
(c) ‘Heriot )} = te S
co > eras qr?
Hint: | ee aS Pia a
0 2a
In applications the use of the Laplace transform is effective only if, given some
function F(s), we can find the inverse transform f(t). In symbols, we write
fO = L NFO}. (10)
We might well wonder if an explicit representation for the inverse Laplace trans-
form exists analogous to the integral representation of the transform itself. Such a
representation does exist, but it requires integrations performed in the complex
plane that involve techniques of complex variables. Therefore, we will rely on other
methods for constructing the inverse transform.
In the process of finding the inverse transform f(t) of a function F(s), there are
three basic questions of concern:
The answer to question | is not necessarily. In order for F(s) to be the transform
of some function, it must possess certain continuity requirements and behave
suitably as s—>00, as stated in the following theorem.
188 CHAP. 6 / THE LAPLACE TRANSFORM
The proof of this theorem follows that of Theorem 6. | and is left to the exercises.
The real significance of the theorem is that if F(s) is any function such that lim so
F(s) # 0, then it does not represent the Laplace transform of any regular function
f(t). This condition immediately rules out many functions as possible Laplace
transforms, such as polynomials in s, e*, cos ks, and so forth.
To respond to question 2, we observe that a discontinuous function can be
transformed into a continuous function by use of the Laplace transform. If f(f) and
g(t) are two identical functions except for a finite number of points, they will have
the same transform, say F(s). Hence, either f(#) or g(t) can be considered the inverse
transform of F(s). That is to say, the inverse transform of a given function is
uniquely determined only up to an additive null function (Lerche’s theorem). Since
null functions are normally of little consequence in applications, the apparent
difficulty of finding unique inverse transforms is of no practical concern.
Remark. A null function n(f) is one satisfying [; n(u) du = 0 for all t¢.
0
&(t)
Figure 6.4
1
yp
SEC. 6.4 / INVERSE LAPLACE TRANSFORMS 189
Hence, either | or g(t) can be considered the inverse transform of F(s) = 1/s. When
a choice can be made between a continuous function f(f) and a discontinuous
function g(f) to represent the inverse Laplace transform of F(s), we always choose
the continuous function. There can only be one continuous inverse transform in any
situation.
In answer to question 3, we find that for many routine problems the inverse
Laplace transform can be obtained directly from existing tables of transforms. (A
short table is provided in the last section of this chapter.) Also, many of the
operational properties used in finding the transform itself can likewise be used in
constructing the inverse transform. For instance, the linearity property for inverse
transforms reads
et 3s + 7 = =1 KY FE
(have Site 1 V5
a oe a oe gee |e
ee
= 3cos V5t + Gesin V51.
e~*(cos 2t — 4sin2r).
a
eee
190 CHAP. 6 / THE LAPLACE TRANSFORM
Ets) a),
Q(s)
where P(s) and Q(s) are polynomials in s. The inverse transform in such a case can
most easily be effected by representing F(s) in partial fractions. The partial fraction
representation is the same as that found in the calculus, for example, as a means
of integrating certain rational functions. It is assumed that P(s) and Q(s) have no
common factors and that the degree of P(s) is lower than that of Q(s). Let us
illustrate the technique with some examples.
OR eas |
ENS) a
SitaramS:
Gg = 6% Ky Gg = |! Suatael
=O ee.) Cae
Solution To determine the inverse transform, we first assume an expansion of the form
1 eA BsiaC
(s +1)? +1) st+1 57417
Clearing fractions gives
1=A(s?+1)+ (Bs + C\(s + 1).
Setting s = —1, we find A = 3, and equating the coefficients of s* and s° yields the
equations
SEC. 6.4 / INVERSE LAPLACE TRANSFORMS 191
QO=A+B, 1=A+C,
: =f Ee
EXAMPLE 19 Find & ee
e 5}
| | +
York: Wiley,
*See, for example, E. Kreyszig, Advanced Engineering Mathematics, 4th ed. (New
1979).
192 CHAP. 6 / THE LAPLACE TRANSFORM
EXERCISES 6.4 _
In problems 1-12, determine the inverse Laplace transform of each function using
Table 6.1 on p. 217 and various operational properties.
h) F(s))=4 2. FS) = 55
Pt FO - Go 4 =a
5) Fs) =a bh ilinerer rie
7) Wore TR 2 Oe ae
(9 Fo = Gea) Ke soxyaae
( 11; F(s) = (ess 12. F(s) = >
In problems 13-24, evaluate the inverse Laplace transform by the method of partial
fractions.
Ming a
“23. Pe ge?
25.\ Given that F(s) = L{f(}, show for constants a, b, and k, that
1
(ers) =
Oso (508 OS oo]
ane If it is known that Lo '{s~'/2e—
15} = (at)7
"7 cos (2t'”), find
Sia hype eaAsal ts ee WY
Hint: Use problem 25.
a .
*30. Show that £ tg(: ‘)|= es
y= t
Hint: Write
lo § se 1 =p als:
Bs aT AGE Sy in
and use
l+x el axe
os(+ =")=2 rere ea
Hest
Find the inverse transform of the series and sum the result.
ae G =i * en = et
(a) Show that by taking Laplace transforms of both sides, we formally obtain
F(s) = [@ + 52)"dx
(b) Evaluate the integral in (a), and by taking the inverse transform of the result,
deduce the value of f(t).
*35. Using the technique of problem 34, show that
a 1/2
(a) Ka
he po eae!
ed 7
sae Pm on pe As
194 CHAP. 6 / THE LAPLACE TRANSFORM
= a \12
(c) [ xo * sintxdx = (z) pam ios 0;
0 ; 2t
*36. Prove that lim sF(s) = lim, f(o.
The Laplace transform is a powerful tool for solving linear DEs with constant
coefficients—in particular, initial value problems. Let us illustrate the method with
a simple example.
Solution We first apply the Laplace transform to both sides of the DE:
Ly'(O} = Le}.
From the tables we find that £{3e'} = 3/(s — 1), and making use of Theorem 6.4,
we get
sY¥(s) —-7 = es
. suit
and solving for Y(s) yields 7
3 7
Y(s) = +—.
(s) SS abd) os
Expressing Y(s) in terms of its partial fractions, we get
Yi) = ——
Sua el
+2,as
and hence the solution of the original problem is symbolically ‘given by
y(t) = 3e’ + 4.
ee
EE ee eee
SEC. 6.5 / SOLUTION OF INITIAL VALUE PROBLEMS 195
Although this example was rather trivial and could be solved more easily by
another method, it does illustrate the basic features of the transform method. The
usefulness of the Laplace transform rests primarily on the fact that the transform of
the DE together with the prescribed initial conditions reduces the differential
system to an algebraic equation in Y(s). Such an algebraic equation is readily
solved, and the inverse transform of its solution then yields the solution of the initial
value problem; i.e., y(t) = L'{Y(s)}. Furthermore, the solution of the DE satis-
fying certain initial conditions is found directly without first producing a general
solution and then solving for the arbitrary constants.
and thus
y(t) = ages°+4
= |
EXAMPLE 22 Solve
y" — 6y + Sy = Te~, 2 ot
OPS Tes
Hence,
196 CHAP. 6 / THE LAPLACE TRANSFORM
It may be of interest to see exactly where each input parameter ends up in the
transform domain by considering the general second-order DE
Here it is interesting to observe that the solution (14) has naturally split into two
parts—the function y,, which is a solution of the initial value problem
a ee eee ee ee ees ee ee ee Ve
EXAMPLE 23 Solve y” + y"”—y'’—y =9e", y0)=2, y'(0)=4, y"(0) =3.
Whj=— 6. 1 6 + te!
= mee 2 =
2m’ m
so that the governing equation now takes the form
y" + 2ay' + b*y = 0, y0)=yo, y'(O) = vo. (19)
Because of the nature of the parameters involved, it is clear that a > O7, and we
may choose b > 0.
*We use a lowercase letter here for f(t) in order to be consistent with our Laplace transform notation.
ta = 0 only in the absence of the damping term, i.e., when c = 0.
198 CHAP. 6 / THE LAPLACE TRANSFORM
ip Krac?)
nie ee Pe Se 21)
It is here where the benefits of using the Laplace transform become evident. In
order to invert the expression in (21), we need to know whether the denominator
s’.+ b* — a? has distinct real factors, has equal factors, or is the sum of two
squares. These three cases correspond to the following:
(vo + ayo)
=e “a(ncoshat + sinh x : (22)
Case II: Critically damped motion results when b = a, and under this condition
y(t) = oe ae (Vo ae
S Sig
Case III. Oscillatory motion can occur only when b > a, which means that the
system is underdamped. Here we find it convenient to put
be--q@e= (en:
SEC. 6.5 / SOLUTION OF INITIAL VALUE PROBLEMS 199
2 Uo +
=e *(sncosLiss. a URL ut) (24)
uv
The reader should compare all three solutions with those obtained in Section 5.2.3
for each of the three cases of damping.
When the forcing function f() is not identically zero, the Laplace transform
method leads to the solution form
EXAMPLE 24 Determine the response of a spring-mass system that is initially at rest and then at
time t = 0 subjected to the sinusoidal forcing function f(t) = P cos wt. Neglect
damping effects.
with solution
Ps/m
Y(s) =
(So a) srk)
No resonance: If w # wo, a partial fraction expansion yields
S As Baw Cs ete)
(tosis ee on sk | sw.
from which we find A = —C = 1/(w3 — w?) and B = D = 0. Hence
yy) = LY(s)}
P
Se ae ay (COS WF a COS Wot).
m(wo = )
whose fre-
This solution represents the superposition of two harmonic motions
frequen cy w of the forcing term and the natural
quencies correspond to the angular
angular frequency w» of the system.
200 CHAP. 6 / THE LAPLACE TRANSFORM
this last result being obtained from the tables. The solution this time corresponds
to resonance in the system, since y>® as 1 (see also Section 533-2).
EXERCISES 6.5
In problems 1-20, solve the initial value problem using the Laplace transform
method.
| | |
nN & < ® =S ~— | _—
Solve this first-order DE to obtain Y(s), and invert it to find the solution y(t). (This
problem is one of the few variable-coefficient DEs for which the Laplace transform
method proves fruitful.)
*33. Apply the method of problem 32 to Bessel’s equation of order zero,
ty” + y! + ty = 0.
(di + so + s¥(s) = 0.
y =y SE ae qe (n 1)? 0 an >
where Jo is the Bessel function of the first kind of order zero. Although a second
solution of Bessel’s equation can be found by other methods, it is not bounded
at t = 0 and so does not arise in the transform method. Because it is not
bounded at t = 0, it is discarded anyway in most applications.
The Laplace transform has the effect of “smoothing” discontinuous functions in the
transform domain. Thus, one of the most interesting and useful applications of the
transform method is in solving linear DEs with discontinuous or impulse forcing
functions. Equations of this type are commonplace in circuit analysis problems as
well as in some problems involving mechanical systems.
In order to effectively deal with functions having finite jump discontinuities, it
is helpful to introduce the Heaviside unit function, or unit step function as it is often
called. We denote this function by the symbol h(t — a) and define it by
202 CHAP. 6 / THE LAPLACE TRANSFORM
On) =a
mea) ={ (26)
len fA Ose
where a = 0 (see Figure 6.5).
EXAMPLE 25_ Sketch the graph of f(t) = A(t — a) — h(t — b), a <b.
Figure 6.6
Suppose we have a function f that has nonzero values only on the interval
Gai <b (ricure 6.7)... that is,
Hove ee Wareb
0, otherwise. (27)
In terms of the Heaviside unit function, we can write
FO ie ft.
(Figure 6.9) in terms of the Heaviside unit function.
Figure 6.9
The Heaviside unit function is also useful in translating a function to the right
a distance of a units. For example, the function y = g(t) is shown in Figure 6.10,
while the function y = g(t — a)A(t — a) shown in Figure 6.11 represents a trans-
lation of g by a distance a in the positive f direction.
Calculating the Laplace transform of h(t — a), we find directly
204 CHAP. 6 / THE LAPLACE TRANSFORM
{h(t -— a} = i“e-h(t — a) dt
[ Ge at
a
or
= [eee ade
‘CO
i eal e * 9(x) dx
e *G(s),
ft) =
Hahaee
IM fe bi 2s
SEC. 6.6 / DISCONTINUOUS FUNCTIONS 205
Hence,
SG] inns oA t*}
2, pal | 4
=4-e(+5+4),
a) Sree Oe ese
ek a
6.9 in
Another way of solving Example 27 is to express the result of Theorem
the form
Fig Ones — a} =e “L{g(t + ads,
3)
x/+ \=> $44 (3 bp(4elt
and then it follows that
L{F(O} = L{t7} + eB =-(t + 24
= Ge +e eae ie 4 ateae ae
yh
Sy
aan os eta 3)
(1+44 eee
first writing 3 — t? asa
Using this latter technique, we avoid the necessity of
function of t — 2.
si ne
es a
EXAMPLE 28 Find the inverse transform of
1 — 3e>8
F(s) => ager as
Bere ecet S|
] ¢ —5s
= 8% —3s
Il pte ie ae
206 CHAP. 6 / THE LAPLACE TRANSFORM
or
ve fo TOSS ie
Oi 1S aie
EXAMPLE 29 Solve the initial value problem
fd =
Solution We can interpret this problem as a spring-mass system (no damping) that is at rest
until time ¢ = 0 and then subject to the sinusoidal forcing function sint until time
t = q/2, after which the forcing function is removed.
Writing f() = sin[1 — h(t = 7/2)] and finding
= I ame 2 Ft 1COS7
set] { }
1 se 7/2
s+ 1s? +1’
it now follows that
Thus,
me & 1 e % ge m2
erat fe + ‘al 4 le + a
car LAN ME take awa 7 TT. \e 1
7 (sine tcos f) sH(t 5)(+ ~ $)
sin(s - 2)
or, equivalently,
SEC. 6.6 / DISCONTINUOUS FUNCTIONS 207
Line 7
=(Sinit
tat? COS) a Olt
— —
- 2
MO
ae Daa poe
Zz 4 : ok
Observe that the amplitude of vibration becomes steady as soon as the external
force is removed. Otherwise resonance would take place, since the forcing function
has the natural frequency of the system.
Closely associated with the Heaviside unit function is the impulse function
5(t — a), also called the Dirac delta function (see Section 5.6). For t # a, the
derivative of A(t — a) is clearly zero, and at t = a this function doesn’t have a
derivative in the usual sense because of the discontinuity. However, in a gener-
alized sense the derivative can be defined in terms of the impulse function. Consider
the following argument.
Since L{h(t — a)} = e~%/s, it follows from Theorem 6.4 that
dh
Fe = a|= sf{h(t —at=e™. (32)
h (34)
tC =10)=10(i—aG)s
The interpretation of (34) is that the unit step function has a jump discontinuity
of unit magnitude at t = a. This is not too surprising since we defined it that way!
The significance of the result, however, is that it allows us to generalize the concept
of differentiation to include functions with finite discontinuities. Wherever a dis-
continuity occurs in the function, a delta function occurs in the derivative of this
function, and moreover, the magnitude of the jump discontinuity appears as a
multiplicative constant of the delta function.
Solution We previously solved this problem in Section 5.6 (Example 12) by using the
method of Green’s function. Now, taking the Laplace transform of each term in the
equation, we get
s7¥(s) +: ¥(s) = e0™
with solution
y( = Cae
oe s? +17
Using the result £-'{1/(s* + 1)} = sint, we have
7 3m Sr Tt ee
Figure 6.12
EXERCISES 6.6
In problems 7-14, express each function in terms of the Heaviside unit function and
find F(s).
ps
—35)
19.
