Garrett Birkhoff, Gian-Carlo Rota - Ordinary Differential Equations-Wiley (1989)
Garrett Birkhoff, Gian-Carlo Rota - Ordinary Differential Equations-Wiley (1989)
DIFFERENTIAL
EQUATIONS
FOURTH EDITION
GARRETT BIRKHOFF
GIAN-CARLO ROTA
ORDINARY
DIFFERENTIAL
EQUATIONS
FOURTH EDITION
Garrett Birkhoff
Harvard University
Gian-Carlo Rota
Massachusetts Institute of Technology
cy
WILEY
Bibliography: p. 392
Includes index.
1. Differential equations. I. Rota, Gian-Carlo,
1932- II. Title.
QA372.B58 1989 515.3’52 88-14231
ISBN 0-471-86003-4
1098 765 4 3 2 1
PREFACE
The theory of differential equations is distinguished for the wealth of its ideas
and methods. Although this richness makes the subject attractive as a field of
research, the inevitably hasty presentation of its many methods in elementary
courses leaves many students confused. One of the chief aims of the present
text is to provide a smooth transition from memorized formulas to the critical
understanding of basic theorems and their proofs.
We have tried to present a balanced account of the most important key ideas
of the subject in their simplest context, often that of second-order equations.
We have deliberately avoided the systematic elaboration of these key ideas, feel-
ing that this is often best done by the students themselves. After they have
grasped the underlying methods, they can often best develop mastery by gen-
eralizing them (say, to higher-order equations or to systems) by their own
efforts.
Our exposition presupposes primarily the calculus and some experience with
the formal manipulation of elementary differential equations. Beyond this
requirement, only an acquaintance with vectors, matrices, and elementary com-
plex functions is assumed throughout most of the book.
In this fourth edition, the first eight chapters have again been carefully
revised. Thus simple numerical methods, which provide convincing empirical
evidence for the well-posedness of initial value problems, are already introduced
in the first chapter. Without compromising our emphasis on advanced ideas and
proofs, we have supplied detailed reviews of elementary facts for convenient
reference. Valuable criticisms and suggestions by Calvin Wilcox have helped to
eliminate many obscurities and troublesome errors.
The book falls broadly into three parts. Chapters 1 through 4 constitute a
review of material to which, presumably, the student has already been exposed
in elementary courses. The review serves two purposes: first, to fill the inevitable
gaps in the student’s mastery of the elements of the subject, and, second, to give
a rigorous presentation of the material, which is motivated by simple examples.
This part covers elementary methods of integration of first-order, second-order
linear, and nth-order linear constant-coefficient, differential equations. Besides
reviewing elementary methods, Chapter 3 introduces the concepts of transfer
function and the Nyquist diagram with their relation to Green’s functions.
Although widely used in communications engineering for many years, these con-
cepts are ignored in most textbooks on differential equations. Finally, Chapter
Vv
vi Preface
Garrett Birkhoff
Gian-Carlo Rota
Cambridge, Massachusetts
CONTENTS
Introduction 1
Fundamental Theorem of the Calculus 2
First-order Linear Equations 7
Separable Equations 9
Quasilinear Equations; Implicit Solutions 11
Exact Differentials; Integrating Factors 15
Linear Fractional Equations 17
Graphical and Numerical Integration 20
The Initial Value Problem 24
*10 Uniqueness and Continuity 26
*11 A Comparison Theorem 29
*12 Regular and Normal Curve Families 31
Bases of Solutions 34
Initial Value Problems 37
Qualitative Behavior; Stability 39
Uniqueness Theorem 40
The Wronskian 43
Separation and Comparison Theorems 47
The Phase Plane 49
Adjoint Operators; Lagrange Identity 54
Green’s Functions 58
*10 Two-endpoint Problems 63
*11 Green’s Functions, II 65
6. Stability 85
7. The Transfer Function 86
*8. The Nyquist Diagram 90
*9,. The Green’s Function 93
Introduction 99
Method of Undetermined Coefficients 101
More Examples 105
Three First-order DEs 107
Analytic Functions 110
Method of Majorants 113
*7 Sine and Cosine Functions 116
*8 Bessel Functions 117
9 First-order Nonlinear DEs 121
10 Radius of Convergence 124
*11 Method of Majorants, II 126
*12 Complex Solutions 128
1 Introduction 170
2 Lipschitz conditions 172
3 Well-posed Problems 174
4 Continuity 177
*5 Normal Systems 180
Equivalent Integral Equation 183
Successive Approximation 185
Linear Systems 188
Local Existence Theorem 190
Contents x
1 Introduction 204
2 Error Bounds 205
*3 Deviation and Error 207
4 Mesb-halving; Richardson Extrapolation 210
5 Midpoint Quadrature 212
6 Trapezoidal Quadrature 215
7, Trapezoidal Integration 218
8 The Improved Euler Method 222
+9, The Modified Euler Method 224
*10. Cumulative Error Bound 226
1 Introduction 261
*2 Movable Singular Points 263
First-order Linear Equations 264
Continuation Principle; Circuit Matrix 268
Canonical Bases 270
Regular Singular Points 274
Bessel Equation 276
The Fundamental Theorem 281
*9 Alternative Proof of the Fundamental Theorem 285
*10 Hypergeometric Functions 287
*11 The Jacobi Polynomials 289
*12 Singular Points at Infinity 292
*13 Fuchsian Equations 294
x Contents
BIBLIOGRAPHY 392
INDEX 395
CHAPTER 1
FIRST-ORDER
DIFFERENTIAL
EQUATIONS
1 INTRODUCTION
o(x,9,9’) = 0
A DE of the form (1) is often said to be of the first degree. This is because, con-
sidered as a polynomial in the derivative of highest order, y’, it is of the first
degree.
One might think that it would therefore be called “linear,” but this name is
reserved (within the class of first-order DEs) for DEs of the much more special
form a(x)y’ + b(x)y + ¢(x) = 0, which are linear in y and its derivatives. Such
“linear” DEs will be taken up in §3, and we shall call first-order DEs of the more
general form (1) quasilinear.
A primary aim of the study of differential equations is to find their solutions—
that is, functions y = f(x) which satisfy them. In this chapter, we will deal with
the following special case of the problem of “solving’’ given DEs.
The problem of solving (1) for given functions M(x,y) and N(x,y) is thus to
determine all real functions y = f(x) which satisfy (1), that is, all its solutions.
(2) x + yy =0
The solutions of (2) can be found by considering the formula d(x? + y*) /dx =
=
x =
Therefore, although the pairs of semicircles in Figure 1.1 appear to join together
to form the full circle x? + x = C, the latter is not a “‘solution curve’ of (1). In
fact, no solution curve of (2) can cross the x-axis (except possibly at the origin),
because on the x-axis y = 0 the DE (2) implies x 0 for any finite y’.
=
=
The preceding difficulty also arises if one tries to solve the DE (2) for y’. Divid-
ing through by y, one gets y’ —x/y, an equation which cannot be satisfied if
=
=
(3) y = F(x,y)
In the normal form y’ = —x/y of the DE (2), the function F(x,y) is continuous
in the upper half-plane y > 0 and in the lower half-plane where y < 0; it is
undefined on the x-axis.
{ In this book, the word “function”’ will always mean single-valued function, unless the contrary is
expressly specified.
2 Fundamental Theorem of the Calculus 3
standing of its capabilities can only be achieved if its definitions and results are
formulated precisely. Some of its most difficult results concern the existence and
uniqueness of solutions. The nature of such existence and uniqueness theorems
is well illustrated by the most familiar (and simplest!) class of ordinary DEs.
These are the first-order DEs of the very special form
(4) y¥ = g(x)
Such DEs are normal; their solutions are described by the fundamental theorem
of the calculus, which reads as follows.
is defined for each fixed x as a limit of Riemann sums; it is not necessary to find
a formal expression for the indefinite integral [g(x) dx to give meaning to the
definite integral {% g(t) dt, provided only that g(t) is continuous. Such functions
4 CHAPTER 1 First-Order Differential Equations
(6) y=y-l
Since »? — 1 = (y + 1)(y — 1), the constant functionsy = —1 and y = 1 are
particular solutions of (6). Since y* > 1 if |y| > 1 whereas y? < 1 if —1 < y
_——
a
—x2
Figure 1.2 Solution curves of y’ = e
< 1, all solutions are decreasing functions in the strip |y| < 1 and increasing
functions outside it; see Figure 1.3.
Using the partial fraction decomposition 2/(y? — 1) = 1/(y — 1) — 1/(y +
1), one can rewrite (6) as 2 dx = dy/(y — 1) — dy/(y + 1) from which we obtain,
by integrating, 2(x — c) = In |(y — 1)/(y + I)|. Exponentiating both sides, we
get te*%*-9 = (y — 1)/(y + 1), which reduces after some manipulation to
1 + e%-9 tanh
(6/) ,= 1 Ft ee4 =
| coth
Jes
This procedure “‘loses’”’ the special solutions y = 1 and y = —1, but gives all
others. Note that if y = f(x) is a solution of (6), then so is 1/y = 1/f(x), as can
be directly verified from (6) (provided y # 0).
dx 1 2
(6”)
—_—
dy
= oe
yy
a
yt]
a
y-1 J |
The DE (6”) can be integrated termwise to give, after some manipulation,
x =4In|1—y?| + Cory = [1 F exp (2x — 77,k = Qe.
EXERCISES A
1 (a) Show that if f(x) satisfies (6), then so do 1/f(x) and —/f(—x).
(b) Explain how these facts relate to Figure 1.2.
Show that every solution curve (6’) of (6) is equivalent under horizontal translation
and/or reflection in the x-axis toy = (1 + e*)/(1 — e*) or toy = (1 — e*)
(1 +
e**),
(a) Show that if y’ = y? + 1, then y is an increasing function and x = arctan y + ¢.
(b) Infer that no solution of y’ = y® + 1 can be defined on an interval of length
exceeding 7.
CHAPTER 1 First-Order Differential Equations
i AN SSS
yr
yr-t
Show that the solution curves of y’ = y* are the x-axis and rectangular hyperbolas
having this for one asymptote. [HInT: Rewrite y’ = y? as dy/y? = dx.]
Sketch sample solution curves to indicate the qualitative behavior of the solutions of
the following DEs: (a) y’ = 1 — y°, (b) y’ = sin xy, (©) y’ = sin? y.
Show that the solutions of y’ = g(y), for any continuous function g, are either all
increasing functions or all decreasing functions in any strip y,_; < y < y, between
successive zeros of g(y) [i.e., values y,, such that g(y) = 0].
Show that the solutions of y’ = g(y) are convex up or convex down for given y accord-
ing as {g| is an increasing or decreasing function of y there.
LLL
——=>
——
+
aoe
EE ye
WV\Y\\
Figure 1.4 Solution curves of y’ = y* — y.
3 First-Order Linear Equations 7
*8. (a) Prove in detail that any nonconstant solution of (6) must satisfy
x=ct+4in
ly — D/yt+ DI
(b) Solve (6”) in detail, discussing the case k = 0 and the limiting case k = © (y =
0).
*9. (a) Show that the choice k < 0 in (6) gives solutions in the strip —1 < y < 1.
(b) Show that the choice k = 1 gives two solutions having the positive and negative
y-axes for asymptotes, respectively.
In the next five sections, we will recall some very elementary, but extremely
useful methods for solving important special families of first-order DEs. We
begin with the first-order linear DE
if and (since e”” ¥ 0) only if y satisfies (8). This proves the following result.
We can treat the general case of (8) similarly. Differentiating the function
P(x)
ey, where P(x) is as before, we get
It follows that, for some constant yo, we must have ¢P(x)J = 9) — fz a(t) at,
whence
(8’)
y = yeP® — 9PO)f ° ea(t) dt
a
Conversely, formula (8’) defines a solution of (8) with y(a) = yo for every yo, by
the Fundamental Theorem of the Calculus. This proves
Initial Value Problem. In general, the “initial value problem” fora first-
order DE y’ = F(x,y) consists in finding a solution y = g(x) that satisfies an initial
condition y(a) = yo, where a and yo are given constants. Theorem2 states that
the initial value problem always has one and only one solution for a linear DE
(8), on any interval a S x S 5b where p(x) and q(x) are defined and continuous.
Remark. There are often easier ways to solve linear DEs than substitution in
(8’). This fact is illustrated by the following example.
(9) yty=xt3
+ See the book by Dwight listed in the Bibliography. Kamke’s book listed there contains an extremely
useful catalog of solutions of DEs not of the form y’ = g(x). For a bibliography of function tables,
see Fletcher, Miller, and Rosenhead.