é
Be, = Gn 0 ; WS;
ee
s°+4
é so = s(1 a e =) he (Stee Chae
21 ehs) = any ee Os yar erer
Ae Oise
t <2
Poti fy 0) iO (0) = 02 awhere f(D ( + 2, ipSD
210 CHAP. 6 / THE LAPLACE TRANSFORM
cos4t, Ost<7
27. y"+4y =f), y(@)=0, y’(0)=1, where f(#) = |
Os 6 Sar
28. y" + 4y =sint — h(t — 2m)sin(t — 2m), y(0)=0, y'(0)=0
29. y" — 5y'+ 6y =A -— 1), yO)=0, y'(0)=1
30. y"+2y’+2y =d(t-—m, yO)=0, y'0)=0
31. y”+2y'+2y =6@¢ -—7m, yO)=1, y'0)=0
32. y" + 4y = 6 — m — 5(t —2m), y()=0, y’0)=0
33. y" +2y’+y =&) + h(t —27), yO)=0, y'0)=1
34. y"+y =6(t — mcost, y(0)=0, y’0)=1
35. A spring-mass system at rest until time t = 0 is then subjected to the sinusoidal force
f() = Psinwt. At time t = 7 seconds, the mass is given a sharp blow from below
that instantaneously imparts an upward impulse of 5 units to the system. Neglecting
damping effects, describe the motion of the system.
In applications we must often find the inverse transform of a function that is the
simple product of two other transforms. Unfortunately, it happens that
L-NF(s)G(s)} # LF (s}L-Y{G(s)}.
We will presently show that the inverse transform of this product is equal instead
to the convolution of two functions f(t) and g(t).
The Laplace convolution of two functions f(t) and g(t) is defined by
(f*g)(t) = [« — u)sinkudu
t 1 t
= ratil = cos kt) ae
in ki Eepoos kt
SR
k jsinkt),
whereas
1 fia. Lage
= yasin kt cos kt + sin’ kt = point = 7300sktsin kt + 7.0087kt
l Ih. 2
= i(:= isint).
Other properties that readily follow from the definition are the following, the
proofs of which are left to the exercises:
f¥(kg) = (kf)*g = k(f*g), & constant (37)
f#(g +h) = f*g + f*xh (distributive law) (38)
f*(g *h) = (f*g) *h (associative law) ©?)
The important result we need is that concerning the Laplace transform of the
convolution. Let us consider
Figure 6.13
Solution Let us select F(s) = 1/s” and G(s) = (s? + k*)~'. From the tables, we find
1
oe = f(t) =t.
Thus we write
ny i
x ee rk ifmM Ae
= [« 2 te
0 k
SEC. 6.7 / THE CONVOLUTION THEOREM 213
The Laplace transform also turns out to be a useful tool for constructing the
one-sided Green’s function (see Section 5.5) for constant-coefficient DEs.
Consider the initial value problem
EXAMPLE 33 Find the one-sided Green’s function associated with the differential operator M =
Da 2D irs:
which leads to
1
k(t) = on On
Hence,
I
k(t — 7) = gi(t,7) = Bo se 77).
214 CHAP. 6 / THE LAPLACE TRANSFORM
EXERCISES 6.7
1. [« — u)sin2udu 2. [« — u)’e du
av 0
In problems 5-10, find the inverse transform of each function using the convolution
theorem.
15. Given that £{f(p} = (s? — a*)~', evaluate [fe — u)f(u) du.
10)
16. Show that[Jo(1 — u)Jo(u) du = sint, where Jo is the Bessel function of the first kind
_ of order zero.
Hint ETD ls eye
17. Solve the integral equation* for y(t),
*The equation receives its name from the fact that the unknown function y appears under an integral.
SEC. 6.7 / THE CONVOLUTION THEOREM 215
where I’ denotes the gamma function (see problem 26, Exercises 6.2).
(b) Take the inverse transform of F(s) and establish the formula
Tol
; et — uy ws 'du = aS
bes .
m — 1)!nte"""
22. Using the convolution theorem, show that the solution of the spring-mass system
my ky — J), (0) = Oey (0) = 0;
can be expressed in the form (@o = V k/m)
23. Using the result of problem 22, determine the response of the spring-mass system
when the forcing function is given by
(a) f(t) = P (constant).
(b) f(t) = Pcosat, w F Wo.
(c) f(t) = Pcos wot.
Determine the current i(t) in a single-loop RLC circuit when L = OHA Rs— 20s
24.
C = 1073 F, (0) = 0, and the impressed voltage 1s
DOA OSS
A Oo Ged
under both an
*The equation receives its name from the fact that the unknown function y appears
integral and a derivative.
216 CHAP. 6 / THE LAPLACE TRANSFORM
fO)= se [o — u) '?k'(u)
du.
7 0
Listed on pages 217-218 is a short table of Laplace transforms and their inverses
that commonly occur in applications.
SEC. 6.8 / TABLE OF SOME LAPLACE TRANSFORMS 217
Se aa)
(sod (Sue) u
S
——_—_————— (a # b)
(s)(s5 0)
1
— sinkt
k
cos kt
1
— sinh kt
k
cosh kt
1
—e sin bt
(s — ay + b* b
S Beene
ae il eee e” cos bt
(See a)’\+ b*
218 CHAP. 6 / THE LAPLACE TRANSFORM
1 1 :
20. x97 £ ptt — sinkt)
1 geen
PAN G+ aie — ktcos kt)
5 ks” nt on
225 (2 + ee pe ee <a pee Dian kt
Sa 1
D3 ware ye + ktcos kt)
S
(a* # b*) 5(cos at — cos bt)
b?— a
1
ae (sinktcosh kt — cos kt sinh kt)
1
| 26. ——— aye in Mt sinh kt
1
ee (sinh kt — sin kt)
1
1 (coshkt — cos kt)
(a > 0) h(t — a)
REFERENCES
Kreyszig, E. R. 1979. Advanced engineering mathematics. 4th ed. New York: Wiley.
LePage, W. R. 1980. Complex variables and the Laplace transform for engineers. New
York: Dover.
Miles, J. W. 1971. Integral transforms in applied mathematics. London: Cambridge Uni-
versity Press.
Sneddon, I. H. 1972. The use of integral transforms. New York: McGraw-Hill.
Spiegel, M. R. 1965. Laplace transforms. New York: Schaum.
The first five sections of this chapter are devoted to solving linear systems of differ-
ential equations with constant coefficients. Several methods can be employed to :
solve such equations. :
In Section 7.2 we introduce a method that changes the system of equations to a f
single DE by successive elimination of the unknowns. The technique can be greatly i
systematized with the aid of Cramer’s rule. We use the Laplace transform in Section
7.3 to reduce the differential system to an algebraic system solvable by elementary
techniques and then apply the inverse transform to produce the desired solution set.
Another method, presented in Section 7.4, consists of assuming a particular trial
solution form that, when substituted into the system, reduces it to a linear system of
algebraic equations. Solving the algebraic system identifies the parameters occur-
ring in the trial solution set. Most of the general theory concerning linear differential
systems is also discussed in this section. In Section 7.5 we briefly discuss the use of
matrix techniques that permit the system to be expressed as a single-matrix DE ina |
form similar to that of a first-order linear equation. J
The qualitative aspects of certain kinds of nonlinear systems are discussed in Sec-
tion 7.6 without any explicit, formal solution functions. The general questions of con-
cern are mainly associated with the idea of stability of a solution. In particular, it is
important to know whether small changes in the input data (initial conditions) pro-
duce only small changes (stability) or large changes (instability) in the output (solu-
tion). When a nonlinear system is approximated by a linear system, it is equally im-
portant to know whether such an approximation is “reasonable.” That is to say, some
nonlinear systems do not iend themselves to reasonable linear approximations.
220
Wal INTRODUCTION
In applied mathematics, many applications involve not one but several dependent
variables, each of which is a function of a single independent variable, usually
time. The formulation of such problems leads to a system of differential equations
rather than a single equation, as we have studied thus far.
For example, the motion of a single particle in space governed by Newton’s
second law of motion leads to the vector DE
—_ =F, (1)
which is equivalent to the system of scalar DEs
mx" = Py,
my is, (2)
mz" = F3,
where m is the mass of the particle, atl he y, Zz) is the position vector of the
particle, and F = (F\, F, F3) is the external force acting on the particle. Systems
of equations are also used to describe the motions of coupled spring-mass systems
and RLC circuits connected in parallel. Examples of such systems will be given
later.
Systems of equations arise in the classical ecological problem of the prey and the
predator, which involves the struggle for survival among different species of
animals living in the same environment. One kind of animal eats the other as a
means of survival, while the other develops methods of evasion in order to avoid
being eaten. For example, suppose x(t) denotes the population of rabbits at any
time on an isolated island, and y(t) the number of foxes at any time on this same
island. The foxes eat the rabbits, and the rabbits eat clover, which is in ample
supply. When the rabbits are plentiful, so too are the foxes, and their population
grows. When the foxes become too numerous and eat too many rabbits, they enter
a period of famine and their population begins to diminish. The rabbits eventually
become relatively safe again, and their population increases, which also initiates
new increases in the number of foxes. Under the proper ecological balance, these
cycles of population increases and decreases can be repeated incessantly. On the
other hand, if the balance of nature is disturbed in the right way, both species could
die out.
Problems of this nature were independently modeled by A. J. Lotka (1880—
1949) in 1925 and by V. Volterra (1860-1940) in 1926. They suggested that the
instantaneous populations x(t) and y(t) of both species are solutions of the system
of equations
Xx
,
D682 16 Bes 3)
Vos ,
Gynt OXY.
where a, b, c, and dare all positive constants. The constants a and c represent the
growth rate of the prey (rabbits) and death rate of the predator (foxes), respectively,
221
222 CHAP. 7 / SYSTEMS OF EQUATIONS
whereas b and d are measures of the effect of the interaction between the species.
These equations are now widely known as the Lotka-Volterra equations.
Finally, another way in which a system of equations can arise is to convert an
nth-order DE to a system of first-order equations. That is, given the nth-order DE
(not necessarily linear)
Lp = J t= Yo, ts SS ie a es
Hence (4) is equivalent to the first-order system of equations
X| = X2,
X2 = X3,
Be Pa ee (5)
Site Ane
Xe UL a ee re Ie
Before proceeding with this section, it may be helpful to review the operator
notation introduced in Section 4.4.1.
Let us consider the simple system of equations
Key =. OF
y' + 4x =0, oI
or, equivalently (using operator notation D = d/dr),
Dx —y =0,
4x + Dy = 0. Om
Suppose we operate on the first equation by D to get
D*x — Dy = 0,
4x + Dy = 0.
Adding the two equations now eliminates y from the system. Hence,
(D* + 4)x =0
with general solution
X(t) ="Cncos 21+ Gorsini ce (8)
In a similar manner we multiply the first DE in (7) by 4, operate on the second
DE by D, and then subtract the results to eliminate x from the system. This action
leads to
SEC. 7.2 / THE OPERATOR METHOD 223
(D? + 4)y = 0,
from which we get
Wit) =1G a COSe2? a Ga sin. Zt: (9)
Although (8) and (9) represent solutions of the system (7), they are not solutions
for every choice of the constants C,, C2, C3, and C4. Therefore, it is necessary to
find a proper relationship between these constants. To do this, we substitute (8) and
(9) into the first equation in (7) to obtain
or
(2Ce Gs) cos 2b = (2G Cy) Sin 2 = 0;
from which we deduce C3 = 2C> and Cy = —2C,. Hence, the solution of (7) is
given by
x (f) = Ci. Cos: 26.4 Cousin: 2h,
(10)
y(t) = 2C, cos 2t — 2C, sin 2t.
In order to systematize the procedure we just used, let us now consider the
general system of equations
where g}(t) = Lal f,()] — L.{fx(d]. In a similar fashion, the elimination of x gives
where g2(t) = LiLfa(d] — Ls[fi(O]. Equations (12) and (13) can now be solved
independently.
The above procedure can be systematized even further by formulating it in
determinant notation. Observe that if we write
L, I, fi@® 14
Ee ? (14)
fo) Ls
which is precisely what Cramer’s rule would yield for an algebraic system, then the
expansion of the determinants on each side reduces (14) to (12). However, it is
important to keep in mind that the determinant on the right-hand side has the proper
meaning only if the operators L, and L, operate on f(t) and f(t), respectively. That
is, the expansion of this determinant must produce the function g,(t) defined above.
Likewise, we can express (13) as
224 CHAP. 7 / SYSTEMS OF EQUATIONS
LT, Ly Ly fit)
(15)
Learns L3 fxd)
The use of Cramer’s rule to solve systems of equations can be applied to systems
of any number n of equations, regardless of the order of the operators L;, I2,. . .,
L,, . . . , Lon. From a practical point of view, it works best on systems of two or
three equations with either first- or second-order operators, since the amount of
computation becomes unwieldly for larger systems. Also, we might note that the
auxiliary polynomials for x and y are always the same, since the determinants on
the left-hand sides are identical [see (14) and (15)].Therefore, with the exception
of the arbitrary constants, the homogeneous solutions (complementary functions)
of x and y are identical. The number of independent arbitrary constants in the final
solutions should be equal to the order of operator determinant on the left-hand
sides. For a 2 X 2 system, for example, this is the order of L,\L4 — L5L3.
Remark. If L,L; — L,L3 = 0, the system either has infinitely many solutions or
no solutions, depending upon whether the determinants on the right-hand sides of
(14) and (15) vanish. Thus the situation here is similar to that of a system of
algebraic equations.
orem (De 2) ys 08
yi Cies 5Cre"
We might point out here that instead of solving both homogeneous DEs forx and
y, we could have solved only for, say, x, and then used the first DE of the system
in the form y = }(x’ — x) to find y. Such observations can be helpful in some
problems.
eye c
D—-2 = || eee — |
tery" AK, =) 4
D—2 =| Die eart
Shewd es es awe;?
After expansion of the determinants, we find
(D?— 4D +3)x =2?—2t + 1,
(D2? — 4D + 3)y = —2t? + 3¢.
The homogeneous solution for x is
Xy(t) = Ge ap Ge
Se ee
AU Pea yom
Hence,
(CS ip (Eee)
1 2 11
= Cie
Cie’ + Cre
Ce* + 3!
or + =t
9! + —.
7
Rather than solve the nonhomogeneous DE above for y, we merely observe that
y =x' — 2x — t, obtained from the first DE of the system. Therefore,
he yal BE
2
ae fp
g
at pence!
16
y(t) Cient Cee 3 9 7
——————
EXERCISES 7.2
1. x= 2x —y 2. x’ = 4x — 3y
y' = 3x — 2y y' = 5x — 4y
3. x' = 4x — 3y 4. x' = 3x — 18y
y' = 8x — 6y y' = 2x — 9y
5. x’ =x— 4y 6. x’ = 3x — 2y
y’=xty Yo Seleoy
7. x’ = 6x — 5y 8. x’ = 3x + 2y
y t= erry yas Ix ay
9, x’ =x —y + 42 "10. x’ =-x+2
A eh Ga Pay Soars y'=-ytz
PgNec DBE* ers 2 =x ey
11. Find the general solution of
x' = ax + by,
y' = bx + ay,
given that b # 0.
In problems 12-28, solve the given system of equations. Make sure the proper
number of constants appears in the general solution.