4 Separable Equations 9
4 SEPARABLE EQUATIONS
(10) yx = g(x)h(y)
d(x) + Wy) = C
of the function U(x,y) = (x) + (y) are solution curves of the DEs (11) and
(11’). Moreover, the Fundamental Theorem of the Calculus assures us of the
existence of such antiderivatives. Likewise, for any indefinite integrals G(x) =
Sg(x) dx and H(y) = Sdy/h(y), the level curves of
G(x) — H(y) = C
(11”) y = HC — Gi]
However, the solutions defined in this way are only local. They are defined by
the Inverse Function Theorem,} but only in intervals of monotonicity of H(y)
where h(y) and hence H’(y) = 1/h(y) has constant sign. Moreover, the range
of H(y) may be bounded, as in the case of the DE y’ = I + 9. In this case,
+ This theorem states that if H(y) is a strictly monotonic map of [c,d] onto [4,6], then H~!('y) is single-
valued and monotonic from [a,b] to [c,d].
10 CHAPTER 1 First-Order Differential Equations
vo — 90) = f 2 — J no ax =e
gives for each c a solution of y’ = g(x)h(y) in the strip y, < y < yg. Near any x
with y, — ¢ < $(x) < yg — ¢, this solution is defined by the inverse function
theorem, by the formula y = y~'(¢(x) + 0).
—1/g(x)h(y).
Critical Points. Points where du/dx du/dy = 0 are called critical points
=
=
of the function u(x,y). Note that the directions of level lines and gradient lines
may be very irregular near critical points; consider those of the functions x? +
y® near their critical point (0,0).
As will be explained in §5, the level curves of any function u € @'(D) satisfy
the DE du/dx + y’du/dy = 0 in D, except at critical points of u. Clearly, their
orthogonal trajectories are the solution curves of du/dy = y’du/dx, and so are
everywhere tangent to the direction of Vu grad u (0u/dx, du/dy). Curves
= =>
= =
having this property are called gradient curves of u. Hence the gradient curves
of u are orthogonal trajectories of its level curves, except perhaps at critical
points.
5 Quasilinear Equations; Implicit Solutions 11
EXERCISES B
1 Find the solution of the DE xy’ + 3y = 0 that satisfies the initial condition
fl) = 1.
(*) y = ay + bsin
kt
(**) y t=
x? + 4x47
(b) Find a solution of the DE (*) that satisfies the initial condition y(0) = 0.
Show that if & is a nonzero constant and q(x) a polynomial of degree n, then the DE
xy’ + y = q(x) has exactly one polynomial solution of degree n.
In Exs. 8 and 9, solve the DE shown and discuss its solutions qualitatively.
In this section and the next, we consider the general problem of solving
quasilinear DEs (1), which we rewrite as
to bring out the latent symmetry between the roles of x and y. Such DEs arise
naturally if we consider the level curves of functions. If G(x,y) is any continuously
differentiable function, then the DE
0G
(12’/
ox
(x,y) + > (x,y) = 0
12 CHAPTER 1 First-Order Differential Equations
is satisfied on any level curve G(x,y) = C, at all points where dG/dy # 0. This
DE is of the form (1), with M(x,y) = 8G/dx and N(x,y) = 8G/dy.
For this reason, any function G which is related in the foregoing way to a
quasilinear DE (1) or (12), or to a nonzero multiple of (12) of the form
'
+ See Apostol, Vol. 1, p. 252. Here and later, page references to authors refer to the books listed in
the selected bibliography.
5 Quasilinear Equations; Implicit Solutions 13
xe + y and there is one such curve, the x-axis y = 0; this divides the plane into
two subdomains, the half-planes y > 0 and y < 0. Moreover, the locus (set) where
du/dy = 0 consists of the points where the circles u const have vertical tan-
=
=
gents and the “‘critical point” (0,0) where du/dx = du/dy = 0—that is, where
the surface z u(x,y) has a horizontal tangent plane.
=
=
This situation is typical: for most functions u(x,y), the partial derivative du/d
y vanishes on isolated curves that divide the (x,y)-plane into a number of regions
where 0u/dy # 0 has constant sign, and hence in which the Implicit Function
Theorem applies.
THEOREM 3. In any domain where du/dy # 0, the level curves of any function
u € @' are solution curves of the quasilinear DE
Proof. By the Chain Rule, du/dx = du/dx + (du/dy)y’ along any curve y =
J (x). Hence, such a curve is a level curve of u if and only if
du Ou ou /
0
=
—> = ——
=
dx ox ay
By the Implicit Function Theorem, the level curves of u, being graphs of func-
tions y = f(x) in domains where df/dy # 0, are therefore solution curves of the
quasilinear DE (13). In the normal form y’ = F(x,y) of this DE, therefore, F(x,y)
=
=
— (du/dx)/(0u/dy) becomes infinite precisely when du/dy = 0.
To describe the relationship between the DE (13) and the function u, we need
a new notion.
+ Courant and John, Vol. 2, p. 218. We will reconsider the Implicit Function Theorem in greater
depth in §12.
14 CHAPTER 1 First-Order Differential Equations
7.
is \,
SZN
Figure 1.5 Level curves
c = 0, +1, +2, +3, +4, +6, +9, +12 of
y? — 2
6 Exact Differentials; Integrating Factors 15
is said to be exact when there exists a function U(x,y) of which it is the ‘total
differential’, so that QU/dx = M(x,y) and 0U/dy = N(x,y), or equivalently
aU 0
(14”) + ey dy = M(x,y) dx + N(x,y) dy
dU = — dx
Ox 0
Since dU = 0 on any solution curve of the DE (14), we see that solution curves
of (14) must lie on level curves of U, just as in the “separable variable”’ case.
Since 0°U/dxdy = 6°U/dydx, clearly a necessary condition for (14’) to be an
exact differential is that 0N/dx 0M/dy. It is shown in the calculus that the
—
=
converse is also true locally. More precisely, the following result is true.
(uM) dx + (uN) dy = du
16 CHAPTER 1 First-Order Differential Equations
is an exact differential. The contour lines u(x,y) = C will then again be integral
curves of the DE M(x,y) + N (x,y)y’ = 0 because du/dx = u(M +Ny’) = 0; and
segments of these contour lines between points of vertical tangency will be solu-
tion curves. Such a function yg is called an integrating factor.
Thus, as we saw in §3, for any indefinite integral P(x) = [p(x) dx of p(x), the
function exp {P(x)} is an integrating factor for the linear DE (8). Likewise, the
function 1/h(x) is an integrating factor for the separable DE (11).
The differential x dy — y dx furnishes another interesting example. It has an
integrating factor in the right half-plane x > 0 of the form u(x) = 1/x?, since
dy/x — y dx/x® = d(y/x); cf. Ex. C11. A more interesting integrating factor is
1/(x? + y*). Indeed, the function
is the angle made with the positive x-axis by the vector (x,y). That is, it is just the
polar angle @ when the point (x,y) is expressed in polar coordinates. Therefore,
the integral curves of xy’ = y in the domain x > 0 are the radii 6 = C, where
—a/2 <0 < w/2; the solution curves are the same.
Note that the differential (x dy — y dx)/(x® + y°) is not exact in the punctured
plane, consisting of the x,y-plane with the origin deleted. For @ changes by 2x
in going around the origin. This is possible, even though 9[x/(x? + y°)]/dx = 0
[—y/(x? + y*)]/dy, because the punctured plane is not a simply connected
domain.
Still another integrating factor of x dy — y dx is 1/xy, which replaces x dy —
y dx 0 by dy/y = dx/x, or In |y| = In |x| + Cin the interior of each of the
=
=
four quadrants into which the coordinate axes divide the (x,y)-plane. Exponen-
tiating both sides, we get y = kx.
A less simple example concerns the DE x(x* — 2y°)y’ = (2x3 — y*)y. Here an
integrating factor is 1/x*y?. If we divide the given DE by x*y®, we get
2 2 _ Qxy _ xty! — y* + Qxy>y!
d x
J
dx ( J x
xy?
EXERCISES C
1 Find an integral of the DE y’ = y?/x?, and plot its integral curves. Locate its critical
points, if any.
Sketch the level curves and gradient lines of the function x° + 3x°y + 9°. What are
its critical points?
1
y= Qx+y
For what pairs of positive integers n,r is the function |x|" of class @’?
In Exs. 8 and 9, solve the DE exhibited, sketch its solution curves, and describe them
qualitatively:
1
8 ¥ = y/x— x? 9. 9 = y/x — In [x] °
10 Find all solutions of the DE |x| + jy[y’ = 0. In which regions of the plane is the
differential on the left side exact?
*11. Show that the reciprocal of any homogeneous quadratic function Q(x) = Ax? +
2Bxy + Cy? is an integrating factor of x dy — y dx.
*12. Show that if u and v are both integrals of the DE M(x,y) + N(x,y)y’ = 0, then so
are u + v, uv except where v =
=
0, Aw + wv for any constants A and p, and g(u)
for any single-valued function g.
*13, (a) What are the level lines and critical points of sin (« + 4)?
(b) Show that for
u = sin (x + 4), (xo.¥o) = (0,0), andé = «= 4 F(x,¢) in the Implicit
Function Theorem need not exist if 7 < }while it may not be unique if n > 4.
dy _ ox +dy
(16) ad
# be
dx ax + by’
18 CHAPTER 1 First-Order Differential Equations
e+ dv
xv’ +o =
a+ bv
(a + bv) dv dx
=0
bv? + (a — dv —c x
Since the integrands are rational functions, this can be integrated in terms of
elementary functions. Thus, x can be expressed as a function of v = y/x: we
have x = kG(y/x), where
More generally, any DE of the form y’ = F(y/x) can be treated similarly. Set-
ting v = y/x and differentiating y = xv, we get xv’ + v = F(v). This is clearly
equivalent to the separable DE
dv dx
d (In x)
_ =
=
Fv) —v ~
cot ycot@ +1
-—=
=
coty =
d0 cot 6 — coty
1dr 1 + tan
y tan 0 _ 1 + (tan 6)F(tan6)
(17) = Q0)
rd tan y —tan 6 F(tan 6) — tan
0
7 Linear Fractional Equations 19
Invariant Radii. The radii along which the denominator of Q(@) vanishes
are those where (16) is equivalent to d0/dr = 0. Hence, these radii are particular
solution curves of (16); they are called invariant radii. They are the solutions
y = 7x, for constant 7 tan 6. Therefore, they are the radii y = rx for which
=
=
y = 7 = (¢ + dr)/(a + br), by (16), and so their slopes 7 are the roots of the
quadratic equation
Ifb # 0, Eq. (18) has zero, one, or two real roots according as its discriminant
is negative, zero, or positive. This discriminant is
In the sectors between adjacent invariant radii, d?/dr has constant sign; this fact
facilitates the sketching of solution curves. Together with the invariant radii, the
solution curves (17’) form a regular curve family in the punctured plane, consist-
ing of the xy-plane with the origin deleted.
similarity transformation of the xy-plane for any fixed k, it follows that the solu-
tion curves in the sector between any two adjacent invariant radii are all geo-
metrically similar (and similarly placed). This fact is apparent in the drawings of
Figure 1.6.
Note also that the hyperbolas in Figure 1.6a@ are the orthogonal trajectories
of those of Figure 1.5. This is because they are integral curves of yy’ x and
=
=
EXERCISES D
1. Sketch the integral curves of the DEs in Exs. C8 and C9 in the neighborhood of the
origin of coordinates.
(cx + dy + e) dx — (ax + by + f) dy = 0 ad
# be
Prove in detail that the solutions of any homogeneous DE y’ g(y/x) have the
Similarity Property described in §7
Show that the solution curves of y’ G(x,y) cut those of y’ F(x,y) at a constant
angle @ if and only if G (1 + F)/(1 — tF), where
r = tan
B
(a) Show that the differential (ax + by) dy — (cx + dy) dx is exact if and only if a +
d = 0, and that in this case the integral curves form a family of coaxial conics
(b) Using Exs. 6 and 7, show that if tan 6 (a + d)/(c — 5), the curves cutting the
solution curves of the linear fractional DE y (cx + dy)/(ax + by) at an angle
B form a family of coaxial conics
at nearby points, weighting the nearest points most heavily (i.e., using graphical
interpolation). Methods of doing this systematically are called schemes of graph-
ical integration.
The preceding construction also gives a graphical representation of the direc-
tion field associated with a given normal first-order DE. This is defined as
follows.
results with roughly the same computational effort by replacing (19) with the
following “improved” Euler algorithm
With h = .001, this “improved” Euler method gives 5-digit accuracy in most
cases, while requiring only about twice as much arithmetic per time step.
Whereas with Euler’s method, to use 10 times as many mesh points ordinarily
gives only one more digit of accuracy, the same mesh refinement typically gives
two more digits of accuracy with the improved Euler method.
As will be explained in Chapter 8, when truly accurate results are wanted, it
is better to use other, more sophisticated methods that give four additional digits
of accuracy each time h is divided by 10. In the special case of quadrature—that
is, of DEs of the form y’ = g(x) (see §2)—to do this is simple. It suffices to
replace (19) by Simpson’s Rule.