In problems 29-31, show that each system is degenerate in that it either has no
solution or infinitely many solutions.
ie, tO eka, lO
x exe
yy =-0
If initial conditions are prescribed along with a system of differential equations, the
Laplace transform method is generally more convenient to use than the operator
method of the last section.
The Laplace transform applied to a differential system reduces it to an algebraic
system in the transformed functions. Standard solution techniques can then be
applied to solve the algebraic system, and the inverse transform of the algebraic
solution functions will produce the solution functions of the original differential
system.
nn nn UIE IIEEEEEEIUIEEStnSa a
2
s°Y(s) — 4Y(s) — 2sX(s) = are
XG)(s) =e§e.1 5
gg? eT
—4
st Ae
Y(s) =. = = ohcan
S ge se Il Ger agai
Now,
x 1
x(t) = -Y pees
i {| si SL Be1 {=1 Fan = I 2
‘} 2 {=
Gland
Sine fi oasin were
and, similarly, we find
having spring constants k, and k,, respectively (see Figure 7.1). Let y; and y2 denote
the vertical displacements of the masses from their equilibrium positions. When the
system is in motion, the stretching of the lower spring is y. — y), exerting a force
of k,(y2 — y,) on the upper mass m, (Hooke’s law). Also, the upper spring exerts
the force —k,y, on this mass, and so, invoking Newton’s second law of motion, we
have
Hence, the coupled spring-mass system in the absence of damping effects and
external forces is governed by the system of second-order DEs
mMy\ (+ (k,
(ky + ky)y,2)Y1 — kyr
22 = 0,7 (18)
ct i a ea ns ne Snr EEE
EXAMPLE 4 Determine the free motions of a double spring-mass system composed of two unit
masses and springs with constants k, = 6 and k, = 4. Assume that both masses
unit initial
start from their equilibrium positions, but that m, has a downward
velocity and m, an upward unit initial velocity. Neglect damping.
proper
Solution The system here is described by Equations (18) above, and when the
parameter values are written in, we obtain
Vt LOV gee: Ayo 0,
yz — 4y, + 4y2. = 0,
230 CHAP. 7 / SYSTEMS OF EQUATIONS
where i), i2, and i; denote the current in each part of the
circuit. However, by
Kirchhoff’s current law,
SEC. 7.3 / THE METHOD OF LAPLACE TRANSFORMS 231
1 = 1» ap 13,
EXAMPLE 5 LetR = 1009, L =4H,C = 10% F, and E(t) = 100 V in the network of
Figure 7.2. If initially the currents i, and i, are both zero, find the currents at all
later times.
Solution Substituting the appropriate values of the parameters into (20), we have
4i; + 100i, = 100,
i, + 100 — i) = 0,
with i,(0) = i,(0) = 0. The Laplace transform applied to each equation gives (after
simplification)
25
sI,(s) + 251,(s) = rae
—100/,(s) + (s + 100)/,(s) = 0,
where J,(s) = L{i,(t)} and 1,(s) = L{i.(D}, and solving these equations, we find
Hence,
HG) ea SOR Oe
pusala Clem ea,
232 CHAP. 7 / SYSTEMS OF EQUATIONS
and
EXERCISES 7.3
In problem 1-14, solve the system of equations by use of the Laplace transform.
1. X=
— ey (0) 0 2. x’=y, x(0)=
y =2x, yOu I y'’ =x, y(0)=0
3. c= yx)= —I 4. 25 4r 2ysex 0)= 2
V part
oe ¥en WU) ae Y= ori 2y, yO)
= 2
59) AP ea Na era | 39(8) Pe
y = 2x +33y +e") yO) =.0
x’ — 44 +2y =e", x(0)=
Va Ox
ly it (0)
=0
2X VSy= ty x(0)=
ictal oemmy (\e—10)
Dea Via re= seer (0) =O
AC heya 13y =92 ay (Oy 0
Xe!xe yu= Ox (0) =0.- x'(0) =
yoriye«= 0, .y(0)=:0, Fy (0), =
x) ay + y" ="6 sin) x0) 0
Lia 2X — 0% Sy(O)e=nOey (0)
= Onn oyOn=1G
Lee Dey
fat 6 ey (en ame (0) nO
A ay) Gye ae (0)
Kiet Oy tery (1), ox(O)=0 er (0) 0
yo Ay = 2x!’ = 0, —y(O) = 0, y'(0)=0
oO, OS %7<—2
where fo = {0 Aas
17. Show that the spring-mass system in the accompanying figure is governed by the
system of equations
yi + 2ky; — kyo = 0,
y2 + Qky2 — ky, = 0.
Problem 17
18. Solve the system of equations in problem 17 if k = 3 and the initial conditions are
yi(O) = 1, yi(0) = 3, y2(0) = 1, y2(0) = —3.
19. Find the currents i;, i2, and i3 in the network in Figure 7.2 when R = 1000, L = 1
H, C = 10°* F, and E(t) = 100 V. Assume that i, and i are initially zero.
20. Find the currents i;, i2, and i; in the network in Figure 7.2 when R = 100 Q,L =4
H, C = 10°* F, i(0) = i,(0) = 0, and
(ay) EG = 100er Ov: (b) E(t) = 100sin607tV.
21. Show that the network illustrated in the accompanying figure is governed by the
system of equations
i,(t)
L,
R,
Problem 21
234 CHAP. 7 / SYSTEMS OF EQUATIONS
Problem 23
*24. Solve the system of equations in problem 23 when the initial conditions are
|
9(0) = 0, 60) =0, 6,(0) = 5 (0) = 0,
and m, = 3, m2 = 1, and b, = bo = 16.
In this section we wish to discuss some of the general theory associated with linear
systems involving first-order DEs. Although the results can be generalized to
SEC. 7.4 / FIRST-ORDER LINEAR SYSTEMS 235
systems of n equations, we will find it convenient to develop the theory for systems
of two equations. Therefore we will be interested in the general nonhomogeneous
system
x’ =ay()x + any + fd,
| (21)
y= an()x + an(ty + fr).
By the associated homogeneous system we mean
"
ic Vi
k
pas + ff)
t '
m m m
and then let y = x,, y’ = x2. Thus, since xj = y’ = %2, we have
x} = X2,
c k fd)
SO hay ier Noel 5) meme
Most linear systems that are not first-order systems can also be rewritten as
first-order systems but will generally involve more than two equations. To accom-
plish this it is usually best to solve for the highest-order derivative in each of the
unknowns before introducing new variables. Let us illustrate with an example.
een ey ele wenn Ae ae es ee
x" — x! te y! Gs en
y=
6x + 4,
and introduce u = x, v = y, w = x’. Thus, uw’ = w and
w’ = 2w + v! — e%,
vo' = 6u +f,
or
u'=w,
v' = 6u +4,
w' = 6u + 2w +t — e”.
Theorem 7.1 | If the functions a;;(0), ai2(t), an,(0), an(t), fi(), and f,(1) are continuous in some
interval a S t < b containing the point fo, then there exists exactly one solution
pair {x(t), y(t)} of the system (21) such that x(t) = xo, y(to) = yo.
Theorem 7.1 is the basic existence-uniqueness theorem for the system (21),
which is comparable to Theorem 5.1 for a single DE.
The superposition principle, which is so important in the study of linear ho-
mogeneous DEs, also applies to systems of DEs as stated in the next theorem. The
proof is left as an exercise (problem 33).
Theorem 7.2 | /f {x,(t), yi(} and {x2(2), y2(t)} are both solution sets of the homogeneous system
(22), then the pair
EXAMPLE 8 The function sets {e~', —e‘} and {e*, }e“} are each solutions of
yh Aa
yous On ar 2s
Verify that x(t) = Cye~! + Cre” and y(t) = —Cye ' + $C e™ are also solutions.
Solution Direct substitution of x(f) and y(t) into the first DE gives
x — x = 2y — —Gier ot 4G se* — Ce * = Coe 4°2C
ie | — 3Che™
= 0.
and, similarly, for the second DE,
Definition 7.1 | The solution sets {x,(t), y\(t)} and {x(t), yo} are said to be linearly dependent
on some interval a <t < b if and only if there exist constants C, and C2, not both
zero, such that
C,x,(t) + Cox2(t) = 0,
Ciyi) + Cry2(t) = 9,
Cy = 0, we
for every t in the interval .* If these relations are true only for C, =
say the solution sets are linearly independent.
ek e ee ee
e
se
dependent on every
EXAMPLE 9 Show that the sets {e”, 2e’} and {—4e*, —8e'} are linearly
interval, while {e*, 2e"} and {—4e”, e’} are linearly indepen dent on every interval.
2C,e' mF 8Ce' = 0,
2C,e' + Coe’ = :
Theorem 7.3 If {x,(2), y,(t)} and {x2(t), y2(t)} are solution sets of the homogeneous system (22)
on some interval a <t <b, a necessary and sufficient condition that these
solution sets be linearly independent is that the Wronskian
Wi =
X(t) x(t)
Oe pate
never vanishes ona St Sb.
Remark. As was the case in Chapter 4, the Wronskian defined in Theorem 7.3
can be shown to be either identically zero on the interval a < t < b or never
vanishing there. Also notice that our present definition of the Wronskian does not
involve differentiation.
Based on the above results, if {x,(4), yi()} and {x2(), y2(} are linearly indepen-
dent solution pairs of the homogeneous system (22) on some interval a <t <b,
then we define a general solution of this system to be the solution pair {xj (t), yy (O},
where
EXAMPLE 10 Verify by the Wronskian test that the solution sets {e~', —e~} and {e*, 4e* given
in Example 8 are linearly independent, and write the general solution of the system
Xe tx ae Dy
y = 3x0 2y.
which is never zero. Hence, the solution sets are linearly independent, and from
Equation (23) the general solution is
Finally, with respect to the nonhomogeneous system (21), we state the following
theorem, which is analogous to Theorem 4.9. The proof also parallels that of
Theorem 4.9 and is left to the exercises (problem 34).
Theorem 7.4 If {xp(t), yp()} is any solution pair of the nonhomogeneous system (21) and {xy (0),
yx (t)} is the general solution of the associated homogeneous system (22), then the
general solution of (21) is given by {x(t), y(t}, where
x(t) = Xy(t) + xp(t) = Cixi(t) + Crx2(t) + xe,
After dividing by the nonzero factor e™ and rearranging the remaining terms, we
are left with the algebraic system
Gi NA apB = 0,
(25)
aA + (ax — AB = 0.
We wish to find values of A such that the system (25) has a nontrivial solution
is possible if and
for A and B, i.e., where A and B are not both zero. This situation
only if the coefficient determinant
Aca) = |" a
=A. a\2
(26)
fig —
For A = A,, we substitute this value back into the algebraic system (25) to obtain
the nontrivial solutions A, and B,. Hence, with these values of A, A, and B, we have
one solution set {x,(1), y,(4)}, where
x(t) = Aje*",
Case I—Distinct real roots: When the roots A, and A, of the auxiliary equation
(27) are real and distinct, there are two linearly independent solution sets corre-
sponding to {A,e*", B,e*"’} and {A,e*”, B,e'} (see problem 25). The constants A,
and B, are found in a similar way as A, and B,, but this time the value A = A; is
substituted into (25). The general solution in this case is therefore
x(t) — GAler ota C,A,e*,
=e +82 ys,
y = 3x 4+ 2y.
2A, + 2B, = 0,
3A, + 3B, = 0,
which shows that B, = —A,. For A; = l, a possible solution pair is {e~', —e~4.
Likewise, with A, = 4, we obtain from (25)
—3A, + 2B, = 0,
3A, — 2B, = 0,
3
or By = 3A). Thus, {e”, 3e*} is another solution pair, and our general solution
4 3 4 . . . . .
is
x(t) aa Gem q= Cre.
SEC. 7.4 / FIRST-ORDER LINEAR SYSTEMS 241
Case II—Equal roots: If the two roots A, and A, are real and equal, then setting
A = A, = Ay produces only one solution set of the system. By analogy with
second-order DEs, we might expect to find a second linearly independent solution
set of the form
x(t) — Aste”,
y2(t) = Byte™.
However, this is not the case. Without providing the details, it turns out that we
must seek a second solution set of the slightly more general form
eese Se
A se B, = 0.
Remark. It should be observed that the special case when aj); = a and
2 = a2, = O leads to the uncoupled systemx’ = a,x, y’ = ayy. Here {A,e™, 0}
and {0, B,e”} are linearly independent solution pairs. However, our general pro-
cedure outlined above is for coupled systems.
Case I1I—Complex conjugate roots: When the characteristic roots are complex
conjugates, A, = p + iq and A, = p — ig, we obtain two distinct solution pairs
1 l
3 ba) + x,(t)] = 5 (a1 + iaz)e”(cos qt + isingt)
and
Le Aare :
sbi) + y,(t)] = e?(b, cos gt — b2 singt).
These sums represent another solution pair of the system involving only real
functions. In a similar fashion, we see that
and
which is also a solution pair. The Wronskian of these solution pairs is found to be
W(t) = e?'(a,b> = ayb;).
For complex A, the constant B; cannot be a real multiple of A, and thus a,b, —
asb, # 0. These last solution pairs are therefore linearly independent, and the
general solution for this case can now be written as
cosgt — azsingt) + C(a,cos gt + a,singt)],
x(t) = e?[C,(a,
cos gt — bzsingt) + C(b2.cos qt + b, sin qt)].
y(t) = e? [C\(b, ee)
then
Therefore, if we choose A; = a, + ia) = 1 and By =o; 4+ 10) = M27,
a, = 1, @ = 0, b; = 1, and bp = —2. The general solution then becomes
x(t) = e*(C, cos 2t + C, sin 21),
y(t) = eriC ie 205) COS 20 (2G) + Cs) sin 2].
244 CHAP. 7 / SYSTEMS OF EQUATIONS
§
Notice that it was not necessary to substitute A, into (25) for this case. (Can you
explain why?)
The results of this section can be generalized to systems of equations larger than
two. Let us illustrate with an example.
y rz
Peace ie ee
ESE 0 1
0 tuaeN. 2 | =A? — 7A? + 6A =A(A — LIA — 6).
1 2 ae 7,"
Hence, we find A, = 0, A, = 1, and A; = 6. The substitution of A; = 0 into the
algebraic system
GloAACH= 0),
Cli HBr
2C* =/0,
At QBS iG AC =0,
yields
A, + C, =0,
B, + 2C, = 0,
A, + 2B, + 5C, = 0,
from which we obtain A; = 1, B; = 2, and C, = —1. Similarly, the substitution
of A, = | and A; = 6, respectively, into the algebraic system leads to A> = 2,
By= 1, Cyi=0; \Ag= 1) By= 2;eand Gy'= 5eiThe general solution of this
system is therefore
x(t) = C, + 2Cr,e' + Ce",
y(t) = 2G} = Cre! ote 266-3
z(t) = = = SGse a
'
x = ax + apy + f,(0),
(34)
y' = ayx + any + f(t).
For the solution technique to be discussed, the coefficients need not be constants,
although that is the only case we will solve. The general solution of this system
(Theorem 7.4) is the pair {xy(t) + xp(t), yu (0) + ye(O}, where {xy (t), yu (D} denotes
the solution set of the associated homogeneous system and {xp(t), yp(} is a
particular solution set yet to be determined.
A general method for finding a particular solution set is the variation of param-
eter method, which is similar to the technique previously given this name in Section
4.7. We seek a particular solution of the form
Xp(t) = uj(O)x\() + un(t)x2(0),
(35)
ye(t) = u(dyi() + ur(Dy2(d),
where we have replaced the arbitrary constants in the homogeneous solution by
unknown functions. The substitution of (35) into (34) gives
ae U5X2 = A U,X| se Q,\U2X2 55 Q\2uUy,y\ ae Q\2U2y2 hi
UX} == uyX) te UX)
_—— ————————
OT
Xu, + XW = fi,
Pe a
yu + yoy = fr.