Xn h
(21) Ynt+1= Yn + 7 [g(x,) + se + g(x, + h))
2
25
1 50 50 50
In 2 = —
150 & | 48 + 2k
+
49 + 2k
+
50 + 2k |
Caution. To achieve 8-digit accuracy in summing 25 terms, one must use a
computer arithmetic having at least 9-digit accuracy. Many computers have only
7-digit accuracy!
a
For the DE y’ = y, since y;, = 9a ~ Yn» this method gives Y,,, = (1 +h + h?/
2)¥,,, and so it is equivalent to the improved Euler method.
For the DE y’ = 1 + 9°, since y” = 2Qyy’ = 2y(1 + 9°), the method gives
This differs from the result given by Euler’s improved method. In general, since
d[F(x,y)|/dx = OF/dx + (OF/dy) dy/dx, Y% =
=
(F, + FF,),. This makes the
method easy to apply.
The error per step, like that of the improved Euler method, is roughly pro-
portional to the cube of h. Since the number of steps is proportional to h', the
cumulative error of both methods is roughly proportional to h”. Thus, one can
obtain two more digits of accuracy with it by using 10 times as many mesh
points.
As will be explained in Chapter 8, when truly accurate results are wanted,
one should use other, more sophisticated methods that give four additional digits
of accuracy when 10 times as many mesh points are used.
e
eth Ax
tan
x + tanh
e = é¢@ and tan (x + h) =
1 — tanx tank’
EXERCISES E
1. For each of the following initial value problems, make a table of the approximate
numerical solution computed by the Euler method, over the interval and for the
mesh lengths specified:
(a) y = y with 9(0) = 1, on [0,1], for h = 0.1 and 0.02.
(b) y’ = 1 + 9’ with 9(0) = 0, on [0,1.6], for A = 0.1, 0.05, and 0.02.
Knowing that the exact solutions of the preceding initial value problems are e” and
tan x:
+ See for example Abramowitz and Stegun, which contains also a wealth of relevant material.
24 CHAPTER 1 First-Order Differential Equations
Find the errors of the approximate values computed in Exercise 3, and analyze the
ratios Y,/h*x (cf. Ex. 2).
Use Simpson’s Rule to compute a table of approximate values of the natural loga-
rithm function In x
=-
=
f dt/t, on the interval [1,2].
“7 In selected cases, test how well your tables agree with the identities arctan (tan x) =
x and In () =x.
*8 Let ¢, be the approximate value of e obtained using Euler’s method to solve y’ = y
for the initial condition y(0) = 1 on [0,1], on a uniform mesh with mesh length 4 =
L/n.
(a) Show that Ine, = nIn (1 + A).
(b) Infer that ne, = 1 —h/2+ h?/3—---
(c) From this, derive the formula
For any normal first-order differential equation y’ = F(x,y) and any “‘initial”
Xo (think of x as time), the initial value problem consists in finding the solution
or solutions of the DE, for x = x», which also satisfy f(x9) = c. In geometric
language, this amounts to finding the solution curve or curves that issue from
the point (xo,c) to the right in the (x,y)-plane. As we have just seen, most initial
value problems are easy to solve on modern computers, if one is satisfied with
approximate solutions accurate to (say) 3—5 decimal digits.
However, there is also a basic theoretical problem of proving the uniqueness of
this solution.
When F(x,y) = g(x) depends on x alone, this theoretical problem is solved by
the Fundamental Theorem of the Calculus (§2). Given xp =>
=
aand yy = c, the
initial value problem for the DE y’ = g(x) has one and only one solution, given
by the definite integral (5).
The initial value problem is said to be well-posed in a domain D when there is
one and only one solution y = f(x,c) in D of the given DE for each given (X9,c)
€ D, and when this solution varies continuously with c. To show that the initial
value problem is well-posed, therefore, requires proving theorems of existence
(there is a solution), uniqueness (there is only one solution), and continuity (the
solution depends continuously on the initial value). The concept of a well-posed
initial value problem gives a precise mathematical interpretation of the physical
9 The Initial Value Problem 25
concept of determinism (cf. Ch. 6, §5). As was pointed out by Hadamard, solu-
tions which do not have the properties specified are useless physically because
no physical measurement is exact.
It is fairly easy to show that the initial value problems discussed so far are
well-posed. Thus, using formula (8’), one can show that the initial value problem
is well-posed for the linear DE y’ + p(x)y = q(x) in any vertical strip a < x <b
where p and q are continuous. The initial value problem is also well-posed for
the linear fractional DE (16) in each of the half-planes ax + by > 0 and
ax + by
< 0.
Actually, for the initial value problem for y’ = F(x,y) to be well-posed in a
domain D, it is sufficient that F € @! in D. But it is not sufficient that F € @:
though the continuity of F implies the existence of at least one solution through
every point (cf. Ch. 6,813), it does not necessarily imply uniqueness, as the fol-
lowing example shows.
(x — a)’, x<a
(22) J,=
ax=x=8
— x > Bp
The rest of this chapter will discuss existence, uniqueness, and continuity
theorems for initial value problems concerning normal first-order DEs
y’ = F(x,y). Readers who are primarily interested in applications are advised to
skip to Chapter 2.
Example 9 shows that the mere continuity of F(x,y) does not suffice to ensure
the uniqueness of solutions y = f(x) of y’ = F(x,y) with given f(a) = c. However,
it is sufficient that F € @'.(D). We shall prove this and continuity at the same
time, using for much of the proof the following generalization of the standard
Lipschitz condition.
for all point pairs (x,y) and (x,z) in D having the same x-coordinate.
The same function F may satisfy Lipschitz conditions with different Lipschitz
constants, or no Lipschitz conditions at all, as the domain D under consideration
varies. For example, the function F(x,y) = 3y2/3 of the DE in Example9satisfies
a Lipschitz condition in any half-plane y = ¢, ¢ > 0, with L = 2e~'/?, but no
Lipschitz condition in the half-plane y > 0. More generally, one can prove the
following.
+R. Lipschitz, Bull. Sci. Math. 10 (1876), p. 149; the idea of the proof is due to Cauchy (1839). See
Ince, p. 76, for a historical discussion.
t Aset of points is called convex when it contains, with any two points, the line segment joining them.
10 Uniqueness and Continuity 27
Proof. The domain being convex, it contains the entire vertical segment join-
ing (x,y) with (x,z). Applying the Law of the Mean to F(x,y) on this segment,
considered as a function of 7, we have
1 |e
| F(x,y) — F(x,z)| = ly
for some n between y and z. The inequality (23), with L = supp|0F/dy|, follows.
A similar argument shows that (23) holds with L = max p 0F/dy.
”
The case F(x,y) = g(x) of ordinary integration, or “quadrature, is easily
identified as the case when L =0 in (23’). A Lipschitz condition is satisfied even
if g(x) is discontinuous.
from the DE. If yy > y), then, by (23), the right side of this equation has the
upper bound L(y — ,)®. Since all expressions are unaltered when y, and yo are
interchanged, we see that the inequality of Lemma3is true in any case.
We now prove that solutions of y’ = F(x,y) depend continuously (and hence
uniquely) on their initial values, provided that a one-sided Lipschitz condition
holds.
28 CHAPTER 1 First-Order Differential Equations
THEOREM 5. Let f(x)and g(x) be any two solutions of the first-order normal DE
y = F(x,y)in a domain D where F satisfies the one-sided Lipschitz condition (23). Then
By Lemma 3, this implies that o’(x) =< 2Lo(x); and by Lemma 2, this implies
o(x) < e*+*-%9(a). Taking the square root of both sides of this inequality (which
are nonnegative), we get (25), completing the proof.
As the special case f(a) = g(a) of Theorem 5, we get uniqueness for the initial
value problem: in any domain whereFsatisfies the one-sided Lipschitz condition
(23), at most one solution of y’ = F(x,y) for x = a, satisfies f(a) = c. However,
we do not get uniqueness or continuity for decreasing x. We now prove that we
have uniqueness and continuity in both directions when the Lipschitz condition
(23’) holds.
In particular, the DE y’ = F(x,9) has at most one solution curve passing through any
point (a,c) €D.
Proof. Since (23’) implies (23), we know that the inequality (23) holds; from
Theorem 5, this gives (26) for x = a. Since (23’) also implies (23) when x goes
to —x, we also have by Theorem 5
EXERCISES F
Let a(t), positive and of class @! for a = t = a + ¢, satisfy the differential inequality
o'(t) = Ko(t) log o(#). Show that o(t) = o(@) exp [K(é — a)].
Let F(x,y) = y log (1/y) for 0 < y < 1, F(y) = 0 fory = 0. Show that y’ = F(x,y)
has at most one solution satisfying f(0) = c, even though F does not satisfy a
Lipschitz condition.
(Peano uniqueness theorem). For each fixed x, let F(x,y) be a nonincreasing function
of y. Show that, if f(x) and g(x) are two solutions of y’ = F(x,y), and 6 > a, then
if® — g®| = lf@ — g(@|. Infer a uniqueness theorem.
identically on the strip 0 < x < a. Show that, if the improper integral [§ k(x) dx is
finite, then y’ = F(x,y) has at most one solution satisfying (0) = 0.
with exact solutions of the DE (3). The following theorem gives such a
comparison.
Proof. Suppose that f(x) > g(x,) for some x, in the given interval, and define
Xo to be the largest x in the interval a < x = x, such that f(x) < g(x). Then
30 CHAPTER 1 First-Order Differential Equations
F(%o) = g(xo). Letting o(x) = f(x) — g(x), we have o(x) = 0 for x9 S x S x); and,
also for x9 = x = x,
where L is the Lipschitz constant for the function F. That is, the function o
satisfies the hypothesis of Lemma 2 of §10 on xy S x S x, with K = L. Hence
a(x) < o(xp)e“*-* = 0 and so o, being nonnegative, vanishes identically. But
this contradicts the hypothesis f(x,) > g(x,). We conclude that f(x) =< g(x) for
all x in the given interval, q.e.d.
Theorem 6, applied to the functions v, u and A(u,v) = —F(x, —v) now yields
the inequality u(x) = u(x) for x = a, or g(x) = f(x), as asserted.
The inequality f(x) < g(x) in this Comparison Theorem can often be replaced
bya strict inequality. Either fand g are identically equal for a = x = x), or else
f(%o) < g(%o) for some xp in the interval (a, x,). By the Comparison Theorem,
the function o,(x) = g(x) — f(x) is nonnegative for a = x < x), and moreover
61(x9) > 0. Much as in the preceding proof
Hence [e'*o,(x)]’ =
=
e*[o{ + Lo,] = 0; from this expression ¢’c,(x) is a non-
decreasing function on a = x = x,. Consequently, we have
COROLLARY 1. In Theorem 6, for any x, > a, either f(x,) < g(x), or f(x) =
g(x) for all x € [a,x;].
Proof. The proof will be by contradiction. 1f we had f(x) = g(x) for some
x > a, there would be a first x = x, > a where f(x) = g(x). The two functions
y = d(x) = f(—x) and z = (x) = g(—x) satisfy the DEs y’ = —F(—x,y) and
z’ = —G(-—x,z) as well as the respective initial conditions ¢(—x,) = ¥(—+x)).
Since —F(—x,y) = —G(—x,y), we can apply Theorem 7 in the interval
[—x,, —a], knowing that the function — F(—x,y) satisfies a Lipschitz condition.
We conclude that ¢(—a) = ¥(—a), that is, that f(a) = g(a), a contradiction.
In this chapter, we have analyzed many methods for solving first-order DEs
of the related forms y’ = F(x,y),M(x,y) + N(x,y)y’ = 0, and M(x,y) dx + N(x,y)
dy = 0, describing conditions under which their “solution curves” and/or “‘inte-
gral curves’’ constitute “one-parameter families’ filling up appropriate domains
of the (x,y)-plane. In this concluding section, we will try to clarify further the
relationship between such first-order DEs and one-parameter curve families.
A key role is played by the Implicit Function Theorem, which shows} that the
level curves u = C of any function u € @'(D) have the following properties in
any domain D not containing any critical point: (i) one and only one curve of
the family passes through each point of D, (ii) each curve of the family has a
tangent at every point, and (iii) the tangent direction is a continuous function
of position. Thus, they constitute a regular curve family in the sense of the fol-
lowing definition.
+ Where du/dy = 0 but du/dx # 0, we can set x = g(y) locally on the curve; see below.
32 CHAPTER 1 First-Order Differential Equations
ficult, and will be deferred to Chapter 6. There we will establish the simpler
result that the initial value problem is locally well-posed for such DEs, after treat-
ing (in Chapter 4) the case that F is analytic (i.e., the sum of a convergent power
series).
In the remaining paragraphs of this chapter, we will simply try to clarify fur-
ther what the Implicit Function theorem does and does not assert about “‘level
curves,”
y=Yotn
x &
u>
= WO
ae
2 UO
u>
¥Y=Yo7-7
xX = Xg—€ X=Xote
Figure 1.8
12 Regular and Normal Curve Families 33
EXERCISES G
Show that if g(y) satisfies a Lipschitz condition, the solutions of y’ = g(y) form a
normal curve family in the (x,y)-plane. [HINT: Apply the Inverse Function Theorem
tox = f dy/g(y) + C]
Let g(x) be continuous for 0 = x < ©, lim,.... g(x) = 5 and a > 0. Show that, for
every solution y = f(x) of y’ + ay = g(x), we have lim, f(x) = b/a.