Solving for uj and uw, we obtain
fi) x(t) xt) fi
a
oe th ee ee e
ee
' 2
Nyt) = Cy + 20re57,
ViDi= 2G — Cree,
Assuming a particular solution of the form
Xp (0) = u(t) + 2u,(He—*,
ye(t) = 2u,(t) — u(de™,
we find
l
- Je
t
ete
u(t) = i = 8 + 1
1 e202 ae Mn
2 -e*
ae
t
Deas : ;
S(t = dash LF HoT«
walt) = Tos ca.
2 -e*
Therefore,
8
u(t) = sé eA Oe
i) =.=
4 St
U( ) 5 ’
8 8
xp(t)
Xp(t) = <t
5 + logt -—55°
16
yp(t) = cz. Sle Peet ~.
Finally,
8 8
a)(0) S.C 1 2G 2€ bet
5 logt -—55°
- 16 4
y= 2Cp— Cre + au + 21097 a ar
e
a a S
EXERCISES 7.4
1. y"+k’y = Psinoat 2) 4y + ye
eX ey AD ey 13.0 x 3x Ly
41
va = het y Vir 2s Sy
y Oxy
14. x' = 6x — 5y 15.0 + -2y
x! 33x 162 x) = 3x = I8y
Veoh oy yi = Se aetery by aaa OY
20 ee = ee
ia oii),
Ft ea ee
21. {e%, 2e*} and {—3e%, —e"} 29. {e*, 2e*} and {—3e*, —6e"}
*24. {1+ t.—2 4 27,4 + 211, —20. 4}, and {3 4 °2t, Gi ar 2 rar|
25. Show that the solution (29) is composed of linearly independent solution sets.
*26. Show that two of the characteristic roots of
Gd ie had A mH
Vesey es,
AE nD Wacay he we
are the same but that three linearly independent solution sets can be found which are
all of the form {Ae*’, Be”, Ce*}.
In this section we assume that the reader is somewhat familiar with matrices and
their basic operations, as taught in a beginning course in matrix methods or in linear
algebra. Those readers lacking this background can skip this section without loss
of continuity.
The theory and solution techniques discussed in the last section can all
be
formulated in terms of standard matrix operations. For example, the system itself
where
x(t) fild)
Y(t)
@
=
2) a
d F(j) =
&3 fy
are both vectors.* [The form of Equation (38) suggests that the notation for a matrix
equation for a system of n equations is as simple as that for a system of two
equations. ]The use of vectors and matrices not only is notationally expedient but
also facilitates calculations and emphasizes the similarity between systems of
equations and first-order linear (scalar) equations. We say that the vector Y is a
solution of (38) provided it is differentiable and its components satisfy the system
of equations (37).
oe
d Yy = 4
aes ae)
= a8
SE
(1)
ey (
e, it follows that
are both solutions of (41). Then, from the superposition principl
the linear combination
Y = CY + GY (43)
is also a solution of (41) for any constants C, and C).
Let us now introduce the matrix
x,(t) x(t) 44
=
ie aay
eu co)
say that Y and Yo
whose columns are the solution vectors Y“) and ¥. We then
7.1 in Section 7.4). How-
are linearly independent if det (X) # Of (cf. Definition
in Theorem 7.3, 1.e.,
ever, this determinant is also the Wronskian defined
det (X) = W(t). (45)
Thus, if we have that Y“’ and Y™ are indeed linearly independent vectors on some
interval / and are also solutions of (41), then a general solution of the homogeneous
equation (41) on J is given by (43).
ME ad a (47)
where A and the constant vector E are to be determined. The substitution of (47)
into (46) leads to
Ane 4 AKeo
which can be rearranged in the form (after canceling the common factor e™’)
(A — ADE = 0, (48)
where
I = (as)
hw)
(49)
49
is the identity matrix. Now from the general theory of matrix equations, we know
that (48) can have a nontrivial solution if and only if
Theorem 7.5 For each eigenvalue A; of the constant-coefficient matrix A and each eigenvector
E” belonging to A,, the function
Y = Ee%!
is a solution of the homogeneous matrix equation
Y’ = AY.
251
SEC. 7.5 / MATRIX METHODS
e
eee re eee ee S
by matrix methods.
Solution (We previously solved this system in Example 11 of Section 7.4.1 by conventional
methods.) Here we assume Y = Ee”, which leads to the system of equations
ae 2D C1a\ ee 0
3 2 eA €7 0 ‘
can be
Thus e, = —e,, and the eigenvector corresponding to the eigenvalue A,
represented by (any constant multiple will also suffice)
e®=(1)).
= 1
ctor
In the same way, corresponding to A, = 4 we find the eigenve
()
E =
2 2
system of equations
Based upon these results, the corresponding solutions of the
are
yo = BE Me, yy = Eee)
The matrix formulation permits us to generalize the solution formula for a single
first-order (scalar) DE with a constant coefficient. For instance, the equation
yi ay (51)
has the general solution (see Section 2.4.1)
y = Ce”, (52)
where C is any constant. Likewise, the matrix equation
Y’ = AY, (53)
where A is a constant matrix, has a solution that can be represented by (where C
is any constant vector)
Y =e'C. (54)
Of course, we must now define the exponential matrix function e“’ so that (54) is
meaningful (see problem 11). This can be done, but the use of this matrix function
as a practical means of solving systems of DEs would necessitate a fairly thorough
knowledge of matrix theory. Such knowledge is considered beyond the scope of
this text.
Wr) = (ie z)
yi(t) y(t)
(56)
whose columns are the vectors Y‘’) and Y. Observe that W is nonsingular, i.e.,
det (W) # 0, since this determinant is also the nonvanishing Wronskian W(t). In
terms of the fundamental matrix, the solution of (55) is simply
Nea) C. (57)
where C is an arbitrary column vector.
In the case of an initial value problem, we seek a solution of (55) such that
C = W'0)Yo, (60)
where W>! denotes a matrix inverse.* Hence the solution of the initial value
problem described by (55) and (58) is formally given by
Y = WNW '(O)Yo. (61)
—
ee
a
Yo = |
] ee. Y” = z Jes
(
=a 3
are linearly independent solutions of the equation. Hence, the fundamental matrix
is
Caer le.
Wi) < G =I 2) i
is given by
— 1 ie nee d A) # 0
=1
ea erier:
-( —e! ae
z Ben Qe
~ \—3e7 + 3e%] ”
Wi) = AOW(D,
and therefore (65) reduces to simply
Wnu'(t) = F(d. (66)
Hence,
is69) (67)
and, upon integration, we deduce that
Y = WC + Wa iW>'()E(2)dt, (70)
SEC. 7.5 / MATRIX METHODS 255
Solution (We previously solved this system in Example 15 of Section 7.4.2 by conventional
methods.) The coefficient matrix A has eigenvalues given by solutions of
det(A — AI) = A? + 5A =0.
Be) ee)
@b) es 1 O) D
I 2
5 5)
Wa) = '
2 St I St
=e —-—e
5 5
5° ae
256 CHAP. 7 / SYSTEMS OF EQUATIONS
Sted
lecee | plead g ns
TA ee Ene 5
5
( ze spurns
=a t
Wy \oe=c ied
25©
8 8
5! t loge 75
16 4
et
es
clogt +
75
—
EXERCISES 7.5
1tgY, ,_-(3
(2 -1
ah 2aFN, ,_-(§
(4 -3
aly WY ,_ (4 3
(3 iy
In problems 8-10, find a unique solution of the system of equations satisfying the
prescribed initial condition.
SEC. 7.5 / MATRIX METHODS 257
8. Y
. — ¢cha ake
@| ¥(0) = (i)
1
9. Y Utpe &=3 tll
|
¥(0) =
(*)
9)
Diener ware 1
pLOgmevee ten Ogee |Ye Y (Oe =| 2.0
Lisaglee a
“11. By analogy with the Maclaurin series
>
: x2 eae
Oye ee Ne er ec
PA BN
we define the matrix exponential function
eee At ANIA i
GAGMe ge
C4)
to compute the matrix exponential function when ¢ = | and
(b) Compute e“’ exactly for the matrix A given in part (a) when t = | and compare
the result with that for (a).
(c) Compute e*’ exactly for any t when
sys-
In problems 13-18, use matrix methods to solve the nonhomogeneous linear
tems.
Re ps
(Ay oe ke(5 eA)
—sin2t nm =I t
13. Y -(i ie ; ‘= ats
1
0
We: Se) Sapa 1 sel beater 18. vie Uh Ble
2
problem
19. Find a unique solution of the nonhomogeneous initial value
oe
aS 10e~* oa
‘= + , YO)= :
uf ies “)¥ ec) ©) (2
20. Find a unique solution of the nonhomogeneous initial value problem
Y ee- (|
cree 3e!
bees ¥ (0) ay ih
©
*21. Verify that Equation (72) satisfies the initial value problem described by Equations
(62) and (71).
A DE provides the basic model from which a mathematical analyst pursues in-
vestigations into the existence, characterization, and construction of solutions to
physical systems. For a large class of problems, either the oscillatory nature of the
system under study or its stability is of greatest interest. But because finding the
solution of many DEs in a convenient and useful explicit form is either difficult or
impossible, we often resort to qualitative methods for much of the analysis, as we
did in Section 5.7 in studying the oscillatory characteristics of certain linear
systems. We now wish to use qualitative methods once again in studying certain
nonlinear systems, but with respect primarily to their stability.
Suppose we consider the initial value problem (primes denote derivatives with
_ respect to 7)
dy _ F(x,y)
(75)
dx yo oF
and then solving this first-order DE.
EXAMPLE 19 Determine the trajectories associated with the simple harmonic oscillator whose
governing equation is
mx takx
= 0.
y
es aE
Li a
dx my
The general solution of this DE can be found by separating the variables, which
leads to kxdx + mydy = 0, or
kx? + my? = C,
y
A
(Ww y(t))
NZ, ‘
> x
Figure 7.4
Remark. Since the parameterization of a curve is not unique, the terms solution
and trajectory are not synonymous. For example, x = e', y = e” and x =1,
y = t’ are both parameterizations of the parabola y = x’, but in the first case only
the right half (x > 0) of the parabola is defined by the parametric equations.
dy _ Qi y) (77)
dvi PO.)
the integration of which provides the integral curves or trajectories except at points
If both
for which P(x, y) vanishes. (If P = 0, we simply consider dx/dy = P/Q.)
the phase plane, that point is said to be a critical
P and Q vanish at a point (xo, yo) of
point, for then a unique slope is not defined. However, in reference to the original
um point
system (76), the critical point (xo, yo) is best described as an equilibri
there. That is to say, once we reach this point,
because both x and y are stationary
a critical point is one for which both velocity
we can never leave it. In mechanics,
and acceleration vanish, and hence the motion stops.
ee ee ae
ee
A
EXAMPLE 21 Determine the critical points of the nonlinear system
ee y
ary
»
!
ey!
x —-y’=0,
le):
we see that there are two critical points, (0, Oywand
of the system
Definition 7.2 |Let {x(t), y(O} denote the solution set
x' = P(x,y), 0) = a,
y’ = Q,y), y0)=B8,
critical point is said to be
which has a critical point at (xo, yo). The
262 CHAP. 7 / SYSTEMS OF EQUATIONS
(a) stable iffor every € > 0 there exists some number 6 > 0 such that
VS Gn 2)? Gy = yy He ate
for all t = 0 whenever
[Og ay yo By = A and
(c) unstable otherwise.
a Se a ee ee oe ee ee
EXAMPLE 22 Discuss the stability of the critical point (0, 0) of the linear system
eee) == 28
y = —2y, y(0) = 1.
Solution Because the system is linear with constant coefficients, we readily find the solution
em wen) =Thuse according to Definition 7.2, given € > 0, we wish to find a 6
such that whenever
(4 1) 5 <6,
we have
1. There exists at most one trajectory passing through any point of the
phase
plane that is not a critical point.
2. A particle starting at a point other than a critical point cannot reach the
critical
point (if indeed it reaches it at all) in a finite amount of time. If the solution
is asymptotically stable, it will approach the critical point as t> (see Figure
SEC. 7.6 / NONLINEAR SYSTEMS AND STABILITY 263
7.5). If the solution is unstable, no matter how close to the critical point the
particle starts, there are solutions that move away from it.
Trajectory
(a, B)
[8 Yo)
3. If a trajectory crosses itself at a point of the phase plane that is not a critical
point, that trajectory is a closed path and corresponds to a periodic solution
of the system (see Figure 7.6).
hes a critical
These statements suggest that a particle (solution) either approac
hes a closed path, or goes off
point as t>*, moves along a closed path, approac
of critical points is most important
to infinity for increasing time. Thus the study
in the analysis of autonomous systems.
n set of the system
In order to test for stability by using Definition 7.2, the solutio
do not have a known solution set,
must be known. Since nonlinear systems often
y without first explici tly finding the
it is desirable to be able to test for stabilit
let us begin by examin ing more
solution of the system. To see how this is done,
extend our ideas to similar nonlinear
closely linear autonomous systems and then
systems.
x'
ff
= ax + by, aa
y =cx
t+dy,
Ul
where a, b, c, and d, are constants such that ad — bc # 0. This system has one
critical point at the origin of the phase plane.
x’ =ax
+ by, ad—bc#0
{?
y =cx + dy,
is stable if and only if both roots of the auxiliary equation
A?
— (a + dd + (ad — bc) = 0
have nonpositive real parts. The critical point has asymptotic stability if
both roots
have negative real parts.
Some of the possible trajectories for the linear system (78) are shown
in Figures
7.7 through 7.11. The critical point (0, 0) is referred to as a node
in Figures 77
and 7.8, a focus in Figure 7.9, a center in Figure 7.10, and a saddle
point in Figure
7.11. These figures make clear that small changes in the initial
conditions produce
only small changes in the solution for stable systems but can lead
to large changes
in the solution for unstable systems.
SEC. 7.6 / NONLINEAR SYSTEMS AND STABILITY 265
y
A A
: Xx |
y y
rN r
: by:
oy J
A
A
‘ © ‘
><
(stable)
yay.
Solution The auxiliary equation is A* + 2 = 0 with roots A= + V2i. Thus the origin is
a center and the solution is stable.
Solution
The auxiliary equation is A* + 6A + 9 = 0 with double root A = —3. Hence the
origin is a node, which is asymptotically stable.
CUO ae
SEC. 7.6 / NONLINEAR SYSTEMS AND STABILITY 267
Our nonlinear system (80) now has the form of a perturbed linear system (nearly
linear system)
y
t/
Ch GV AUN y). (81)
In most cases, stability behavior of the perturbed linear system in the neigh-
borhood of a critical point (0, 0) is closely related to the stability behavior of the
corresponding linear system arising when p(x, y) = q(x, y) = 0. In particular, we
have the following theorem, which we state without proof.
Theorem 7.7 Let (0, 0) be a critical point of the perturbed linear system
,
X= aXe DY +,DUX..y),
Uy
y ay rag):
x = ax by dd = be 0,
y' =cx + dy,
Even though the stability of a perturbed nonlinear system may be the same as
that of the associated linear system, the trajectories of the nonlinear system may
differ greatly from those of the linear system. Also, if we move sufficiently far from
a stable critical point of the nonlinear system, it may no longer be a point of stability
of the system. This is in sharp contrast with linear systems, where stability is not
localized.