Show that if a < 0 in Ex. 5, then there exists one and only one solution of the DE
such that lim,. f(x) = b/a.
Let F, G, f, g be as in Theorem 8, and F(x,y) < G(x,y). Show that f(x) < g(x) for x
> a, without assuming that F orG satisfies a Lipschitz condition.
Show that the conditions dx/dt = |x| 12 and x(0) = —1 define a well-posed initial
value problem on [0,a) if a S 1, but not ifa > 1.
*11 (a) Prove that there is no real-valued function u € @! in the punctured plane of
Ex. 10 whose level curves are the integral curves of xy’ = y.
(b) Show that the integral curves of y’ = (x + y)/(x — 9) are the equiangular spirals
r= he =e
2 #0,
(b) Prove that there is no real-valued function u € @! whose level curves are these
spirals.
CHAPTER 2
SECOND-ORDER
LINEAR
EQUATIONS
1 BASES OF SOLUTIONS
du
(1) po(x)
—
dx”
+ pix)
“% + polx)u = ps(x)
The coefficient-functions p,(x) [i = 0, 1, 2, 3] are assumed continuous and real-
valued on an interval J of the real axis, which may be finite or infinite. The inter-
val J may include one or both of its endpoints, or neither of them. The central
problem is to find and describe the unknown functions u = f(x) on J satisfying
this equation, the solutions of the DE. The present chapter will be devoted to
second-order linear DEs and the behavior of their solutions.
Dividing (1) through by the leading coefficient po(x), one obtains the normal
form
du
(1’) dx? + p(x) = + q(x)u = r(x)
p=
pi q=
Ps r=
Ps
Po Po Po
(*) 2la-s%]+m=0
has singular points at x = +1. This is evident since when rewritten in the form
(1) it becomes (1 — x”)u” = Qxu’ + du —
=
0. Although it has polynomial solu-
34
1 Bases of Solutions 35
tions when A = n(n + 1), as we shall see in Ch. 4, §1, all its other nontrivial
solutions have a singularity at either x =
=
lorx
=
=
—1.
Likewise, the Bessel DE
has a singular point at x = 0, and nowhere else. More commonly written in the
normal form
wt ta + 1-3 Jano
its important Bessel function solution Jo(x) will be discussed in Ch. 4, §8.
Linear DEs of the form (1) or (1’) are called homogeneous when their right-
hand sides are zero, so that ps(x) = 0 in (1)—or, equivalently, r(x) = 0 in (14.
The homogeneous linear DE
obtained by dropping the forcing term p,(x) from a given inhomogeneous linear
DE (1) is called the reduced equation of (1). Evidently, the normal form of the
reduced equation (2) of (1) is the reduced equation
du
(2’) dx? + p(x)
as
% + q(x)u = 0
for any nontrivial solution of u” + k°u = 0. The constantA in (2) is called the
amplitude of the solution; ¥ its initial phase, and k its wave number; k/2z is called
its frequency, and 22/k its period.
(5’) o” + (q — pr/A)u =
=
0, v= ey
Case 1. IfA > 0, then (5) reduces to v” = k?v, where k = VA/2. This DE
has the functions v = e*, ¢— hx as a basis of solutions whence
(VE—p)x/2 (-VA-p)x/2
(6a) u=e u=e
are a basis of solutions of (5). Actually, it is even simpler to make the “exponen-
tial substitution” «
—
=
* in this case. Then (5) is equivalent to (4? + prat
ge = 0; the roots of the quadratic equation \? + pr + q = 0 are the coeffi-
cients of the exponents in (6a).
With differential equations arising from physical problems, one is often inter-
ested in particular solutions satisfying additional initial or boundary conditions.
Thus, in Example 1, one may wish to find a solution satisfying u(0) = up and
u’(0) = ug. An easy way to find a solution satisfying these initial conditions is to
use Eq. (4) with a = uy and b = uj/k. In general, given a second-order linear
DE such as (1) or (14), the problem of finding a solution u(x) that satisfies given
initial conditions u(a) = up and u’(a) = up is called the initial value problem.
satisfies the original DE u” + u = 3 sin 2x. Such a function will satisfy u(0) =
0 if and only if a = 0, so that
u’(x) = bcos
x — 2 cos 2x
uaixt qd —2pe
q
—~—
38 CHAPTER 2 Second-Order Linear Equations
Finally, u”
=
=
¥(0)
(0) | = $(0)¥/(0)—¥(0)6(0) ¥ 0
¢/(0) ¥(0)
EXERCISES A
Solve the initial value problem for each of the DEs of Exercise 4, with the initial con-
ditions u(0) = u’(0) = 4.
Solve the initial value problem for each of the DEs of Exercise 7, and the initial con-
ditions u(0) = u’(0) = 0.
Show that any second-order linear homogeneous DE satisfied by x sin x must have a
singular point at x = 0.
3 Qualitative Behavior; Stability 39
Note that when A < 0 (i.e., in Case 2), all nontrivial solutions of (5) reverse
sign each time that x increases by +/k. Qualitatively speaking, they are oscillatory
in the sense of changing sign infinitely often. These facts become evident if we
rewrite (6b) in the form (4), as
THEOREM 1. ifA 2 0, then a nontrivial solution of (5) can vanish at most once.
If A < 0, however, it vanishes periodically with period x/\/ —&
Proof. This result can be proved very simply if complex exponents are used
freely (see Chapter 3, §3). In the real domain, however, one must distinguish
several possibilities, viz.:
(A) 1f q < 0, then A > 0 and A? + pr + gq = 0 must have two real roots of
opposite sign. Instability is therefore obvious.
(B) If p < 0, instability is obvious from (6a)-(6b), if one keeps in mind the sign
of p in each case.
(C) If p = 0 and q > 0, then we have Example 2: the DE (5) is stable but not
strictly stable.
(D) If p > 0 and q > 0, there are two possibilities: (i) A S 0, in which case we
have strict stability by (6b) and (6c); (ii) A > 0, in which case VA< p since
A = p? — q < p’, and strict stability follows from (6a).
q=up/4 _
q=p/4
Unstable Stable
focal point focal point
—_
Ay way hie;
eo
Unstable Stable
nodal point nodal point
w<M <0 O<m<d
4 i
Saddlepoint (unstable)
M<0<ds
4 UNIQUENESS THEOREM
We are now ready to treat rigorously the initial value problem stated in §1.
The first basic concept involved is very general and applies to any normal sec-
ond-order DE u” = F(x,u,u’), whether linear or not.
Think of x as time, and of the possible pairs (u,w’) as states of a physical system,
which is governed (or modeled mathematically) by the given DE. Since wu’
expresses the rate of change of u at any “time” x, while u” = du’/dx gives the
rate of change of w’, it is natural to surmise that the present state of any such
system uniquely determines its state at all future times. Indeed, the theoretical
initial value problem is to prove this result as generally as possible.
In this section, we will prove it for second-order linear DEs of the form (1)
having continuous coefficient-functions p,(x) and no singular points. Since fo(x)
¥ 0, it suffices to consider the normal form (1’).
One would like to prove also that there always exists a solution for any initial
(u,45); this will be proved for second-order linear DEs having analytic coefh-
cient-functions in Chapter 4, and (locally) for linear DEs having continuously
differentiable coefficient-functions in Chapter 6. For the present, we will have to
construct “particular” solutions and bases of solutions for homogeneous DEs
u” + p(x)u’ + q(x)u = 0 by other methods.
= pof”
+ pif’ + bof
4 Uniqueness Theorem 41
(for continuous p,) is a transformation from one family of functions [in our case,
the family @?(J) of continuously twice-differentiable functions on a given inter-
val J] to another family of functions [in our case, @(J)]. Such a functional trans-
formation is called an operator, and is written in operator notation
The proof is trivial, but the result describes the fundamental property of lin-
ear operators. Its use greatly simplifies the solution of inhomogeneous linear
DEs.
Proof. Let v and w be any two solutions of (1’) that satisfy these initial con-
ditions; we shall show that their differences u —
=
v — w vanishes identically.
+ The phrase “particular solution”’ is used to emphasize that only one solution of (1) need be found,
thus reducing the problem of solving it to the case ps(x) = 0.
42 CHAPTER 2 Second-Order Linear Equations
—2p(x)u + 21 — g(x))uw’
and
By Lemma 2 of Ch. 1, §10, it follows that o(x) = 0 for all x € [a, 6]. Hence
u(x) = 0 and v(x) = w(x) on the interval, as claimed.
For some x Xo, let (flxo), f’(%o)) and (g(xo), g’(Xo)) be linearly independent
_
In other words, the general solution of the given homogeneous DE (8) is cf(x)
+ dg(x), where c and d are arbitrary constants.
Proof. By the Superposition Principle, any such h(x) satisfies (8). Conversely,
suppose the function h(x) satisfies the given DE (8). Then, at the given point xg,
constants c and d can be found such that
where we have used the abbreviations fo = fixo), f6 = f’(%o), and so on. For this
choice of c and d, the function
5 THE WRONSKIAN
fx) f'®)
(9) Wf g *) = flxe’(x) — gf") = g(x)
g(x)
THEOREM 5. The Wronskian (9) of any two solutions of (8) satisfies the identity
Equation (10) follows from the first-order homogeneous linear DE (11) by Theo-
rem 4 of Ch. 1, §6.
44 CHAPTER 2 Second-Order Linear Equations
We now relate the Wronskian of two functions to the concept of linear inde-
pendence. In general, a collection of functions f|, fo... , f, is called linearly
independent on the interval a < x < 6 when no linear combination ¢, f\(x) +
Cofglx) +... + Cyfy (x) of these functions gives the identically zero function for
a = x = |, except the trivial linear combination where all coefficients vanish.
Functions that are not linearly independent are called linearly dependent. 1f f
and g are any two linearly dependent functions, then cf + dg = 0 for suitable
constants c and d, not both zero. Hence g = —(d/d)f or f = —(c/dg; the func-
tions f and g are proportional.
LEMMA. [If f and g are linearly dependent differentiable functions, then their
Wronskian vanishes identically.
Proof. Suppose that fand g are linearly dependent. Then there are two con-
stants ¢ and d, not both zero, which satisfy the two linear equations
The interesting fact is that when fand g are both solutions of a second-order
linear DE, a strong converse of this lemma is also true.
Proof. Suppose that the Wronskian W(f, g; x) vanished at some point x).
Then the vectors [ f(x,), f’(x,)] and [g(x)), g’(x))] would be linearly dependent
and, therefore, proportional: g(x,) = kf(x;) and g’(x,;) = kf’(x,) for some con-
stant k. Consider now the function h(x) = g(x) — kf(x). This function is a solu-
tion of the DE (8), since it is a linear combination of solutions. It also satisfies
the initial conditions h(x) = h’(x,) = 0. By the Uniqueness Theorem, this func-
tion must vanish identically. Therefore, g(x) = &f(x) for all x, contradicting the
hypothesis of linear independence of f and g.
f? f?
5 The Wronskian 45
This suggests that the ratio of two functions is a constant if and only if their
Wronskian vanishes identically. However, this need not be true if f vanishes: the
ratio of the two functions x° and |x|° is not a constant, yet their Wronskian
Wx3, |x|°) = 0. (Note also that both functions satisfy the DEs xu’ = 3u and
8xu” — Qu’ = Q.)
Nevertheless, the connection between W( f,g) and g/f is a useful one. Thus,
it allows one to construct a second solution g(x) of (8) if one nontrivial solution
is known. Namely, if P(x) = JSp(x)dx is any indefinite integral of p(x), then the
function
e7P®)
(12) a = fof | *(x)
|e
is a second, linearly independent solution of (8) in any interval where f(x) is
nonvanishing. This is evident, since (g/f)’ = W(fig)/f?, whence
— P(x)
Ja=JIf?@) |
gh)
_ W( fg)
fix) S| f?
6x
, 72
Uu
(
v’
(13)
=- —
p(x)v — q(x) — v?
=
_—
=
Uu
The quadratic first-order DE (13) is called the Riccati equation associated with
(8); its solutions form a one-parameter family. Conversely, if v(x) is any solution
of the Riccati equation (13) and if u’ = vu(x)u, then u satisfies (8). Hence, every
solution u(x) of (8) can be written in any interval where u does not vanish, in
the form,
EXERCISES B
1. Show that all solutions of (8) have continuous second derivatives. Show also that
this is not true for (1).
2. Find a formula expressing the fourth derivative u” of any solution u of (8) in terms
of u, u’, and the derivatives of p and g. What differentiability conditions must be
assumed on the coefficients of (8) to justify this formula?