EEE
EXAMPLE 26 Discuss the stability of the simple pendulum whose equation of motion is
6” + k? sin@ = 0.
x'=y,
y' = —k?*sinx.
'
SEC. 7.6 / NONLINEAR SYSTEMS AND STABILITY 269
This system has an infinite number of critical points (0, 0) and (nz, 0), where
n= +1, +2,.... The associated linear system near the critical point (0, 0) is
(sinx =x as x0)
'
x M5
eed coe
which
Here we find that the roots of the auxiliary equation are pure imaginary,
7.7, however,
means that (0, 0) is a center of the linear system. Based on Theorem
(0, 0).
we cannot conclude anything about the stability of the nonlinear system near
about the
At the critical point (77, 0), the situation is different. Expanding sinx
point x = 77 yields
i
Sue Tee) Gay Tek) ote
y =k yo):
— k* = Oare
Letting z = x — 77, we see that the roots of the auxiliary equation A?
system.
A, = kand A, = —k, so that (77, 0) is an unstable critical point of the linear
ar system.
Therefore, it is an unstable critical point of the nonline
At the point x = 77, the pendulum is pointing vertica lly upward. Clearly, such
cause large motions of the
a position is unstable, since a small displacement will
pendulum hanging vertically
system. Also, the point x = 0 corresponds to the
downward. It is equally clear that such a positio n is stable based on physical
considerations in spite of the fact that we could not derive this conclusion from
points of the form (n7r, 0) will
Theorem 7.7. In fact, it is evident that all critical
of n and of the type (77, 0) for odd
be of the same type as (0, 0) for even values
stable (but not asymptotically
values of n. Thus the critical points are alternately
stable) centers and unstable saddle points.
7.12. Close to the stable
The trajectories of this system are shown in Figure
critical points are the closed paths suggesting periodic motion. The trajectories that
cross at the unstable critical points are called separatrices. Finally, the wavy paths
outside the separatrices correspond to whirling motions of the pendulum.
For purposes of contrast with Example 26, it may be interesting to consider the
effect of frictional forces acting on the pendulum. Assuming such a force is
proportional to the angular velocity 6’, the equation of motion becomes
6" + co’ +ksind =0, c>O0, (82)
or, equivalently,
/
Am kame) 5
Va=
,
ke cy, (84)
with auxiliary equation A? + cA + k? = 0. The roots are
Or ti(C = a4)
di, Az » 5)
and hence the origin is a stable node if c > 2k, or a stable focus if c < 2k.
According to Theorem 7.7, in all cases the origin of the nonlinear system is also
asymptotically stable.
A more direct method than presented here is also available for analyzing the
stability of nonlinear systems. This method, which rests upon the construction of
a suitable auxiliary function, is a more powerful method in that it provides more
global type of information. This method is a generalization of the physical prin-
ciples associated with a conservative system and is due to A. M. Liapunov
(1857-1918). The Liapunov theory, however, would take us too far afield for our
purposes, so we refer the interested reader to the references.
EXERCISES 7.6
In problems 1-4, sketch the trajectory corresponding to the solution Satisfying the
prescribed initial conditions, and indicate the positive direction of motion.
Less Oe Ss PEE ae Sa
y' =x, y(0)=0 y' =x, yO) =0
3.nitesty,~ Tx(0)) Sted 4. x'=2x+ 4y, x0) =4
y= Xe). 0 y' =—2x
+ by, 3(0).=0
SEC. 7.6 / NONLINEAR SYSTEMS AND STABILITY 271
5. x'=y Gi x=
Ye ax yea xy
Lk = Soy Bat, Skee
y’ =x + 3y yi pax 1 2y ite!
oO. x=A xy 10, x0 =e — x7 xy
11. x’ =y 12a al a 2y
pix | Vo teDyan)
13.7% = 3x 14, x’ = —2x’y
y' =x + 3y y’ =x*+y*
15. 2% =2 —\y* 16. — siny
x' =2xy
yi ny y'=1-y?
TJ = 3x yx ty
y' = y(3 — 2xy’)
18. Consider the linear autonomous system
yo S54
t
yo =x ey-
conditions x(0) = 1,
(a) Find the solution of this system that satisfies the initial
y(0) = 3.
(b) Repeat (a) for x(4) = e, y(4) = 4e.
both represent the same
(c) Show that the two different solutions in (a) and (b)
trajectory.
ly 20. + 4y
x' =2x
19 sex = 2h
BY y pr 2x + Oy
yl = xr
+ 3y 222k aay
21. x’ =x
y'=3x ty Vespa = ge)
ey 24 ety
23 eee =
y' =x +5y
y' = 4x + 5y
25x = xray
to"
In problems 27-32, determine the nature of the critical point (0, 0) by analyzing
the related linear system.
Determine the critical points in problems 33-35, and discuss their nature and
stability.
X30. DN
'
Na Se wsban
Lf
(a) Show that the change of variables x = cX/d, y = aY/b leads to the system
X' = a(X — XY),
YaA= =) aX):
Birkhoff, G., and Rota, G.-C. 1978. Ordinary differential equations. 3rd ed. New
York: Wiley.
Boyce, W. E. and DiPrima, R. C. 1977. Elementary differential equations and bound-
ary value problems. 3rd ed. New York: Wiley.
Cullen, C. G. 1979. Linear algebra and differential equations. Boston, Mass.: Prindle,
Weber & Schmidt.
Finney, R. L., and Ostberg, D. R. 1976. Elementary differential equations with linear
algebra. Reading, Mass.: Addison-Wesley.
Ross. S. L. 1980. Introduction to ordinary differential equations. 3rd ed. New York:
Wiley.
273
Numerical methods are becoming more and more important in applications, partly
because of the difficulties encountered in obtaining exact analytical solutions but
also, more recently, because of the ease with which numerical techniques can be
used in conjunction with today’s high-speed automatic computers.
Several numerical procedures for solving initial value problems involving
first-order DEs are discussed briefly in Section 8.2, all of which are based upon
Taylor series approximations. They include Euler’s method, the improved Euler's
method, and the Runge-Kutta method. Of these, the Runge-Kutta method is the most
widely used because it is far more accurate.
These same numerical techniques are generalized in Section 8.3 to include
systems of first-order equations and higher-order equations, which are solved by
reducing them to a system of first-order DEs.
Although error analysis is an important part of any numerical procedure, we have
limited our discussion primarily to the use of the procedure itself. The theory of errors
is sometimes fairly complex and goes beyond the intended scope of this chapter.
The interested reader should consult a text on numerical analysis.
274
6
A
are,
8.1 INTRODUCTION
In applications we must often solve an initial value or boundary value problem that
is either difficult or impossible to solve exactly by analytical methods. For this
reason it becomes either convenient or necessary to employ some method that
yields accurate numerical estimates of the true behavior of the system. However,
the method itself can involve a considerable amount of analysis concerning the
errors involved in using the method as well as errors due to rounding off in the
computations. We do not intend to discuss these matters deeply, as is done in
courses on numerical analysis; instead, we will simply present some techniques that
provide quantitative information about certain systems we wish to study.
where the function F is suitably “well behaved.” We will assume the problem has
a unique solution in some interval containing the point fo (see Theorem 2.1).
The methods to be discussed here are step-by-step procedures, wherein each
calculated value makes use of the previously calculated value. We start at the initial
point on the solution curve (fo, Yo) and increase t) by the fixed (positive) number
h to get t; = tf + h, and then compute a value y, that approximates the solution
value y(t,). In the second step of the procedure, we find an approximate value y2
for y(t), where tf = t; + h = f + 2h, and continue in this fashion.
The calculations are all done by the same formula at each step of the process.
These formulas are suggested by the Taylor series
—
d
eee i aL
dy
pee
”
with respect to
and so forth. The subscript variables indicate partial differentiation
e, we set ¢ = % and
the designated variable. In the first step of the procedur
y(t.) = yo to calculate
fer
yo) cor at ee)
Vite 0 a hF (to, Yo) “12 5h As
275
hi
276 CHAP. 8 / NUMERICAL METHODS
whereas, in general,
For computational purposes, the series in (3) must be truncated after a certain
number of terms. This leads to an error, which is appropriately referred to as a
truncation error. If h is picked to be sufficiently small, terms involving h’, h°, and
higher powers of h can often be neglected in (3), making a first-order approxi-
mation. The truncation error per step is then of the order h*. When higher-order
terms of (3) are required for greater accuracy, we must compute the derivatives of
F. Since these derivatives must be done by hand, using (3) directly is often not
feasible when F is of a complicated nature. In such instances it is usually preferable
to replace the derivatives by certain numerical equivalences so that the formulas
lend themselves to computer calculations. That is precisely the approach used in the
improved Euler method and Runge-Kutta methods to be discussed.
obtained from (3) by truncating all terms involving powers of h higher than 1.
Geometrically, this technique approximates the true solution curve with a polygon
whose first side is tangent to the solution curve at (f, yo) (see Figure 8.1). Hence
it is also called the method of tangent lines.
S
True solution
(to, Yo)
Figure 8.1
oe
eee
EXAMPLE 1 Use Euler’s method to approximate the solution of the initial value problem
n=0: y= Yo — 2(0.1)toy
1 — 2(0.1)(0)(1)?
= |,
= 0.98,
Tiel f2 y3 = y2 — 2(0.1)tay3
kes = 2(0.1)(0.2)(0.98)
= (0).
= (0)Sil).
1.0000
1.0000
0.9800
0.9416
0.8884
0.8253
1.0000
1.0000
0.9950
0.9851
0.9705
0.9517
0.9291
0.9032
0.8746
0.8440
0.8119
278 CHAP. 8 / NUMERICAL METHODS
Since we had the exact solution in Example 1, we could calculate the amount
of error incurred in the method at each step of the procedure. For instance, over the
interval in which we made the calculations, the maximum percentage error in the
first case is found to be
|error|
elles socal ee _ 0.0263
= ye = 3.05%,
exact value oY 0.8621 Bt : 4
O0125
0.862] We
x 100 ceil 1.45% !
We might conclude that the Euler method works quite well, since the errors are
acceptable in many applications. However, this kind of accuracy is usually not
realized by Euler’s method in practice (see Example 2).
a ee Ped ee hee, ee
EXAMPLE 2. Use Euler’s method to solve the initial value problem
Vasa
last titty my O)eals
on the interval O S¢ <0.5 withh = 0.1.
Table 8.3 Euler’s method fory’ = 1—1+ 4y, y(0) = 1, withh =0.1
Exact
Value
SEC. 8.2 / NUMERICAL METHODS FOR FIRST-ORDER EQUATIONS 279
oe.1) Gt 016)
Beh.
Additional values are provided in Table 8.3 along with the exact values for com-
parison.
Euler’s method can be made more accurate for a fixed value of h by first computing
the auxiliary value
Vie =I Yn ats hF (th, Yn) (5)
This technique, called the improved Euler method (or Heun’s method), is an
step we
example of what is called a predictor-corrector method. That is, at each
predict a value by (5) and then correct it by (6).
true
The geometric interpretation of this new method is that we approximate the
(tf, y,) with
solution y in the interval from f, tot, + sh by the straight line through
slope F(t,+1, Ye+1) UP to tr+1 (see
slope F(t,, Yn), and then along a new line with
sum Fh, Va) + Fass y*.1)] as
Figure 8.2). Therefore, we might interpret the
y True solution
280 CHAP. 8 / NUMERICAL METHODS
some average slope over the interval t, < t < t,+4,. It can be shown that (6) is
equivalent to a Taylor series through the term containing h?, but it has the advantage
that dF/dt does not have to be calculated.
ee
ee eee
EXAMPLE 3_ Use the improved Euler formula to obtain an approximate solution of
Vee ye ty(Oars
Solution We will illustrate the technique for h = 0.1 on the interval 0 < t < 0.5. We first
calculate
na 0; yi = yo — 2htoys
= 1 — 2(0.1)(0)(1)’
— ¢
and then
1
y= Yo, 5 hl —2toyo eI 2n(yi)"]
1 — (0.)[)C)* + O.1)d)]
EO:
The remaining calculations are provided in Table 8.4 along with those obtained by
Euler’s method and the exact values.
Improved
Euler
Example 3 shows that the results using the improved Euler method with
h = 0.1
are better than those obtained with the Euler method even for h = 0.05. In
general
this is the case, and although a few more calculations are required at each
step using
the improved Euler’s method, the greater accuracy is worth the extra effort.
SEC. 8.2 / NUMERICAL METHODS FOR FIRST-ORDER EQUATIONS 281
Perhaps the most commonly used as well as most accurate technique is the Runge-
Kutta method .* At each step we must first compute four auxiliary quantities
ky = F(t, Vas
1 1
k, = F(t ar ah, yer sh),
: (7)
= F(t + ah, Vice ss),
kg = Filta Yn XG hk3),
y of a Taylor
The purpose of the Runge-Kutta method is to achieve the accurac
to calculat e higher- order derivati ves. For example,
series expansion without having
(8) was derived by obtaini ng appropr iate constants
the algorithm given in (7) and
A, B, C, and D so that
Ynt1 = Yn + Ak, + Bky + Ck; + Dk,
term of the series (see
agrees with the Taylor expansion out to h4, or the fifth
-order Runge-Kutta
problem 26). Thus the technique is often called the fourth
+ k4)/6 appear ing in (8) can be interpreted as
method. The term (k; + 2k, + 2k3
an average slope over the interval t, = t S t+1-
our previous formulas,
Although the Runge-Kutta formula is more complex than
even the improved Euler
it yields results that are many times more accurate than
es this accura cy with larger comparative in-
formula. And it generally achiev
computer, the calculations
crements of h. Furthermore, with the use of a modern
in a few second s for most problems.
are routine and can be performed
on F does not explici tly depend on y, then
Finally, we note that if the functi
1
ky = F(t), ke = k3 = F(t +5h), ky = F(t, + h),
ky = ive
02
y=1- “6 (2(0.2) + 2(0.192) + 0.37] = 0.9615,
The remaining calculations are given in Table 8.5 along with similar results ob-
tained by the Euler methods. The improved Euler method with h = 0.1 is also
included for comparison.
k, = F(0,1) = 5,
ky = F(O + 0.05, 1 + 0.25) = 5.95,
ke = (040.05, 140.2975) =.6.14,
k, = F(0.1, 1 + 0.614) = 7.356.
Thus,
| 0.1
me =1+ a. + 2(5.95) + 2(6.14) + 7.356]
1.6089.
A tabulation of the remaining values is given in Table 8.6 as well as the results
obtained from the Euler methods. The superiority of the Runge-Kutta method is
clearly demonstrated by this example.
Improved Runge-
Euler Kutta Exact
Values.
requires
Remark. The Runge-Kutta method has certain drawbacks in that it
calculations of the function F(t, y) at successive steps of the
time-consuming
expensive. These calcu-
procedure. In some applications this technique may be too
rrector methods, such
lations can be greatly reduced by use of certain predictor-co
However, these methods create some inconve-
as the Adams-Moulton method.
change in step size as the
niences of their own in many cases by requiring a
choose will greatly depend
calculations proceed. The proper numerical method to
the necessary computer
upon the application and the budget allowed for making
calculations.
284 CHAP. 8 / NUMERICAL METHODS
EXERCISES 8.2
In problems 1-10, use the Euler formula to obtain an approximation to four places
after the decimal, to the indicated value of y with h = 0.1.