For the solution pairs of the DEs specified in Exs. 3-5 to follow, (a) calculate the
Wronskian, and (b) solve the initial-value problem for the DE specified with each of the
initial conditions u(0) = 2, u’(0) = 1, and u(0) = 1, u’(0) = —1 (or explain why there is
no solution).
Let f(x), g(x), and h(x) be any three solutions of (8). Show that
ff f"
gg’ Bg" =0
hh”
Ah
10 (a) Show that if p,q € @", then every solution of (8) is of class @**?.
(b) Show that if every solution of (8) is of class @"*? then pe @" and g€e@".
11 Let f(x), g(x), h(x) be three solutions of the linear third-order DE
ff f*
w(x) =
ge 8”
h”
AW
6 Separation and Comparison Theorems 4
*12. Let y” + q(x)y = 0, where g(x) is “piecewise continuous” (i.e., continuous except
for a finite number of finite jumps). Define a “solution” of such a DE as a function
y = fix) € @' that satisfies the DE except at these jumps.
(a) Show that any such solution has left- and right-derivatives at every point of
discontinuity.
(b) Describe explicitly a basis of solutions for the DE y” + g(x)y = 0, if
+1 when x > 0
q(x) =
—l when x < 0
The Wronskian can also be used to derive properties of the graphs of solu-
tions of the DE (8). The following result, the celebrated Sturm Separation Theo-
rem, states that all nontrivial solutions of (8) have essentially the same number
of oscillations, or zeros. (A ‘‘zero”’ of a function is a point where its value is zero;
functions have two zeros in each complete oscillation.)
THEOREM 7. [Jff(x) and g(x) are linearly independent solutions of the DE (8),
then f(x) must vanish at one point between any two successive zeros of g(x). In other
words, the zeros offix) and g(x) occur alternately.
since f and g are linearly independent; hence, f(x,) # 0 and g’(x,) # 0 if g(x,)
= 0. If x, and x» are two successive zeros of g(x), then g’(x,), g’(x2), f(x,), and
Jix2) are all nonzero. Moreover, the nonzero numbers g/’(x,) and g’(xs) cannot
have the same sign, because if the function is increasing at x = x,, then it must
be decreasing at x Xg, and vice-versa. Since W(/, g; x) has constant sign by
=
=
the Corollary of Theorem 4, it follows that f(x,) and f(x.) must also have opposite
signs. Therefore f(x) must vanish somewhere between x, and x9.
THEOREM 8. Let f(x) and g(x) be nontrivial solutions of the DEs u” + p(x)u = 0
and v” + q(x)v = 0, respectively, where p(x) = q(x). Then f(x) vanishes at least once
between any two zeros of g(x), unless p(x) = q(x) and fis a constant multiple of g.
48 CHAPTER 2 Second-Order Linear Equations
Proof. Let x, and x» be two successive zeros of g(x), so that g(x,) = g(xo) =
0. Suppose that f(x) failed to vanish in x; < x < x9. Replacing f and/or g by
their negative, if necessary, we could find solutions f and g positive on x, < x
< x9. This would make
p-q=Wf gx) =0
whose solutions vanish when u does (for x # 0). Applying the Comparison Theo-
rem to (15) and u” + u = 0, we obtain the following.
0 depend on
the sign of q(x) is illustrated by Figures 2.2 and 2.3, which depict sample solution
curves for the cases q(x) = 1 and q(x) = —1, respectively.
7 The Phase Plane 49
y |
7S
RESKRRS
ERO DM \i
i
SEEK
VEY x
Figure 2.2 Solution curves of u” + u = 0
4ex
4cosh
z
4 sinh
z
sez
rT
—}te-z
— sinh x
z
— 4teosh
det
The trajectories just defined have some important general geometrical prop-
erties. For example, since u is increasing when wv’ > 0 and decreasing when w’
< 0, the paths of solutions must go to the right in the upper half-plane and to
the left in the lower half-plane. Furthermore, paths of solutions (‘‘trajectories’’)
must cut the u-axis u’ = 0 orthogonally, except where F = 0.
We will treat in this chapter only homogeneous second-order linear DEs (8),
deferring discussion of the nonlinear case to Chapter 5. Using the letter v to
signifly wu’, this DE is obviously equivalent to the system
du d
(16) —_ =
’
—_—_—
=
P(x)o — q(x)u
dx d.
d 0 1
(3) =|
u
dx — q(x) — p(x) I v
Note that if g(x) < 0, then du’/dx = —q(x)u has the same sign as u on the u-
axis. It follows that if q(x) is negative, then any trajectory once trapped in the
first quadrant can never leave it, because it can neither cross the u-axis into the
fourth quadrant nor recross the u’-axis into the second quadrant. The same is
true, for similar reasons, of trajectories trapped in the third quadrant.
Even more important, two nontrivial solutions of (8) are linearly dependent if
and only if they are on the same straight line through the origin in the (u,v)-
plane. It follows that each straight line through the origin moves as a unit. The
preceding facts also become evident analytically, if we introduce clockwise polar
coordinates in the phase plane, by the formulas
—(csc? 0)6’ =
=
—pcot 6 — q — cot?6
TY = =
=
=
r® cos O[(1 — g(x)) sin 6 — p(x) cos 6]
7 The Phase Plane 51
1 dr
(19)
—_—— =
From (20a) it follows that, after the graph of any @(x) has crossed the line 6 =
nw, it can never recross it backwards. Where u(x) next vanishes (if it does), we
must have 6 = (n + 1)z; in other words, successive zeros of u(x) occur precisely
where 6 increases from one integral multiple of 2 to the next!
After verifying that the right side of (18) satisfies a Lipschitz condition, we
see that this inequality can never cease to hold; hence, in any interval where
6,(x) increases from nz to nw + 7, 6.(x) must cross the line 6 = na + 7 and so
ug must vanish there. Sturm’s Comparison Theorem follows similarly: if q(x) is
increased and p(x) is left constant, the Comparison Theorem of Ch. 1, applied
to (19), yields it as a corollary.
Focal, Nodal, and Saddle Points. Even more interesting than Sturm’s theo-
rems are the qualitative differences between the behavior of solutions of differ-
ent second-order DEs that become apparent when we look at the corresponding
trajectories in the phase plane (their so-called phase portraits). We shall discuss
these for nonlinear DEs in Chapter 5; here we shall discuss only the linear, con-
stant-coefficient case. We have already discussed this case briefly in §§2~3, pri-
marily from an algebraic standpoint.
In the linear constant-coefficient case, using the letter v to signify u’, we obvi-
ously have
du du
(21) —_ = —-
=
i
po
— qu.
dx Ya
for nonlinear DEs of the general form oe F(u,v), of which the form
d
dv _ ~po
|
(21’)
du v
A=2
/
0
ss
/ /
(a) u" +0.2u' + 4.0lu = 0. (b) 2u"—5u'+ 2u=0
When > 4q (i.e., in Cases B and C), the two lines u’ = au and u’ = Bu in
the phase plane are invariant lines (Ch. 1, §'7). These lines, which correspond to
the solutions e** and e**, divide the uu’-plane up into four sectors, in each of
which @& is of constant sign and so @ is monotonic. If ¢ = a6 > 0, the two invar-
iant lines lie in the same quadrant; if g < 0, they lie in adjacent quadrants.
Case B. In this case, the trajectories in each sector are all tangent at the
origin to the same invariant line, and have an asymptotic direction parallel to
the other invariant line at 00. Fig. 2.4b depicts the phase portrait for 2u” — 5u’
+ Qu 0. The lines v 2u and u 2v are the invariant lines of the corre-
= = —
= => =
sponding linear fractional DF, dv/du = (5v — u)/v. In Case B, the origin is said
to be a nodal point.
Case C. In the saddle point case that p® > 0 > 4g, the two invariant lines lie
in different quadrants, and all trajectories are asymptotic to one of them as they
come in from infinity, and to the other as they recede to it. Figure 1.5 depicts
the phase portrait for the case u” u, with hyperbolic trajectories u 2 2
= —
—vd
=
EXERCISES C
1. (a) Show that if g(x) = f’(x), then g(x) vanishes at least once between any two zeros
of f(x).
(b) Show how to construct, for any n, a function f(x) satisfying (0) = fl) = 0,
Six) # 0 on (0,1), yet for which f(x) vanishes n times on (0,1).
54 CHAPTER 2 Second-Order Linear Equations
Show that there is a zero of J,(x) between any two successive zeros of Jo(x).
for all functions u € @®. An integrating factor for the DE (22) is a function v(x)
such that vL[u] is exact. [Here and later, it will be assumed that py € @ and that
Pi-Po € @' in discussing the DEs (22) and (22’).]
The DEs (23) and (23’) can be solved by a quadrature (Ch. 1, §6). Hence, if an
integrating factor of (22) can be found, we can reduce the solution L[u] = r(x)
to a sequence of quadratures.
Evidently, L[u] = 0 is exact in (22) if and only if pp = A, p) = A’ + B, and
po = B’. Hence (22) is exact if and only if
pr = BY = (hy — AY = Bh ~ (pay
LEMMA. The DE (22) is exact if and only if its coefficient functions satisfy
p> — pi + po = 0
DEFINITION. The operator M in (24) is called the adjoint of the linear oper-
ator L. The DE (24), expanded to the DE
Since wu” — uw’ = (wu’ — uw’ and (uw)’ = uw’ + wu’, this can be simplified
to give the Lagrange identity
(26)
(xu’ + [x — (n?/x)]u =0
EXERCISES D
(pu’y + q(x)u = 0
y” + Ty = 0, _P
I(x) =q—-
4— p’/2
*5 Show that two DEs of the form (8) can be transformed into each other by a change
of dependent variable of the form y = v(x)u,v # 0, if and only if the function
I(x) = g(x) — p'(x)/4 — p’(x)/2 is the same for both DEs [J (x) is called the invariant
of the DE].
Reduce the self-adjoint DE (pu’)’ + qu = 0 to normal form, and show that, in the
notation of Ex. 5, I(x) = (p? — 2pp” + 4pq)/4p°.
(a) Show that, for the normal form of the Legendre DE [(1 — x*)u’])’ + Au = 0
(1 +A
— Ax’)
I(x) = (Use Ex. 6.)
(1 — x)?
(b) Show that, if \ = n(n + 1), then every solution of the Legendre equation has
at least (2n + 1)/x zeros on (—1, 1).
Let u(x) be a solution of u” = q(x)u, q(x) > 0 such that u(0) and w’(0) are positive.
Show that wu’ and u(x) are increasing for x > 0.
Let h(x) be a nonnegative function of class @'. Show that the change of independent
variablet = f% h(s) ds, u(x) = v(t), changes (8) into v” + p,(é)v’ + ¢,(v = 0, where
Pil) = [plx)a(x) + h'(e)]/h(x)? and q(t) = g(x)/h(x)’.
10 (a) Show that a change of independent variable ¢ + f Iq(x)|'? dx, q # 0,
=
=
du q + 2pq du
(*) dt?
( 2\q|*” dt
~
9 GREEN’S FUNCTIONS
2
u
(27) L{u] =
—
2
u
(27’) Lf{u] = —
by the nonzero function r(x) on the right side. In applications to electrical and
dynamical systems, the function r(x) is called the forcing term or input function.
By the Uniqueness Theorem of §4 and Lemma 2 of §4, it is clear that the solu-
tion u(x) of L[u] = r(x) for given homogeneous initial conditions such as u(0)
u’(0) = 0 depends linearly on the forcing term. We will now determine the
=
=
satisfy given homogeneous boundary conditions, provided that the latter define
a well-set problem.
The kernel G(x, £) of Eq. (28) is then called the Green’s functiont associated
with the given boundary value problem. In operator notation, it is defined by
the identity L[G[r]] = r (Gis a “right-inverse’’ of the linear operator L) and
the given boundary conditions.
Green’s functions can be defined for linear differential operators of any
order, as we will show in Ch. 3, §9. To provide an intuitive basis for this very
general concept, we begin with the simplest, first-order case. In this example,
the independent variable will be denoted by ¢ and should be thought of as rep-
resenting time.
+ To honor the British mathematician George Green (1793-1841), who was the first to use formulas
like (28) to solve boundary value problems. Cauchy and Fourier used similar formulas earlier to solve
DEs in infinite domains.
9 Green’s Functions 59
du
a bet wo
1f the account is opened when ¢ = 0 and initially has no money: u(0) = 0, then
one can calculate u(7) at any later time T > 0 as follows. Each infinitesimal
deposit r(f) dt, made in the time interval (¢, + dé), increases through compound
interest accrued during the time interval from t to T by a factor e7~® to the
amount e?—r(¢) dt. Hence the account should amount, at time T, to the integral
(limit of sums)
0 t< ty
u(t)
-| (vp/v)e™"— sin v(t ~ ty) t=t
Now suppose the mass is given a sequence of small velocity impulses Av, =
=
Proof. We must prove that, for any continuous function r, the definite
integral
+ Kaplan, Advanced Calculus, p. 219. In our applications, dg/dt has, at worst, a simple jump across
t=fT.