1. y’ =2y, yA) =1, ys) =7
y =1 + y yO) =U, yOs) = 7
ye tery Sl)s yO) = 20 yOs) a4
y ‘= pty, yO)
= 1) [email protected]) =?
y '=??+y’, yO)=1, yO.5)=?
y '=ty+Vy, yO)=1, y(0.5)=?
11. Repeat the calculations in problems 1-10 using the improved Euler formula.
12. Repeat the calculations in problems 1-10 using the Runge-Kutta method.
13. Using Euler’s formula with h = 0.2, find an approximate value for y(1) where y is
the particular solution of the initial value problem
y 2 yw O= 1
(Your answer approximates the value of e. Can you explain why?)
14. Solve problem 13 using the improved Euler’s formula.
15. Solve problem 13 using the Runge-Kutta method.
16. Using Euler’s formula with h = 0.2, find an approximate value for y(1) where y is
the particular solution of the initial value problem
y' =(? +1)", yO) =0.
Use your answer to approximate the value of 77.
Hint: The exact solution is y = Arctan (tf).
17. Solve problem 16 using the improved Euler’s formula.
18. Solve problem 16 using the Runge-Kutta method.
“19. Derive Euler’s formula (4) by integrating both sides of the DE y’ = F(t, y) from x,
to x,+, and then approximating the integral by
[feoax = (b — a)f(a).
Give some justification for using the above approximation for the integral.
20. Using the step size h = 0.2, approximate the value y(1.4), where y is a particular
solution of the initial value problem
y=r?ty’, y)=1,
SEC. 8.3 / SYSTEMS OF EQUATIONS 285
The methods discussed in the preceding section for solving initial value problems
featuring first-order DEs can be extended to a system of first-order equations. For
instance, let us consider a system of two equations
x' = F(t, x, y), x(t) = 0,
(9)
Ve ae G(t, x, y), y (to) = die
As before, we will assume that F and G are suitably behaved so that (9) has a unique
solution (for linear systems, see Theorem 7.1).
For illustration purposes we will generalize the Euler method, since it is the
simplest to apply. Thus, we first calculate
X, = Xp + hF (to, Xo, Yo);
(10)
x, oi See DE Gn)
where ¢, = t + nh.
ee ee ee ee eee ee Se
1+ 0.Dd -—0+ 1)
= 1.2
and
V1 = yo + h(2x% + 3yo + e ”)
= 0 + (0.1)[2(1) + 3(0) + 1]
= 0.3.
Similarly,
ApS Ae es!)
= 172) (O81) Ce 2e— Ona en)
= 1.4005
and
Vi VA ap h(2x, ap 3y, ar e ")
= 0.7205.
ft) =
1
—
1
x(t) 10° 2t Qi cost =— 13isin'
. ey
tire te a
10° .
l
y(t) = go (A cost +17 sing) eo— ae"
4 =7);, X(0) = 2)
, (12)
PS Ns) 2 (0) = yo,
and thus can be solved by methods already discussed.*
x'=y, x(0)=1,
y’ =t—3x-—t’y, y(0) =2.
From (10) we have
xX; = X + hyo
= 1+ (0.1)(2)
=al22.
yi = yo + (to — 3x0 — fYo)
So (OUD SO o ew)
I,
and
*Some numerical analysts believe that greater accuracy is achieved by applying a numerical procedure
directly to the higher-order DE rather than reducing the DE to a system of first-order equations and then
applying a numerical procedure. See Peter Henrici, Discrete Variable Methods in Ordinary Differential
Equations (New York: Wiley, 1962).
288 CHAP. 8 / NUMERICAL METHODS
D9) a 6) ar hy,
= 1.2 + (0.1)(.7)
= 1.37,
yo = yi + A(t, — 3m — fi)
= 167-42 (O5L) (0: Deas Gle2 (0.1)?(1.7)
ll 1.3483.
EXERCISES 8.3
In problems 1-5, use the Euler method with h = 0.1 to determine approximate
values of the solution at t = 0.1 and t = 0.2.
where
K, = I Cg Xny Male
K>
oe
— F(tn
1
qe 5h, Xn
A
ate 7 hKi, Yn a
u5 Li),
eet
ce ik y Ne, n 5) 9 An D 29 ay,osSing
Yn D 2)>
SEC. 8.3 / SYSTEMS OF EQUATIONS 289
Change the initial value problems 11 and 12 to a system offirst-order DEs, and use
the Euler method with h = 0.1 to determine approximate values of the exact
solution at t = 0.1 and t = 0.2.
REFERENCES
Acton, F. S. 1970. Numerical methods that work. New York: Harper & Row.
Burden, R. L., Faires, J. D., and Reynolds, A. C. 1978. Numerical analysis. Boston,
Mass.: Prindle, Weber & Schmidt.
series, and for that reason the procedure is referred to as the power series method.
Because of computers, this method is no longer as useful as it once was. Never-
theless, the theory is still very important since it can be used to determine regions
where the solutions are analytic, and this information is essential even in the applica-
tion of numerical techniques.
The general method is discussed in Section 9.2 for the case of power series
expansions about ordinary points. For this case we always find two linearly indepen- nha
290
0
9.1 INTRODUCTION
where Co, C1, -- - » Cn» - - - are called the coefficients of the series and xo is the
center of the series. The series has the sum co when x = Xo, but generally we are
interested in whether the series also has a sum for other values of x.
Definition 9.1 A power series == cn (x — Xo)" is said to converge for a particular value of x if
N
limDen(x — Xo)"
2
exists. Otherwise the series is said to diverge.
The values of x for which the series (1) converges is called the interval of
convergence. That is, to each series corresponds a number R, called the radius of
convergence, with the property that the series converges if |x — x9| <R and
diverges if |x — x| > R. The radius of convergence of many power series can be
found by means of the ratio test.
Ratio Test If
Coico) a |= |x — xo|lim
Cnt+l
L = lim )
roa
Cyl FxG)" now
Cn
> 1.
then the series > Ca(X — Xo)" converges when L < | and diverges when L
t= Sx",
MiaBh alent
s1<x <1, 3)
foe] x"
e Sri DN
2 met
nt 1GOOl. OX
x Pome, (4)
(oe) antl (5)
in
sinx 2 ((eee
1) On # DI’ Oran (
xn
cosx =ae ere yp OO xy <0, (6)
Many of the series of interest have centers at x = 0 as in (3) through (6), but
not all power series do; for example,
co Hy!
log(1 + X) = Deora
— 2 paki Xe. (8)
For this reason much of the pe esetin subsequent sections will be confined to
power series for which x9 = O without loss of generality.
y= » Cue (9)
n=0
It follows that power series for y’, y”, . . . can be obtained from (9) by termwise
differentiation. That is,
The method now is somewhat like the method of undetermined coefficients in that
all that remains is to determine the constants c, appearing in (9) in order to have
the solution.
Let us illustrate the basic procedure by applying it to the simple equation
foo} foo}
To combine these two series, we must have the same exponent for x in each
corresponding term, and both summation indices must start at the same value. Let
us begin by replacing the index n by (n — 1) in the second sum to get
foe] oo
OMe ag eas:
ICH acral (12)
n=1
Now, since x can take on various values, the coefficient of x”! must vanish in order
to have (12) identically zero; 1.e.,
NC Cp ON Na yo,ees (13)
while cy can remain arbitrary since the term in which it appears is already zero. The
relationship (13) is called a recurrence formula for the unknown coefficients. Since
n # O, we can rewrite (13) as
Cp =e Na Ooi
2)e (14)
n
Cee CO,
C2
Peli
he. 5 0
Lae ac oape 60
Eas 2003 amet
a ee COME CO
BON TSE TTD ate
tle Mee etier eto, wiememie§ es ene) leuce,
2 3 4
Xx Xx
Stet tion ait gyt 1a
294 CHAP. 9 / THE POWER SERIES METHOD
or
00 n
x
fy a= co» Hid
n=0
In this case we recognize the infinite series as that of e* and can therefore write the
general solution as
y = coe’, (15)
in agreement with our previous result.
Although the above example was quite elementary, it illustrates the basic manip-
ulations required to solve a DE by the power series method. Even when the
coefficients of the DE are polynomials in x rather than constants, the calculations
required differ little from those used in this simple example.
The power series method does not provide a general solution to all variable-
coefficient DEs, even for the special case when the coefficients are polynomials in
x. Before discussing the method any further, however, let us first clarify the two
rules of manipulation that were used in the above example and will be used in all
future examples.
Rule 1: When making an index change, always make all exponents on x equal
to the smallest one occurring in the various series.
Rule 2: When combining series under one summation sign, start all series with
the largest of all the beginning values. Terms preceding the new first value of the
summations must then be added outside the summation sign.
ee
ee ee ee ey Oe ee ee
Solution In (a), we set 1 — x* = Oand find x = +1 as the singular points. All other points,
real and complex, are ordinary points.
The only singular point in (b) is x = 0, and thus all others are ordinary points.
The DE in (c) has singular points at the solutions of x? + 4 = 0, or x = +2i.
Again, all other points are ordinary.
More generally, if the coefficients in (16) are not polynomials, we first rewrite
the equation in normal form
EXAMPLE 2 Determine whether the DE 2xy” + (sinx)y = 0 has any singular points.
we see that all points are ordinary points, including x = 0, since (sin x)/2x has the
power series expansion
pine
2x Zz
etn asthe
St a |
which converges for all x and therefore must be an analytic function.
BA EE ——————E——————e—e———— en
In the remaining sections of this chapter, we will confine our discussion to only
those DEs that have polynomial coefficients. Furthermore, we will present the
theory only for second-order DEs and make no attempt at generalizing to higher-
order equations.
296 CHAP. 9 / THE POWER SERIES METHOD
For most problems we are mainly concerned with whether the point x = 0 is an
ordinary point or a singular point of the DE, since this point is the easiest point at
which to apply the power series method. If x = 0 is an ordinary point the following
theorem is applicable. We state it without proof.
Theorem 9.1
n=0
Moreover, these series will converge at least in the interval |x| < R, where R is the
distance from 0 to the nearest singular point.
Remark. If the nearest singular point is complex, then we define the modulus of
this point as the distance from the origin. For example, the modulus of 1 — 2i is
[1 — 2i] = V5.
Theorem 9.1 is easily generalized to any ordinary point x = Xo. The series
solution in such a case will then be of the form
Y= Gea ook
n=0
te S|Nip e =
n=2
or
+ 2c,-2|x""? = 0.
Clearly, co and c, are arbitrary, but the remaining constants must satisfy
n(n — 1)c, — ([(n — 2) — 3) + 2(n — 2) — 2]e,-2 = 0.
C, = (AF) ens Fe a ne ie
Oy = Wy
C4
ye
3° 3 60
Z
Coa— ie = 0,
C6
ae
a
ee5 Or
4
C7 6a: = 0,
> 1
Oy = mice aw
Thus we get
y = Cy + C1x + gx? + C3x? + cax* + Cx
5
+--
CHAP. 9 / THE POWER SERIES METHOD
Co 4 Co 5 Co 7
=c t+ cx — Ox? — —xt- =x x
Bue : 3 5 7
y = coyi(x) + cy y2(x).
If the pattern in each series is clear, it is sometimes useful to write the solutions
in summation notation. For instance, in the above example we can write
xn
whereas y2(x) = x is already in this form. The ratio test can readily be applied to
solutions written in this form to determine the radius of convergence. In the present
example it is easy to verify that the series converges for |x| < 1.
Remark. In some cases the interval of convergence extends beyond the closest
singularity.
or
We now see that co and c, are arbitrary, and the recurrence formula is
nA — 1) Ce hs (1 —52) Cen, Ge Os
or, upon simplification,
SEC. 9.2 / THE GENERAL METHOD 299
i2 7 60
joe
3 30
] C
pa
5 5
heBLE
3 2: ’
ee
: 6x 2-4-6’
eee
c7 = ae eae!
aR IEG
y = Coy(x) + cyy2(x).
Although it is not obvious, the first series can be written in terms of an elemen-
tary function. That is,
~-L4+—-
3g
+
LD) sia) Sa Sire aPC
N Ww
ll | | ab | te
NI Sake epi NI
+
Se bedi
fe
—
aS Se ae
a
or
series.
This is not the case for the second series, so it must be left as an infinite
Hence we write
erste
3
X
Bex agraPahl nx 5) 7
+e).
ae + a(x Be Caer
300 CHAP. 9 / THE POWER SERIES METHOD
Imposing the first initial condition y(0) = 3 on this general solution, we imme-
diately find that
Co 5.
In order to impose the second initial condition y’(0) = —7, we first calculate
4 6
7) x Xx
y’= —coxe r+ a1 7 fot fees ,
If the initial values of y and y’ in Example 4 were prescribed at some point other
than x = 0, then the solution technique would be altered. For example, suppose we
wish to solve the initial value problem
ig Os eee a (19)
n=0
Venki) y 4 1) yf
about the point x = 1.
Assuming
we find that
SG DG et Ancacae acA = 0,
n=0 =0 n=0
nn 3 in 3
Be ye 2a
Once more Cp and c, are arbitrary, but we must set cz = 0. The remaining constants
are then determined from the recurrence formula
n(n — 1)c, + (n — T)ep-3 = 0
or
lumen) =} oe) ee
CcPoe nh 3 1) Gaeasmlt
n—39
ye RS) eae -
C3 3.2 Co 3-2 0>
SAAC, ON
Oe MUU nae g
—2)
Ck = 54 c, = 0,7
DE (ean RIT A
Coin ae 6154 ee 69523220 4
Gia ’
C3 = 0,
ee 24
OOS eg O 78 9655:6 0.
3
= — = 0
C10 10:9.” >
Cie 0,
302 CHAP. 9 / THE POWER SERIES METHOD
Because the original DE has no singular point in the finite plane, this solution is
valid for all finite values of x.
In some cases the recurrence formula involves more than two terms in the
unknown coefficients. Consider the next example.
Solution Making the appropriate substitutions for y, y’, and y”, we find after simplification
foc) io) ie)
Co
Oana?
Cheats
¢, = oo an =3,4,5,....
nie)
To simplify matters here and still obtain two linearly independent solutions, let
us first set c; = O and let co remain arbitrary. Therefore
1
Se A Sei
2
SEC. 9.2 / THE GENERAL METHOD 303
A
: A aera
Cc BRR C2 SRC ie WO
4-3 4!’
C3 + C2 4co
Cia => —
eC ee Cee eee Oe a ne
c, = 0,
pet aa
: 3-2 B17
pee Sone ed
4 4-3 At;
EXERCISES 9.2
o~
&
oo
ts ee) ee spa
ere as a series in x”.
\A. > Dues 2
n=2 n
ES (Hece | ee
2. y a ea as a series in x”.
n=0
n
41 Le as a series in x”*'.
> @yay-- Gay
n=1
In problems 5-8, list all of the singular points in the finite plane.
‘5.
\
In problems 9-12, find a power series solution about x = 0 for the given first-order
DE.
at 3)
y" Hay = 05, 14, y"—4y =0
45) at 4x2)y" — 8y = 0 16.5 yy" Felxy = 5y = 0
17. (1 — x2)y" — 2xy' + 12y =0 18, (x? + 4)y" + 2xy — 12y = 9
19. y" + xy’ + (x? + 2y =0 202, Hat ixry tray = 0
21. «@-—Dy"’+y'
=0 22. @ +3)y"
+ Ge 2)yas y= 0
In the initial value problems 23-26, find the first four nonzero terms in the power
series expansion about x = 0 of the solution.
we
In problems 31 and 32, find the power series expansion about x = 0 of the general
solution.