9 Green’s Functions 61
Applying this rule twice to the right side of formula (32), we obtain successively,
since Git, t) = 0,
where Wf, g3 7) = fe’) — g(t)f’@ is the Wronskian. This gives for the
Green’s function the formula
COROLLARY. Let f(t) and g(t) be any two linearly independent solutions of the
linear homogeneous DE (27’). Then the solution of L[u] = r(t) for the initial conditions
62 CHAPTER 2 Second-Order Linear Equations
Consequently, if we define the functions o(t) and Y(t) as the following definite
integrals:
‘ {@ * g(r) r(r) dr
ot) =
~~
r(r) dr, vO =~
S_
a Wg) a Wig)
when substituted into (28), will give a solution of the inhomogeneous DE L[u]
= r(t)? Since c(r) and d(r) may be regarded as “variable parameters,”’ which vary
with 7, formula (33) is said to be obtained by the method of variation of
parameters.
EXERCISES E
Show that the general solution of the inhomogeneous DE y” + k’y = R(x) is given by
y = (1/2); sin ke — YR di] + c, sin kx + cy cos kx.
Solve y” + 3y' + 2y = x? for the initial conditions y(0) = y'(0) = (0).
Show that any second-order inhomogeneous linear DE which is satisifed by both x?
and sin” x must have a singular point at the origin.
Construct Green’s functions for the initial-value problem, for the following DEs:
(a) uw” = 0 (b) u” =u () u® +u=0
(d) x7” + (x? + 2x)u’ + (x + 2)u = 0 [HINT: x is a solution.]
Find the general solutions of the following inhomogeneous Euler DEs:
(a) xy — Qxyl + Q2y = x7 + px tq (b) xy” + 3xy’ + y = R(x)
[Hint: Any homogeneous Euler DE has a solution of the form x’.]
8. Show that, if g(f) < 0, then the Green’s function G(t, 7) of u, + q(tu =
=
0 for the
initial-value problem is positive and convex upward for ¢ > r.
aand
x = Bb, are
more natural. For instance, the DE y” =
=
0 characterizes straight lines in
the plane, and one may be interested in determining the straight line joining
two given points (a, c) and (6, d). That is, the problem is to find the solution
y = f(x) of the DE y” = 0 which satisfies the two endpoint conditions f(a) = ¢
and f(b) = d.
Many two-endpoint problems for second-order DEs arise in the calculus of
variations. Here a standard problem is to find, for a given function F(x,y,y’), the
curve y = f(x) which minimizes the integral
d OF
(34’)
(ay’ (oF
=
=
dx dy
For example, if F(x,y,y)) = V1 + y” so that I(f) is the length of the curve, Eq.
(34’) gives zero curvature:
/ 72 a
d
)-9 | aa J
(
dx Viry
J
a + wy?
| = dl + wy?
endpoints a and b > a? When this is so, the resulting two-endpoint problem is
called well-set. Clearly, the two-endpoint problem is always well-set for y” = 0.
Example 5. Now consider, for given p, q, r € C', the curves that minimize the
integral (34) for F = ${p(x)y + 2q(x)yy’ + 1(x)y®]. For this F, the Euler-
Lagrange DE is the second-order linear self-adjoint DE (py’)’ + (7 — ny = 0.
The question of when the two-endpoint problem is well-set in this example is
partially answered by the following result.
with continuous coefficient-functions have two linearly independent solutions.t Then the
two-endpoint problem defined by (35) and the endpoint conditions u(a) c, u(b) = d
=
=
u(b) = 0
=
(36) u(a) =
have one and only one solution vector (a, 8) if and only if the determinant
f@g) — g@f® # 0. The alternative f(a)g(6) = f®)g(@ is, however, the con-
dition that the homogeneous simultaneous linear equations
have a nontrivial solution (a, 8) # (0, 0). This proves Theorem 11.
When the DE (35) has a nontrivial solution satisfying the homogeneous end-
point conditions u(a) = u(b) = 0, the point (, 0) on the x-axis is called a con-
jugate point of the point (a, 0) for the given homogeneous linear DE (35) or for
a variational problem leading to this DE. In general, such conjugate points exist
t In Ch. 6, it will be shown that this hypothesis is unnecessary; a basis of solutions always exists for
continuous p,(x).
11 Green’s Functions, II 65
for DEs whose solutions oscillate but not for those of nonoscillatory type, such
as u” = g(x)u, q(x) > 0.
Thus, in Example 5, letp = 1, ¢ = 0, and r = —k? < Q. Then the general
solution of (35) for the initial condition u(a) = 0 is u = A sin [k(x — a)]. For
u(b) = 0 to be compatible with A # 0, it is necessary and sufficient that b = a
+ (nx/k). The conjugate points of a are spaced periodically. On the other hand,
the DE y” — Ay = 0, corresponding to the choice p = 1, q = 0,7 A in
—
=
We now show that, except in the case that a and b are conjugate points for
the reduced equation L{u] = 0, the inhomogeneous linear DE (27) can be
solved for the boundary conditions u(a) = u(b) = 0 by constructing an appro-
priate Green’s function G(x, &) on the square a < x, € <= b and setting
+ For a more thorough discussion, see J. L. Synge and B. A. Griffith, Principles of Mechanics, McGraw-
Hill, 1949, p. 99.
66 CHAPTER 2 Second-Order Linear Equations
where ¢, is set equal to w,/T in order to give a jump in slope of w,/T at the point
x = £, Passing to the limit as the Aé, | 0, we are led to guess that
9) = J" Cle,ule) dé
where
G(x,§) = |&(x
(§ — 1)x/T
~ 1)/T
Ox=xxsé
é=x=1
Gx, ) = |e)f(x) g)
«(E)fg(x)
asxx<t
é=x<b
will satisfy L[G] = 0 in the required intervals because L[f] = L[g] = 0; it will
satisfy (ii) because f(a) = g(b) = 0; and it approaches the same limit «(£)f(é)g(é)
from both sides of the diagonal x = &; hence it is continuous there. For the
factor e(£) to give 0G/dx a jump of 1/po(x) across x = &, a direct computation
gives the condition
where W = fg’ — gf’ is the Wronskian of f and g. Observe again that since f(a)
= g(b) = 0, Ga, §) = GO, §) = 0 for all & € [a, 8].
11 Green’s Functions, I 67
THEOREM 11’. For any continuous function r(x)on [a, b], the function u(x) €
@? defined by (39) and (39’) is the solution of pou” + piu’ + pou =
=
r that satisfies
the boundary conditions of u(a) = u(b) = 0, provided that W(f, g) # 0, i.2., that
there is no nontrivial solution of (35) satisfying the same boundary conditions.
The proof is similar to that of Theorem 10; the existence of two linearly inde-
pendent solutions of (35) is again assumed. Rewriting (38) in the form
r(x)
u”(x) = f Gy(x, &)r(&) d& + po
~~
)
r(x)
since L[G(x, §)] = 0 except at x = & Here the operator L acts on the variable
x in G(x, &); thoughG is not in @?, the expression L[{G] is meaningful for one-
sided derivatives and the above can be justified. This gives the identity (38).
Since G(x, ), considered as a function of x for fixed &, satisfies the boundary
68 CHAPTER 2 Second-Order Linear Equations
u(b) = 0,
completing the proof of the theorem.
(41) ff be — 9 ax = (, if £ € (a, b)
if£ ¢ [a, 4]
and the same initial or boundary conditions (in x). Now consider the function
EXERCISES F
In Exs. 1-3, (a) construct Green’s functions for the two-endpoint problem defined by the
DE specified and the boundary conditions u(0) = u(1) = 0, and (b) solve for 7(x) = x7:
1. u” — u = r(x) 2. u” + 4u = r(x)
io
3. u” — u = r(x)
x
— 1 2 1
In Exs. 4-5, find the conjugate points nearest to x = 0 for the DE specified.
4, u” + 2u’
+ 10u = 0
Show that the Euler-Lagrange DE for F(x, y, y) = gp(x)y + 4 Ty” (g, T constants) is
Ty” = gp(x). Relate to the sag of a loaded string under tension T.
ADDITIONAL EXERCISES
1. Show that the ratio v = f{/g of any two linearly independent solutions of the DE u”
+ q(x)u = 0 is a solution of the third-order nonlinear DE
a vi
Vv 3
(*) Sfv] = —
/
2 (
—_—
/ = 2q(x)
The Schwarzian S[v] of a function v(x) being defined by the middle term of (*), show
that S[(av + b)/(cu + d)] = S[0°] for any four constants, a, b, c, d with ad # be.
*3 Prove that, if vo, v,, vo, v3 are any four distinct solutions of the Riccati DE, their
cross ratio is constant: (v9 — v)(vs — v2)/(v9 — v9)(v3 — v1) = ¢.
on (1, +0) if B > }and a finite number if B < 4. (Hint: The DE is an Euler DE.]
10 For the DE u” + q(x)u = 0, show that every solution has a finite number of zeros
on (1, +00) if g(x) < 4x”, and infinitely many if g(x) > B/x*, B > 4.
*11. For the DE u” + g(x)u =
=
0, show that every solution has infinitely many zeros on
(1, +00) if
[P| -L]ae= 4
*12, Show that, if p, ¢ € @?, we can transform the DE (8) to the form d*z/d# = 0
in some neighborhood of the y-axis by transformations of the form z = f(x)y and
d& = h(x) dx. [HINT: Transforma basis of solutions to y, = 1, yo = &.]
CHAPTER 3
LINEAR EQUATIONS
WITH CONSTANT
COEFFICIENTS
So far, we have discussed only first- and second-order DEs, primarily because
so few DEs of higher order can be solved explicitly in terms of familiar func-
tions. However, general algebraic techniques make it possible to solve constant-
coefficient linear DEs of arbitrary order and to predict many properties of their
solutions, including especially their stability or instability.
This chapter will be devoted to explaining and exploiting these techniques.
In particular, it will exploit complex algebra and the properties of the complex
exponential function, which will be reviewed in this section. It will also apply
polynomial algebra to linear differential operators with constant coefficients,
using principles to be explained in §2.
The nth order linear DE with constant coefficients is
Here u stands for the kth derivative d*u/dx* of the unknown function u(x);
a), , 4, are arbitrary constants; and r(x) can be any continuous function. As
in Ch. 2, §1, the letter L in (1) stands for a (homogeneous) linear operator. That
is, Law + 6v] = aL[u] + BL[v] for any functions u and vu of class @” and any
constant a and 6.
As in the second-order case treated in Chapter 2, the solution of linear DEs
of the form (1) is best achieved by expressing its general solution as the sum u
Up, + u, of some particular solution u,(x) of (2), and the general solution u,(x)
=
=
71
72 CHAPTER 3 Linear Equations with Constant Coefficients
roots X ~p/2 wv =
=
V4q — p*), the real functions e~**/? px form
abasisoofsolutions.Wewillnowshowhowtoapplytheexponentialsubstitution
u =e to solve the general DE (2), beginning with the second-order case.
Loosely speaking, when the characteristic polynomial p() has n distinct roots
AL »Ay the functions ¢,(x)
= e* form a basis of complex solutions of the
DE (2) ‘By this we mean that for any “initial”
x =
=
%q and specified (complex)
numbers up, U4, uf, mere exist unique numbers ¢,, , ¢, such that
the solution f(x)
(a-1)
=
=
fO-M%)
Moreover, the complex roots A, of p(A) occur in pairs y, +3 vy just as in the
second-order case treated in Ch."9. Therefore, the real functions e* cos Vx,
e“* sin v,x together with the é* corresponding to real roots of(A) = 0, form a
basis of‘real solutions of (2).
Initial Value Problem. By the “‘initial value problem” for the nth order DE
(n— ) a solution
(1) is meant finding, for specified xp and numbers wp, uj, » Uo
u(x) of (1) that satisfies u(x9) up, and u(xo) = uf? for j = 1, n-1
If a basis of solutions ¢,(x) of the “‘reduced”’ DE (2) is known, together with
one “particular” solution u,(x) of the inhomogeneous DE (1), then the sum
u(x) = u,(x) + Le,b{x), with the c; chosen to make u,(x) = Lc,@,(x) satisfy
U,(%p) = Uy — uupl%)“uklxo) = ug ‘Uhl (n—1)
» U;, (%o) = uP ~ ue-PGxo),
constitutes one solution of the stated initial value problem. In §4, wewill prove
that this is the only solution (a uniqueness theorem), so that the stated initial
value problem is ‘“‘well-posed.”’
Clearly, the roots of the equation p(A) = 0 are the A,. The exponent &, in (4) is
called the multiplicity of the root \,; evidently the sum of the &, is the degree of
p. When all d, are distinct (ie., all k, = 1 so that m =
=
n), the DE has a basis of
complex exponential solutions ¢(x) = eM 7 =
=
1,2 , n; see §4 for details.
isA* = 1, with roots +1, +i. Therefore, a basis of complex solutions is provided
by e*, e~*, e™, and e”. From these we can construct a basis of four real solutions
20
Because e° cos @ + isin 6, one also often writes z x + ty as
z
= =
= =
=r,
+ 0
+ 4) =Inr
Inz = In(x
e* = gixtwx
(5)
=
=
d
t =Inx, _ =
=
> eM = xh
d d.
Corresponding to the real solutions é, te, ... of (2), we have real solutions
x, x In x, of (6).
Moreover, these can easily be found by substituting x* for u in (6). This sub-
stitution yields an equation of the form J(\)x* = 0, where J() is a polynomial
of degree n, called the indicial equation. Any A for which [(A) = 0 gives a so-
lution x* of (6); if X is a double root, then x* and x* In x are both solutions, and
so on.
Trying u
=
=
Cs) AA—
1) +pr(+q=0
du
(8)
—
dt?
+ (p- 1)
a 7 + mu =0, t=Inx
2 Complex Exponential Functions 75
since
d®
d d oa
de dx (*dx dx? + x
If (p — 1? > 4q, the indicial equation has two distinct real roots \ = a and
\ = 8, and so the DE (7) has the two linearly independent real solutions x* and
x®, defined for positive x. For positive or negative x, we have the solutions |x |*
and |x|® since the substitution of —x for x does not affect (7). Note that the
DE (7) hasa singular point at x = 0 and that |x| has discontinuous slope there
ifa 31.
When the discriminant (p — 1)? — 4g is negative, the indicial equation has
two conjugate complex roots \ = » + w, where w = (1 — p)/2 andy = [4q —
(p — 1)*)'/7/2. A basis of real solutions of (8) is then e“ cos vt and e“ sin vt; the
corresponding solutions of the second order Euler homogeneous DE (7) are
x" cos(v In x) and x" sin(v In x). These are, for x > 0, the real and imaginary
parts of the complex power function
A oe hth (usw) In x
x =e = x*[cos@ In x) + i sin( In x)]
as in (5). For x < 0, we can get real solutions by using |x| in place of x. But for
x < 0, the resulting real solutions of (7) are no longer the real and imaginary
parts of x, because In(—x) = In x + ix; cf. Ch. 9, §1.
General Case. The general nth-order case can be treated in the same way.
We can again make the change of independent variable
t
x =e, t = Inx, x=
dx di
This reduces (6) to a DE of the form (2), whose solutions ve give a basis of
solutions for (6) of the form (In x)’x’.
EXERCISES A
1. u” + 5u’ + 4u = 0 2,.u” =u
3. ua’ =u *4.u"+u=0
5. u” + Qiu’ + 3u = 0 6. u” — 2u’ + 2u = 0
In Exs. 7-10, find the solution of the initial value problem specified.
11 xu”
+ 5xu’ + 3u
=0 12. x*u” + 2ixu — 3u = 0
LEMMA. Linear operators with constant coefficients are permutable: for any con-
stants ayby if p(D) Xa;D) and q(D) <b,D*, then p(D)q(D) q(D)p(D)
Labpit
Proof. Iterate the formula D[b,D[u]] = b,D°[u]. It follows that, for anytwo
constants a; and 5, and any two positive integers j and k, we have a,D/b,D* =
Moreover, the operators (D — },)" being permutable, we can write, for any i
However, when all the coefficients a, are real numbers, more detailed infor-
mation can be obtained about the solutions, as follows.
LEMMA. Let the complex-valued function w(x) = u(x) + iv(x) satisfy a homo-
geneous linear DE (1) with real coefficients. Then the functions u(x) and v(x) [the real
and imaginary parts of w(x) both satisfy the DE.
Proof. The complex conjugatet w*(x) = u(x) — iv(x) of w(x) satisfies the
complex conjugate of the given DE (2), obtained by replacing every coefficient
a, by its complex conjugate af, because L*[w*] = {L[w]}* = 0. If the a, are real,
then a¥ = a, and so w*(x) also satisfies (2). Hence, the linear combinations
COROLLARY 1. If the DE (2) has real coefficients and ¢ satisfies (2), then so
does é”*, The nonreal roots of the characteristic polynomial (3) thus occur in conjugate
pairs >, = pw, £ iv, having the same multiplicity k,.
These solutions differ from the solutions ¢* with real) in that they have infi-
nitely many zeros in any infinite interval of the real axis: that is, they are oscil-
latory. This proves the following result.
+ The complex conjugate w* of a complex number w = u + iv is u — iv. Some authors use w instead
of w* to denote the complex conjugate of w.
78 CHAPTER 3 Linear Equations with Constant Coefficients
THEOREM 2. [If the characteristic polynomial (3) with real coefficients has 2r non-
real roots, then the DE (2) has 2r distinct oscillatory real solutions of the form (9).
4 SOLUTION BASES
We now show that all solutions of the real homogeneous linear DE (2) are
linear combinations of the special solutions described in Corollary 2 above. The
proof will appeal to the concept of a basis of solutions of a general nth order
linear homogeneous DE
The coefficient-functions p,(x) in (10) may be variable, but they must be real and
continuous.
The aim of this section is to prove that the special solutions described in Cor-
ollary 2 form a basis of real solutions of the DE (2). The fact that every nth order
homogeneous linear DE has a basis of n solutions is, of course, a theorem to be
proved.
First, as in Ch. 2, §2, we define a set of n real or complex functions f,,
Se . »f, defined on an interval (a,b) to be linearly independent when no linear
combination of the functions with constant coefficients not all zero can vanish
identically: that is, when Lf.) ¢,f,(x) = 0 implies c, = cg = +: - =
=
= 0. A
C,
set of functions that is not linearly independent is said to be linearly dependent.
There are two notions of linear independence, according as we allow the coef-
ficients ¢, to assume only real values, or also complex values. In the first case, one
says that the functions are linearly independent over the real field; in the second
case, that they are linearly independent over the complex field.
Proof. Linear dependence over the real field implies linear dependence over
the complex field, a fortiori. Conversely, the f(x) being real, suppose that Xc,/, (x)
= 0 fora < x < 6, Then [Xe,f-(x)]* = 0, and hence Xc*f(x) = 0. Subtracting,
we obtain X[(c, — c¥)/t] f(x) = 0. If all c, are real, there is nothing to prove. If
some c,J is not real, some real number (c, — c¥)/i will not vanish, and we still have
a vanishing linear combination with real coefficients.
A set of functions that is linearly dependent on a given domain may become
linearly independent when the functions are extended to a larger domain. How-
ever, a linearly independent set of functions clearly remains linearly indepen-
dent when the functions are so extended.
4 Solution Bases 79
where the r are nonnegative integers and the d, complex numbers, is linearly indepen-
dent on any nonvoid open interval, unless two or more of the functions are identical.
Proof. Suppose that Xc, f,(x) = 0. For any given A, choose R to be the larg-
est r such that c, # 0. Form the operator
where for each i # j, k, is the largest r such that x’e** is a member of the
set of functions in (11). It follows that g(D)[f,] = 0 unless 7 = j, and that
q(D)[f,] = 0 for r < R. Hence, we have
QD)[Xepfrl*)) = cpg(D)ix"e*)}
On the other hand, as in the proof of Theorem 1, we see that
Therefore, substituting back, we find that cz, = 0. Since we assumed that cp, #
0, this gives a contradiction unless all c, = 0, proving linear independence.
The analogous result for real solutions of DE of the form (2) with real coef-
ficients can be proved as follows. For any two conjugate complex roots A = 4
+ w and A* # — w of the characteristic equation of (2), the real solutions
=
=
x’e cos vx and x’e“ sin yx are complex linear combinations of x’e and xrox
and conversely. Hence, they can be substituted for x’e* and xrone in any set of
solutions without affecting their linear independence. Since linear indepen-
dence over the complex field implies linear independence over the real field,
this proves the following.
We now show that all solutions of the real homogeneous linear DE (2) are
linear combinations of the special solutions described in Corollary 2. (The proof
80 CHAPTER 3 Linear Equations with Constant Coefficients
LEMMA 38. Let f(x) be any real or complex solution of the nth order homogeneous
linear DE (1) with continuous real coefficient-functions in the closed interval [a,b]. If
f@ =f'@=-+ + =f" @ = 0,then f(x) = 0 on [4,0].
satisfies the initial condition o(a) = 0. Differentiating o(x), we find, since o(x) is
real, that
o(x) < (f? +f” + (f? +f") tee et fe??? + Lf") + af e-Dee
Since L[f] = 0, it follows that f™ = —LXf., pf”. Hence, the last term can
be rewritten in the form
We now show that any n linearly independent solutions of (10) form a basis
of solutions.
THEOREM 3. Let u;, » Uy, be n linearly independent real solutions of the nth
order linear homogeneous DE (10) with real coefficient-functions. Then, given arbitrary
real numbers a, Ug, Up, ., ug), there exist unique constants c,, . » Cy Such that
u(x) = Xc,u,(x) is a solution of (10) satisfying
; u® Yea) = ue)
(12’) Uos u’(a) = ud,
=
u(a) =
Theorem 3 follows readily from the lemma. Suppose that, for some a, uo, u4,
, uf’), there were no linear combination Xc,u,(x) satisfying the given initial
conditions (12’). That is, suppose the 7 vectors
(n—
[u,(a), uj(a), » Uj %@), k=1,...,n
were linearly dependent. Then there would exist constants y,, ȴw not all zero,
such that
fix) =
obtain u(x) = v(x) and v(x) = Xe,u,(x), proving the second conclusion of Theo-
rem 3.
0, s k
J
— 1, are a basis of complex solutions of (2).
COROLLARY 2. If the coefficients of the DE (2) are real, then it has a basis of real
solutions of the form xe, xe cos vx, and x’e™ sin vx, where d, p, and v are real
constants.
EXERCISES B
1. Solve the following initial-value problems:
(a) uw” —u =
=
—2
(c) u” + u” =
=
(a) Find a DE L[u] = 0 of the form (2) having e~4, te“, and ¢ as a basis of solutions.
(b) For this linear operator L, find a basis of solutions of the sixth-order DE
L?[u] = 0 and the ninth-order DE L?[u] = 0.
Find bases of solutions for the following DEs:
(a) u" = u (b) u” — 3u” + 2u = 0
(c) u” + 6u” + 12u’ + 8u = 0
(d) u” + 6u” + 12u’
+ (8 + du = 0
Knowing bases of solutions L,[u] = 0 and L.[u] = 0 of the form given by Theorem
1, find a basis of solutions of L,[L,[u]] = 0.
Show that in every veal DE of the form (2), L can be factored as L = Lily... L,
where L, = D, + 6, or L, = D? + pD + q with all b, p, q, real.
Extend Lemma 2 of §4 to the case where the r are arbitrary complex numbers.
“7 State an analog of Corollary 2 of §4 for Euler’s homogeneous DE, and prove your
statement without assuming Corollary 2.
+ The preceding result can be proved more generally for linear DEs with constant complex coeffi-
cients, by similar methods; see Ch. 6, §11.
5 Inhomogeneous Equations 83
5 INHOMOGENEOUS EQUATIONS
u du
(13) L{u] = + a, —__—
df" 1
+--+ + a4,-; au7 T Gntt =
=
r(t)
dt d
already introduced in §1. As in the second-order case of Ch. 2, §8, the function
r(f) in (13) may be thought of as an “input” or “source’’ term, and u(é) as the
“output’’ due to r(é). We first describe a simple method for finding a particular
solution of the DE (13) in closed form, in the special case that r(#) = Lp,(terk!
is a linear combination of products of polynomials and exponentials.
We recall that, by Lemma 2 of §3, &
(D — »NeMAH) = Mf’
As a corollary, since every polynomial of degree s is the derivative r(t) = q’(i) of
a suitable polynomial q(é) of degree s + 1, we obtain the following result.
(D — A)[u] = er)
has a solution of theform u = eq(t), where q(t) is a polynomial of degree s.
Applying the two preceding lemmas repeatedly to the factors of the operator
-+(D—),)*
L = p,(D) = (D — d)"(D — d»)”
Knowing the form of the answer, we can solve for the coefficients 5, of the
unknown polynomial q() = £8,t" by the method of undetermined coefficients.
Namely, applying p(D) to u(é) = e™(r,t'), one can compute the numbers P,, in
the formula
p(D)[u] = MI(Pybyt"
using formulas for differentiating elementary functions. One does not need to
factor p,. The simultaneous linear equations “P,,b, = ¢, can then be solved for
the b,, given r(t) = ct", by elementary algebra. Theorem 4 simply states how
many unknowns 5, must be used to get a compatible system of linear equations.
that satisfies the initial conditions u(0) = —17/3, w’(0) = u”(0) = 1/3. First,
since the two-dimensional subspace of functions of the form (a + Bie’ is
mapped into itself by differentiation, the constant-coefficient DE (*) may be
expected to have a particular solution of this form. And indeed, substituting u
= (a + Bie’ into (*) and evaluating, we get
EXERCISES C
In Exs. 1—4, find a particular solution of the DE specified. In Exs. 1—3, find the solutions
satisfying (a) u(0) = 0, u’(0) = 1 and (b) u(0) = 0, w’(0) = 1.