*31. ve cox 3xy' =’ y = 0 339% ye ate Noy aL Sx! ae 3y = 0
"33. The DE
Voy, reeny 10,8 yu;
is called Hermite ’s equation .* Obtain two linearly independent solutions for the cases
when
(a) n=l.
(b) n= 4.
(c) Show that when n is any nonnegative integer, one solution of Hermite’s DE is
always a polynomial of degree n.
*34. The DE
(1 — x*)y"— xy’ + n’y = 0, n 20,
is called Chebyshév’s equation .+ Obtain two linearly independent solutions for the
cases when
(a) ore — al
(b) n=4.
(c) Show that when n is any nonnegative integer, one solution of Chebyshév’s DE
is always a polynomial of degree n.
some
The solution of a DE in the neighborhood of a singular point usually exhibits
type of peculiar behavior, and for this reason the behavior of the physical system
singular
governed by such an equation is frequently most interesting around this
technique ,
point. Therefore, rather than avoiding the singular points in our solution
we need to investigate precisely these points in many situations.
be possible
Unfortunately, when x = 0 is a singular point of the DE, it may not
to find a power series solution of the form
Under proper conditions, however, we may be able to find a solution of the more
general form
Remark. In the special case when a(x) and b(x) are rational functions reduced to
lowest terms, an R.S.P. is a singular point for which the factor (x — x9) appears
at most to the first power in the denominator of a(x) and at most to the second power
in the denominator of b(x).
Solution Clearly, x = 0 is the only singular point, and dividing the DE by x shows that
|
a(x) = 2 and b(x) = 1.
Thus, since both xa(x) = 1 and x*b(x) = x’ are analytic atx = 0, we conclude that
x = Ois anR.S.P.
Solution The singular points are x = 0, 1, —3. Putting the DE in normal form identifies
oer
EET RG LLY
SEC. 9.3 / SOLUTIONS NEAR A REGULAR SINGULAR POINT 307
and
2x2 th)
eax 423)"
Therefore, since x and (x + 3) appear at most to the first power in the denominators
of a(x) and b(x), it is clear that x = 0 and x = —3 are both R.S.P.’s. Likewise,
since the factor (x — 1) appears to the second power in the denominator of a(x),
x = 11s anI.S.P.
we find
and
y"=SYatsnts— Dox eS
n=0
Assuming x = 0 is anR.S.P., both xa(x) and x7b(x) have power series expansions;
1263)
xa(x) = do + ax + @x?+---,
DG) = boa Dyan Dyk ae
If we substitute these expressions into (23), the smallest exponent of x occurring
in (23) is (s — 2), corresponding to n = 0. Let us separate this term from the rest
and write
This important quadratic equation is called the indicial equation of the DE (21).
There are two roots of the indicial equation, s, and s;, leading to three different
procedures for generating a second linearly independent solution, depending on the
nature of the roots:
SEC. 9.3 / SOLUTIONS NEAR A REGULAR SINGULAR POINT 309
We will consider the three cases separately under the assumption that s; and s, are
real solutions of the indicial equation (24). Although these roots can be complex,
we will not discuss this case.
This case always leads to two linearly independent solutions of the form
and
constant for
These solutions are clearly linearly independent, since y)/y2 cannot be
are obtained
our choice of s, and s). The two sets of coefficients c, (s,;) and c,,(s)
recurrence
independently by replacing s by s; and s by 52, respectively, in the
formula.
Ee Le ee ae
ee
EXAMPLE 9 Solve 2xy” + (1 — 2ny’ — y = 0.
the infinite series
Solution We first observe that x = 0 is an R.S.P. Next, substituting
expressions for y, y’, and y”, we find
n—>n-1 n—>n-1
2s? — 5 = 0*
with roots s; = 0 and 5) = 5. The recurrence formula is
(n + s)[2(n +s) — Je, = [2(n + 5) — 1en-1,
or, since 2(n + s) — 1 # 0 for either choice of s,
(a + se, = Een SH 27 ee
ey = els
Cy = co = 1,
St
2 5 J?
Cae fl
2 mipane:
5 eecee ak
ult 4e eate
aenee
| 3 Be
e GMs
Daeees a5
é ee 2:
AES PDE TL
and thus
9x 2x? Dee
yo(x) = x7( 1 + — +
Ne ( 3 S. 3u5et
The general solution is therefore
*Observe that we could find the indicial equation directly from (24) by noting that xa(x) = 3 — x and
x?b(x) = —x/2, and hence ap = 4 and by = 0.
SEC. 9.3 / SOLUTIONS NEAR A REGULAR SINGULAR POINT 311
y = Ayi(x) + By2(x)
Dane S400
= Asspsl tenet ec ee eal}
where A and B are arbitrary constants.
When s, = 52, the procedure used in the last case will yield only one solution.
Nonetheless, two linearly independent solutions can be shown to exist in this case
corresponding to
Pe ae S=S|
where -
Remark. By (29), we simply mean the general function obtained for arbitrary s,
which becomes y; upon setting s = 5).
pe oes ee all
312 CHAP. 9 / THE POWER SERIES METHOD
Setting the coefficient of co to zero, we see that the indicial equation reduces to
s? = 0 with double root s,; = s, = O. For this choice of s, the coefficient of c; does
not vanish and so we are forced to set c; = 0. For values of n greater than 1, we
have the recurrence formula
(n? + S)'Ce 4 Cee 0, Bh = 2, 3, Al ea
or
In this last form of the recurrence formula we are emphasizing the dependence of
these constants on the parameter s. Since the solution formula (28) requires deriv-
atives of the c’s with respect to s, we will not substitute the value s = O into the
recurrence formula until later. From the recurrence formula, there follows (with
Co = | once again for convenience)
C(s) = ee Oem 2 ee
Y GE) Gey *
c3(S)
teeee
(s aE 3) 0,
c4(s) = eel ea ee
(s+ 4)? (s + 2)°Xs + 4)’
c5(S) = 0,
Cc (s) = ee hart = — Be ee
; (GEG) Soa)
while in general we deduce that c, = O forn = 1,3,5,... , and for even values
of n we write
us alle
Ch
lig 3 m
Semen
= EP
(ten
EP WL
een?
ir a
Eq (0) = ee
—j| m =j| m
Com (S) = i Oh et
eer _S Com(S),
SSP 0
SEC. 9.3 / SOLUTIONS NEAR A REGULAR SINGULAR POINT 313
1 ae | 1
| |Htkcactl MG ensseigen iat ag chiidadtm
iS) m
for the partial sum of the harmonic series. Thus, from (28) we obtain the second
solution
y(x)
; = yi) logx + ec
(7m) ae Jane
The general solution, valid for x > 0, is
y = Ay,(x) + By2(x)
(—1)"(x/ 2)?” ro) (-1)" 'An x 2m
= (A + Blog Se ene 3
ce ta lt Sl lage See ae I a 2 a
An alternate solution technique for finding y2(x) in Example 10 rests upon the use
of (30), which leads to
dx
Yo 0) | ag:
a ‘as a sao k ae Ee
See 1) OE Ai fie 198 13456
By x2 «5x* = 23x°® 1B G6
eea ee e | ee |.
=
yi) logx + ( 4 64 MG * 128 ” 3456
or
314 CHAP. 9 / THE POWER SERIES METHOD
e be ee ‘i ise
yalx) = vila) logx + 7 — 58 + F3e54 (31)
We have therefore obtained the first few terms of the second solution by this
alternate technique. Although it is not immediately obvious, this is the same
solution found in Example 10 for y2, the verification of which is left to the reader.
The solution function y, in Example 10 defines a particular function called the
Bessel function of the first kind of order zero and is denoted by the special symbol
yi = Jo(x). (32)
In most applications involving the Bessel function, it is customary to choose a
certain linear combination of Jp and y, and to take this combination as the “second”
solution of the DE rather than simply to take y). This special combination is called
the Bessel function of the second kind of order zero and is defined by
y a, in ( 1
eSSee 1 ee 1
z og) = (0. S772 :
noo
Using the Bessel functions, we can then express the general solution of
xy" + y' + xy = 0; (34)
Let us suppose s; — s, = N, where N is a positive integer. Thus s, > s>, and in this
case there exist two linearly independent solutions of the form
and 5
yale) Z = - S = 8)
Therefore, we are always assured of one solution of the form (36) corresponding
to the larger root s, of the indicial equation. Using the smaller root, however, we
can always produce two solutions of the DE, although there are two subcases to
consider here. In most situations we can produce the second linearly independent
solution by a procedure similar to that used in Case II in Section 9.3.3 which leads
to a logarithmic type of solution as suggested by (37). An essential point of
dissimilarity between the two cases is the choice co(s) = 8 — 52 in constructing
y(x, s), rather than setting co(s) = 1 as before. By constructing the function y(x, 5)
in this manner, it can be shown that
a » Noe as 0,
Bes > (n ae Genie
n=0 n=0
> ——_——_——_
a,
n—>n-1 n—n-1
which simplifies to
+ s)cn
Be = ir Aslegx® 5 + 5 [7 + s\n + 5 — Ic, — 4
n=1
s; = 5 and 52 = 0 as roots. If
Therefore, the indicial equation s? — 55 = 0 gives
formula reduces to
we try the smaller root s,; = 0, the recurrence
n(n — 5)cn = (n — 3)en-1,. 1 = ee Oe, ole
5, it is best to write out
Since division by (n — 5) is not permitted when n =
separate relations for the c’s through n = 5. Thus,
316 CHAP. 9 / THE POWER SERIES METHOD
n = 3: —6c3 = —0°c2,
cy = ue C2 re oe Con. 3 =c,=0,
and since 0-c; = 0 is satisfied for any value cs, we see that cs is arbitrary. For
n > 5, it follows that
=
(n — 3)Cn-1 Sag)
a7 tae ens
ie Ce) Lae
Proceeding now as usual, we find
cau
6 6:1 5>
4 3-4
a ane OO ee
Polat, |e eenase)
RE per
@: 19" a) (eo) 6; Ue: 6:8 6)iyo) 0) ome 0), (ee) fal. of rer ..0-<e:
y =o ol(1+ *
=x + =x?
|
3 3-4 3-4-5
+ ¢5x5( 1)+is—ex + 7 r r-
CRESS om
es ih eR,
ee) co 1
Fs42 se
+ See = rine ‘= 0,
n=0 n=0
n—n-2
. |
ri, C + s(n +s — Mey + (nm + S)ep + Cn-2 — mas
4 20:
n=2
Cn-2
Ch = n = 2, 3, 4,
‘ nin — |)
We obtain successively
Co Co
——
Oe 2!
C\ Ci
QS SS SS=
caper 320 owae 73!
C= -
C2 =
Co
oS ae
E ey al
C3 Ci
a, pg
5-4 me Nem
5!
and thus
Viera (Cot top On tsk te Cv te)
= = 1/2
Kien 2
cA gs2
kgpice?4 x6
SSeS ew. VP =
CE2 3 ea 5
44 22 es
S7 Ee 6 6 ‘
co 2 Al 6h )+s as 305!
which can also be expressed in the form
cos x sinx
Ween ONE. x >.
yields
Ni —=2)C
= Cay eat Slee
ee
n = 2: O-co
= C1,
and
Cn-1
Bl ees eo Rear! Fake Pete te
: n(n = 2)
These relations can be satisfied only for c} = co = 0, which contradicts our as-
sumption that cp # 0. Therefore we have failed to produce two solutions this time
that do not involve logarithmic terms. (It may at first seem that the way to proceed
is to set co = c; = O and leave c, arbitrary. Doing so will only produce one
solution, which will be reobtained automatically when solving the DE by the
method discussed here.)
Let us write the recurrence formula above in the form
mstnta?
GF "Ps
Cei(S)
CAS a=
iC) epee Oa as
(s + I)(s +3) (s + 1X(s + 3)’
(5) = ao a a
(Set 2) ire 4) (Se 1) (8 Sse a)
c4(S) = =
(s + 4)\(s + 6)
SEC. 9.3 / SOLUTIONS NEAR A REGULAR SINGULAR POINT
Z 1
~ (s + Is + 3)%X(s + 4)°(s + 5)(s + 6)’
and so on. We have
y(x, s) = ey C.CS)x"
n=0
(s ak ace x5t2
(See ei
(Sta 1CSa 3) sea) Gra 3) t- 4)
xst3 +
Gaines Sac
from which we can generate y, by setting s = —2; i.e.,
yi) = ys S2)
= 2 Spee 1Oe eee
% ee Xeae oe x?
on ee!
EPR ESS Sonam Ten ST
or
oc xe .
eure ae
Also,
dy (x, 5)
a
See 5 + 2
ee
(st 2)! |
1 1
eee, ee
1
Oe De A TEENIY ©EEN hee
a 5ue Seis reel za
(sa seats 4) Savi ly se-asiy sy 4
x5*3
and therefore,
320 CHAP. 9 / THE POWER SERIES METHOD
(Ore dy (x, S)
va Os s=-2
1 1 1 11 213
= y,(x) logx + see
ae eh + ee4 + ee 576.Yo cate
36° ee
EXERCISES 9.3
In problems 1-8, locate and classify the singular points of the DE.
1. x°@? — Dy” = x0 + Dy + @— Dy =0
Qi Katee |)Y airy es ye 0
3. x4(0x? + ID — 1)?y" + 40°e — Dy’ + & + Dy =0
4. y" +x =0
5. x*(x — 4)?y" + 3xy’ — @ — dy =0
6. x(x + 1)?y" + (x? — ly’ + 2y =0
7. xy" +(x - 3)"47 =0
8. (x + x4 — 6x>)y” + 3x’y' + — 2)y =0
In problems 9-18, show that the roots of the indicial equation do not differ by an
integer and obtain two linearly independent solutions by the method of Frobenius
about the point x = 0.
9. 2x"
+ (1+ xy’ —2y =0 10. 4xy” + 3y’ + 3y =0
13. 2x?y"
+ 3xy’ —y =0 14. 2x?y”
+ xy' -—y =0
15. 2x*y"— xy’ + (1+ xy =0 16. 2x’y"-—xy’ +(x — 5)y = 0
17. 2x*y" + xy’ + (x? — 3)y =0 18. 2xy” + (1 + 2x)y’ — Sy = 0
In problems 19-26, show that the roots of the indicial equation are equal and
obtain two linearly independent solutions by the method of Frobenius about the
point x = 0.
In problems 27-36, show that the roots of the indicial equation differ by a nonzero
integer and obtain two linearly independent solutions by the method of Frobenius
about the point x = 0.
Nonlog cases
2f. x*y" + 2x(x — 2)y’ + 2(2 — 3x)y = 0 74% aah! So lbs SA yy shy aa)
29. xy” — (x + 3)y’ + 2y = 0 30. x?y" + x’y’ — 2y =0
31. x*y” + x?7y’ H& = 2)y = 0
Log cases
32. xy" +y =0 RK G3) Salers
yyy arty 0)
34. x7y" + (x? — 3x)y’ + 3y =0 BOs etcy eK lat) yeh Sven 0
36. x*y" + xy’ + @? -— Dy =0
*37. The DE
x1 — xy" +[c —- (a + b + I)x]y’ — aby = 0,
where a, b, and c are all constants, is called the hypergeometric equation, and its
solutions are called hypergeometric functions.
(a) Show that x = 0 and x = 1 are R.S.P.’s of the DE.
(b) Assuming c #0, —1, —2,... , show that one solution of this DE is
REFERENCES
Boyce, W. E., and DiPrima, R. C. 1977. Elementary differential equations and bound-
ary value problems. 3rd ed. New York: Wiley.
CHAPTER 1
Section 1.2
9. linear, first-order itl linear, second-order 13. linear in x, first-order 15. nonlinear, third-order
Section 1.3
isle Se 18)
23. y= 25. y = Asinx + Bsin3x 29. xy’ — 3y = 12 31. y’ — 2y = 4x
ae EE
33. (y’? — 9y7=1 35. DVN — 2 Aeetny 37. y =O (singular solution) 39. no
Section 1.4
A ee ke Oe its
CHAPTER 2
Section 2.2
mare : il ae 59 j (-aF *)
ex to = : =
oe A ley OX,
13.
(x? ik {2 = Cy ae ive.
1S: = C G7) Pea a G
323
324 ANSWERS TO ODD-NUMBERED EXERCISES
(singular solution)
(c) no solution
Section 2.3
xy(y — 1) =C
a it y-¢
13. (xy #y2e7 =C 15. 2x — log? +y7)=C -=-x7¥°\+ C
y 2
3y 4,3 5,,3 y a Gremean
19. @+y)+—=C a. xy Grxyj=e
Y 4 2 1 2 1
5. y =Cie~* 27. y=e*|C,+ ae im Aesr ie + log(y + V1 + y’)]
y
1 u
33.
33
2logx
Ogx
+>y -—-
x
ty? =
3 2
5 log u + v*) + tan! (“) G
1 x
logy + —=tan '(
va") =C (x — y)logx +ylogy = Cx + y
& J) y
41. 2x + 2y)
+ (@ + y)log + y) =0 x’y3 — 3xy + 2y7=C
1
ae a + le ty dee yaad
a re gh Bee
4x3y? = may a
49. ES —y)= a +C
» 3
ax”
— dy? + 2bxy = 2c
Section 2.4
Section 2.5
1. 1 Cie bs 3
y area Lee
3. y =-(sinx + C;) = cosx 5. y=x‘[@ - le*+ Ci]
xX
Ve .
13. y=(@+Ci)e™ ecole dleeach (3 17. y= b k
Gamat Gixe
a
ANSWERS TO ODD-NUMBERED EXERCISES 325
E 1
19. yy = 5) + Ciel*
21. i(t) = = es By Hy 23. y = 5(2x + 3)" log (2x + 3)
a5 sinx — seca 1+ 77/4
x
be?
ten 2 OS = 1+ x?’
27. = 350 yi
lat Cem
2 fe=G Ye 40 feet Ae? i pete
1+ x?
Section 2.6
LG:
3. y?=— + C,x* §&. xy =2+.Cy'"
1 ee
CHAPTER 3
Section 3.2
Con x?
9. = log{3 -—>]
2logy
2x?
+k Wye
5/3
= 5
ek 15. r = b(1 + sin@) 17. r?>=bsin@
uy
1 — g~2Vcgimt
yell) 1]
29. 3.1 mi/sec Sills v = 493 ft/sec 33) approx. 16 hr, 16 min
27. Wo
o=sfres[(3) - Th .
k k]2) 1/2
y? = 2Cx + C’ (parabolic shape)
2\\a k
Section 3.3
dM
M(0)=M, 11. v(t) = 400 — els)
(b) k = 1: M(10) = 1.790Mo (c) ae = 0.06M,
9.
M(t) = Meo
k = 4: M(10) = 1.814Mo
k = 365: M(10) = 1.821 Mo
M(10) = 1.822Mo
EoCw w
m Doc
13. (a) 1 = hoa(1+2) 15.) i(t) = 1 — R°C2w? T= R200.
Cc mg
(b) 0.884
ANSWERS TO ODD-NUMBERED EXERCISES
E,
(a) i(t) = rl aa emtey te ine —Rt/L 19. 58.5°
CHAPTER 4
Section 4.2
Section 4.3
y= e 2 3. y= xe>*
1. 5. y2 = cosx 7. yo = e*sin2x
Ih sp oe ae
9. yo=x WW. y=x'? 13. y. = log 15. y.’=x ““cosx
re
Section 4.4
2 Bt AVA]
y =e SNS ye! + re! 23. 4D* — 17D — 15 25. D?-1
Section 4.5
V3 V3
y = (C, + Cox + Cax2)e™* 15, “yas Cer i e.* (cscos*Ss + Cassin)
a y=
y=
C, aia Gene
Gien ae Grew:
ar C3e*
te Gren
19. ye C\ + (C2 =i Crew
25. .y = Clem + Coen + Cre? + Cre” 27. y =C, + Cox + Cx? + (Ca + Csx) cos 3x + (Co + C7x) sin 3x
+ (Cg + Cox + Cyox? + Cix*)e*
29. y = Ci + (C2 + C3x) cosx 31. y =(L—}x)e™
+ (Cy + Csx) sinx 33. y =2+ 3x + 2x? + 3x?
QS) y = (Ci + Cox + Cax? + Cux*Je™ + e*(Cs + Cox) cosx + (Cy + Cax)sinx + Coe + Cyoe™*
37. (D*® — 7D? + 17D — 15)y =0 39. y =e *[{(C, + Cox) cos2x + (C3 + Cyx) sin2x]
41. (a) y=0 (b) y = Csinnzx, where C is arbitrary andn = 1, 2,...
Section 4.6
Section 4.7
Section 4.8
CHAPTER 5
Section 5.2
5 27 - 7 1 ‘
paar —ed
—Hz, 5) OUOM a
(a) —=s OE
(b eee
= —cos(8V3t 7. 29 ticks
23. 1.069; no
Section 5.3
‘ 2s cosi2 ta OL tae
eae ee ae pene 4) 2003 | ae oO
i aa) + — = L y — a —
P
7. 0.0056 Hz, 2m 13. 5.56s, 8.85 15. e ‘sint (transient), 2sint(steady-state) 17. ’
Section 5.4
Se (a) 0. (b) ip(t) = cos2t + Hsin 2t 5. (a) i(t) =e *(3sin4t — 4cos4t) + 4cosSt
(b) i(t) = Se = sint
7. q(t) =3—4e ' (cos
10¢ + sin102), 3 coulombs 9. R = 9.41 X 10°-*ohms
Section 5.5
1. y =sint 3. y = 2e*
— 3e' 5. y=3-e"
1
7. y = (2 — 3logt)t? 9. g(t, T)=t—T 11. g,(t, rT) = —=sinh
V5(t — 7)
F : V5
t
13. gi(t, 7) = —e 7? — e 2 15. g(t, T) = =|() - (:)
3 8L\r t
1 (ett) Fa) 1
17. gilt,t, T)
T) =
= 5
-(1 — eet
7?) log Ki en 19. as
SS 211. ye Cs
=e al
3) Wiis 00 ental = ote mat (Se tiem en! 29. y=aro—gr' +t
31. y=pr?+ eet %— Zlogt 37. gi(t, rT) = He" — e” ) — (t — 7)
(a + 1)e""? — (b + Ie" + (b — ade —3 — V33 —3 + V33
3900210.) ee
—— where a = ————_,,_ 6 = —————
(GE 70) (Gaeta) (Dasa) 4 4
41. git, 7) = pe 7 — ce 9 + Fe 45. y =jcos2t + ¢sin2t — 4
ANSWERS TO ODD-NUMBERED EXERCISES 329
Section 5.6
2sint, O<t ee
5. y(t) = a Son
2sint — A costae tf ae
V3
(t + 3)e'— 2eN cos——t, Oma
V3
7% yM=4(4+3)e'- 2e" cost +e!
= | Leh — 1) + — ayesin—(t
cos——(f Ney, — | : fal
2 se
CHAPTER 6
Section 6.2
1 a 3 n! 5 Ss i= @ §
Section 6.3
3 2k(Bs? — k? s? 3(s + 4 +4
12(s
1. 3. ae ft rarer @ oT
(s +4?+16 x [(s + 4)? + 16]
(s — 2) (s? +k’? s* + 4k
s+1 (s +17? +2 fl F 1 i
9. | 144.. ——_—_—_—_—. 13. (a) -+—-(l-e” b) -(l-e?
ae 1 (s + I[(s + 1)? + 4] (a) 5 sa ) ( A )
ise
3 caret
17 . —arctan— 25. : (b) T er(== )
s? + 16 sae 2Vs
Section 6.4
CYR sO —4
Whe 1
\ —t
4,..=3
g 5. e'sint Te Wr
1. aie (3) 2 e
9. Whossinst
e-2'( 7cosSt ~ = | —21/3
41) aS : 2
Scos=V5t ae
= 4V5sin5V51 iksk il>@ a
12
«(3
3
ay ee ser
a ie 4s
2
OP
—2t a
Sa Lie has 2)
330 ANSWERS TO ODD-NUMBERED EXERCISES
Section 6.5
y(t) = 3(e7 — 3) 3. y(t) = sinht 5. y(t) =sint 7. y(t) =e' — (cost + sint)
1.
ti 1 sae
y(t) =(e ‘— 1)sint 11. pipet cet +e (eos Vir_ V21)
sin
3V2
1
3) y(t) =¢(1
8 + 20) + fe7"(1 — 20 15. y(t) = 8e '— 9e-* + 3e°*
17. yQ=Fe'—Se WV +he *+ he" 19. y@= 4(sinht Se sinit) at
Section 6.6
\ x Site POD)
Ol eens 11.
se s+
49) en — 3)
21. cos2i[1 + h(t — 7)
Ws 26] Qo = 2) Se
23. y#=e'+[1l—e “ne - 1) 95) y(t) = cos2t +t —3sint — [ft — 1 — 5sin2(t — IJa(t — 1)
7) y(t) = (cos2t — cos 4t)[1 + A(t — )] + 3sin2r 29. y =e" = e% + 2[1 + 20 —3e™ he
31) y@)i— ex (Cost + sind) —e) “aa sinth( = am)
33. y(t) =2te'+[1—e *? —(¢ —2m)e h(t — 277)
Ip 1 1 k
5) y(t) = — (“sin Wot — —sin ar)- sin wo(t — 7), where wo =—
m(@~ — wo) \@Wo @ MW m
Section 6.7
2 S = ies P
a.) ——— Si So 5 Cede ie ee — tea Sig
s(s* + 4) (GS se AIG? ae 1}
9 a.75| 2 + St)e"!
(4 ak 50
Ree ad cos Di eosin
air sin f) 15 A ee at
pain
oe 17 * y(t)
) = =t + aeaL
sindt
2 P
CHAPTER 7
Section 7.2
Section 7.3
Section 7.4
Kee =%
3. x1 5. u’=w
x3 = —kx, + Psinwt a Xa Demat aa 10
Xj =Xq nae AM Gn:
xq = k*x, + fd)
(7) ees 9. x(t) = Cie’ + 3C2e' Tioex) = Cres ot ere
1 ka y(t) = Cie’ ae SGoeme y(t) = —3Cie "+ Cre™
x3 = =——(ki AP kz) x, 3
m m
X3 = X4
ky kp
xj =—xX1 - —%3
Section 7.5
ue 2cos 3t , 2 sin 3t 24
SE EA Bae, = ee sf“Aer ie sna)
(1 1
1h Lew
7 Y@=Gl + Cl 9! |e*
6: |)46 —2: |e" 9. veo =35 (flew +(
=13 i =|
ioe 1.543 1.175 cosht sinht
ee) i ‘| (b) tee a | (c) ea on
13. Y= (2)cost + (?)sint|ar cl
c;| ()cost + (3)sint|+ ie, aa + (axa )
Section 7.6
CHAPTER 8
Section 8.2
Section 8.3
CHAPTER 9
Section 9.2
Suh ah
ra Q > — exe
1 ne ce 2)(n +k >)
Se a us
. ee
= aera ae (oi \)))!
Section 9.3
11. y =A(U —x + x?
15, y = Axl —x + bx? —- ++) + Bx(l — 3x 4 ox’ — °°")
17 ays AG (let 3h ak oe ) + Bx2(1 — yx? + oex*— ++ *)
19. + 4x + 4x7 +--—
y =(A + Blogx)(1 BBx + 12x? + ex? +--+ -)
+)
334 ANSWERS TO ODD-NUMBERED EXERCISES
21. y =(A + Blogx)xe* — B(x? + 3x? + 3x4 ee) 23. y =(A + Blogx)x”
25. y =(A + Blogx)x — B(x? — 4x° + 7gx* — + -*)
27. yy = colx — 2x? + 2x?) + Cn(x* — Fx? + 4x°— >>")
29. y =co(l + 3x + x2) + ca(x* + 3x5 + oxo t+: °°) 31. y = cox”! + c3(x? + 3x? + xt +: °
1 4
33. y=(A + Blogry ~ 12+ 32x ~ 40s? +) + B(5 42-4 16x ++)
go
5 1
35. y =(A + Blogs)(~3e ~ 4x 3? = +++) + a(S -2-x 3+)
x
ABEL, NEILS H., 89 Curves of linear pursuit, 63-65
Abel’s formula, 89, 103 Cycloid, 69
system of DEs, 248
Amplitude, 131 Damped motion, 130, 198
maximum, 142 forced, 141-144
Asymptotic stability, 262 free, 134-137
Autonomous system, 261 Decay, radioactive, 70
Auxiliary conditions, 14, 127 Delta sequence, 164
Auxiliary equation, 94, 104 Differential equation, 2
system of DEs, 239 classification of, 2
exact, 25—28
Beats, 140 first-order, 19-54
BERNOULLI, JAKOB, 52 general solution of, 9-12
Bernoulli’s equation, 52-54 higher-order, 82-125
Bessel function, 187, 201, 214 homogeneous, 31-33, 36-39, 83, 101
first kind, 314 linear, 3
second kind, 314 nonhomogeneous, 36, 41-50, 83, 101, 108-114
zeros of, 168 nonlinear, 3
Bessel’s equation, 83, 201 order of, 3
standard form, 166 ordinary, 3
Boundary value problem, 14-16 partial, 3
Brachistochrone, 69 solution of, 6-12
systems of, 220-273
Catenary, 69 Differential operator, 36, 98
Cauchy-Euler equation, 121-123 normal, 36, 101, 127
Center, 264 Dirac, PAut A. M., 160
Characteristic equation. See Auxiliary equation Dirac delta function. See Impulse function
CHEBYSHEV, PAFNUTI LvovicH, 305 Discontinuous function, 49, 201-208
Chebyshev’s equation, 305
Clairaut’s equation, 35 Eigenvalues, 250
Comparison theorem, Sturm, 167 Eigenvectors, 250
Complementary solution, 108 Electrical networks, 230-232
Confluent hypergeometric equation, 321 Electric circuits, 75, 146-148
Confluent hypergeometric function, 321 Electromotive force, 147
Constant-coefficient equations, 94-99, 104-106
Equidimensional equation, 121-123
linear system of DEs, 239-247 Error truncation, 276
Convolution theorem, 210-212 Error function, 47, 186
Critical damping, 135, 198 complementary, 186
Critical point, 261 Escape velocity, 62
335
336 INDEX
i
= a
XN
i,
“ as 2: 4 .
: “Oe
ns ee re
a a
<
; pete a na
= ost ’
x
: x
;
; i.
> \ Ss
1S,
3
‘-
f ‘
%
2cat 4 <a
ee
~
ie
¥. i
=) * +
ce x ite ra a
pe
ss med : 3
+ oe : if y /
f os =
i a
+ \ ‘
rhe :
oe
oo
f
ia
: ; ©
| 7
: i. ‘
el Aa te
- : Rohs Nz
ii
a‘ i
;
eae
1OOU
:
@)
yA aoe Ene : . ; a i . ; :