1. u” = te 2.u”+u= te
3. u” — u = te’ 4. uu” = fd
Stability 85
In each of Exs. 9-12, find (a) the general solution of the DE specified four exercises
earlier, and (b) the particular solution satisfying the initial condition specified.
9 u(0) = 0 fory = 0, 1,2, 3,4 u(0) = 1
10. u(0) = u’(0), u”(0) = 1
6 STABILITY
dy ad”*u
(14)
+
ap
T ag7yn?
—_——-
Foss tay + au = 0
dt"
THEOREM 5. A given DE (14) is strictly stable if and only if every root of its
characteristic polynomial has a negative real part. It is stable if and only if every mul-
86 CHAPTER 3 Linear Equations with Constant Coefficients
tiple root dX, [with k, > 1 in (4)] has a negative real part, and no simple root (with
k, = 1) has a positive real part.
Polynomials all of whose roots have negative real parts are said to be of stable
type.t There are algebraic inequalities, called the Routh-Hurwitz conditions, on
the coefficients of a real polynomial, which are necessary and sufficient for it to
be of stable type. Thus, consider the quadratic characteristic polynomial of the
DE of the second-order DE (5) of Ch. 2, §1. An examination of the three cases
discussed in §2 above shows that the real DE
au du
u + au + agu = 0,
=
u =
a’ at
is strictly stable if and only if a, and a, are both positive (positive damping and
positive restoring force). That is, when n = 2, the Routh-Hurwitz conditions
are a, > 0 and a, > 0.
To make it easier to correlate the preceding results with the more informal
discussion of stability and oscillation found in Ch. 2, §2, we can rewrite the DE
discussed there as u + pu + qu = 0. We have just recalled that this DE is strictly
stable if and only if p > 0 and q > 0. It is oscillatory if and only if g > p’/4,
so that its characteristic polynomial A? + pdr + gq has complex roots
—(p £ Vp" — 49)/2.
In the case of a third-order DE (n 3), the test for strict stability is provided
=
=
by the inequalities a, > 0 G = 1, 2, 3) and a,a) > as. When n = 4, the condi-
tions for strict stability are a, > 0 =
1, 2, 3, 4), aja@q > ag, and a,a,a3 > ara,
+ aj.
When n > 2, there are no equally simple conditions for solutions to be oscil-
latory or nonoscillatory. Thus, the characteristic polynomial of the DE # + & +
ua + u = 0is A + 1)Q? + 1); hence its general solution is
t
acosit+ bsint + ce~
Unless a = 6 = 0, this solution will become oscillatory for large positive ¢, but
will be nonoscillatory for large negative ¢. Other illustrative examples are given
in Exercises C.
Inhomogeneous linear DEs (13) are widely used to represent electric alter-
nating current networks or filiers. Such a filter may be thought of as a “black
+ See Birkhoff and MacLane, p. 122. For polynomials of stable type of higher degree, see F. R.
Gantmacher, Applications of Matrices, Wiley—Interscience, New York, 1959.
7 The Transfer Function 87
box’’ into which an electric current or a voltage is fed as an input r‘t) and out of
which comesa resulting output u(i).
Mathematically, this amounts to considering an operator transforming the
function r into a function u, which is the solution of the inhomogeneous linear
DE (13). Writing this operator as u = F[r], we easily see that L[F[r]] = r. Thus,
such an input-output operator is a right-inverse of the operator L.
Since there are many solutions of the inhomogeneous DE (13) for a given
input r(¢), the preceding definition of F is incomplete: the preceding equations
do not define F = L™' unambiguously.
For input-output problems that are unbounded in time, this difficulty can
often be resolved by insisting that F[r] be in the class B(—0o, +00) of bounded
functions; in §§7—8, we will make this restriction. For, in this case, for any two
solutions u, and wu, of the inhomogeneous DE L[u] = 1, the difference v = u,
— Uy would have to satisfy L{v] = 0. Unless the characteristic polynomial p, (A)
= 0 has pure imaginary roots, this implies v = 0. Hence, in particular, the DE
L{u] = rhas at most one bounded solution if the DE L{u] = 0 is strictly stable—
an assumption which corresponds in electrical engineering to a passive electrical
network with dissipation. Moreover, the effect of initial conditions is ‘‘tran-
sient’: it dies out exponentially.
For initial value problems and their Green’s functions, it is more appropriate
to define F by restricting its values to functions that satisfy u(0) = w’(0)
=
=
A is called the amplitude, k/2x the frequency, and a the phase constant. The fre-
quency k/2z is the reciprocal of the period 2x/k.
Except in the case p,(tk) = 0 of perfect resonance, there always exists a
unique periodic solution of the DE (13), having the same period as the given input
function r(¢) = ce. This output function u({) can be found by making the sub-
stitution u = C(k)ce™, where C(k) is to be determined. Substituting into the inho-
mogeneous DE (14), we see that L[C(k)ce™] = ce™ if and only if
The reason for this terminology lies in the relationship between the real part
of u(t) and that of the input r(é). Clearly,
This shows that the amplitude of the output is p(k) times the amplitude of the
input, and the phase of the output lags ~y = —arg C behind that of the input at
all times.
In the strictly stable case, the particular solution of the inhomogeneous linear
DE L{u] = ce found by the preceding method is the only bounded solution;
hence F[{ce“] = C(k)ce™ describes the effect of the input~output operator F on
sinusoidal inputs. Furthermore, since every solution of the homogeneous DE
(14) tends to zero as t > +00, every solution of L[{u] = ce™ approaches C(k)ce™
exponentially.
(*) [L[u]
=
=
u” + eu’ + pu = sin kt, e< 1]
The transfer function of (*) is easily found using the complex exponential trial
function e. Since
we have C(k) = 1/[(p? — k?) + eik]. De Moivre’s formulas give from this the
gain function p = 1/[(p? — k*)? + ¢k*]'” and the phase lag
ek
Y = arctan
——_—_——
(p? _— i)
|
The solution of (*) is therefore p sin (kt — ‘y), where p and y¥ are as stated.
One can also solve (*) in real form. Since differentiation carries functions of
the form u = a cos kt + b sin ki into functions of the same form, we look for a
periodic solution of (*) of this form. An elementary computation gives for u as
before:
To make the coefficient of cos kt in (*) vanish, it is necessary and sufficient that
a/b = ck/(k® — p*), the tangent of the phase advance (negative phase lag). The
gain can be computed similarly; we omit the tedious details.
Finally, note that the characteristic polynomial of any real DE (2) can be fac-
7 The Transfer Function 89
and since all b, p, and q, are positive in the strictly stable case that all roots of
pi) = 0 have negative parts. Therefore
L{te“| = | 2 *| = = 16 —
= (pa = PLOe™ + piQyte™
—
=
we obtain, setting A = ik
L{te™] = pi(ik)e™
Periodic Inputs. The transfer function gives a simple way for determining
periodic outputs from any periodic input function (2) in (13). Changing the time
90 CHAPTER 3 Linear Equations with Constant Coefficients
co
Je
|
ke
(17) ut) = >>
k= —00 pith)
provided that no p,(ik) vanishes. The series (17) is absolutely and uniformly con-
vergent, since p,(ik) = 0(k-") for an nth-order DE. We leave to the reader the
proof and the determination of sufficient conditions for term-by-term
differentiability.
EXERCISES D
In Exs. 1-4, test the DE specified for stability and strict stability.
6. D?+4D+4+
41 7, D+ 6D?
+ 12D + 81
8. D? + 2D + 1017 9. Dt-—I
10. Fora strictly stable L[u] = u” + au’ + bu = r(i), calculate the outputs (the responses)
to the inputs 7() = 1 and r(#) = ¢ for a? > 46 and a? < 40,
|
from which these graphs are easilyplotted. Figure 3.1a depicts thegain function
and phase lag of the DE
THEOREM 6. The DEL [u] = r(t) of order n is strictly stable if and only if the
phase lag-y = —arg C(k) increases by nm as k increases from —©0 to + 00. In this
case, the phase lag increases monotonically with k.
if the differential operator L has real coefficients, then y(—k) = —y(&) and
+
p(—k) = p(k), since the complex roots A, occur in conjugate pairs y, ~_ wv, In
particular, we have 7(0) = 0, and the change in (A) as k increases from 0 to
00 is (m — p)x/2. This proves the following specialization of Theorem 6.
If all roots \, of the characteristic polynomial are real, then one easily verifies
that all In |ik — d,| increase monotonically as k increases from 0 to 00, Hence,
+ For purely imaginary roots, the change of argument of A is undefined (it could be x or —z). In
this case, we make the convention that the change in the argument is zero. The following theorem
is true with the proviso that, whenever the argument is undefined, the change is taken to be zero.
92 CHAPTER 3 Linear Equations with Constant Coefficients
v(k)
e(k)
3x ¥(k)
1.0
0.5
In p(k)
0.2
0.1
05 (a)
0.5 1.0 15 2.0 2.5 k
him {C(k)}
k=-1
k=—-0.8
k= —1.2
k=-2
k=0.5
k=0
Re {c(k)}
k= 0.5
km 1.2
15 kL2
k=0.8
1
k=
(b)
Figure 3.1 (a) Phase-lag and gain functions, (6) Nyquist diagram.
in this case, the gain p(k) is a monotonically decreasing function of the fre-
quency. In the case of complex roots A, = #, + i, with p, very small, the gain
function p(k) is very large near k = »,; this is due to near resonance, as in Example
3 above.
Another useful way to visualize the transfer function is to plot the curve
z = C(k) in the complex plane as k ranges through all real values. The curve thus
obtained is called the Nyquist diagram of DE (14). Figure 3.1 depicts the Nyquist
diagram of the DE below (18b).
9 The Green’s Function 93
Since C(x) is the inverse of a polynomial, it tends to the origin as k > +00
that is, it “starts” and “ends” at the origin. It is a continuous curve except when
the characteristic equation has one or more imaginary roots A, = ik, From
Theorem 6 we obtain the following Nyquist Stability Criterion
COROLLARY 2 The equation L{u] 0 is strictly stable if and only if the Nyquist
diagram for C(k) turns through —nx radians as k increases from — 00 to +00
If L is real, then C(—A) = C*(k), and it suffices to plot half of the Nyquist
diagram. The operatorL is strictly stable if and only if the Nyquist diagram turns
through n/2 radians as k increases from 0 to 00
The concept of the Green’s function for initial value problems was intro-
duced in Ch. 2, §9. For any inhomogeneous linear DE L[u]
=
=
r(é), it is a func-
tion Gt, 7) such that
THEOREM 7. The Green’s function for the initial value problem of the nth order
real linear differential operator with continuous coefficients
-1
is zero ift <1. For t = 7, it is that solution of the DE L[G]= 0(for fixed + and
variable t) which satisfies the initial conditions
= dG/at ve = OG /A? =
=
In the case p,¢) = a, of linear DEs with constant real coefficients, the exis-
tence of sucha solution follows from the results of §3. Green’s function is easily
computed as a sum of polynomials times exponentials. Thus, if (20) is uw” + 3
+ 3u’ + u, the Green’s function is
ifi<r
G(, 7) {° —_ ry? T 12 ift=rT
94 CHAPTER 3 Linear Equations with Constant Coefficients
For variable coefficient-functions, existence will follow from the results of the
next chapter.
We omit the proof of Theorem 7. It follows exactly the proof for second-
order differential operators, given in Ch. 2, §9. This can be easily extended to
the present case: one simply applies Leibniz’s rule for differentiation under the
integral sign n times instead of twice.
The computation of Green’s functions for linear DEs with constant coefhi-
cients is most easily performed and its significance best understood by using the
following result.
THEOREM 8. The Green’s function for the initial value problem of any linear
differential operator with constant coefficients is a function G(t, r) = T(t — 7) depend-
ing only on the difference t — T.
Proof. Let T(@é) = G@, 0); then T(@) = 0 ift < 0. If t = 0, the function I
is the solution of the DE L{[T] = 0, which satisfies the initial conditions
We now remark that, if u(f) is a solution of the DE L{u] = 0, for each fixed
7 the function u(t + 7) of the variable ¢ is also a solution of the DE. It follows
that the function Fi) = G( + 7, 7) (for fixed 7) is a solution of the DE. This
function satisfies the same initial conditions as the function I, because of the
way Green’s functions are defined. By the uniqueness theorem (Theorem 3),
it follows that I) = G(t + 7, 7). Hence, setting s t + 7, we obtain I'(s — 7)
=
=
COROLLARY 2. [fr(t) vanishes for t < a, and is bounded and continuous for a
XS t, then the function
OO,
(21a) fo = f- TG — r)r(r) dr = f —_
~ T(s)r(é — 5) ds
is a solution of the inhomogeneous linear DE with constant coefficients (13) on [a, 00],
which satisfies f(a) = f@=--- =f Ng) = 0. (Note that, unless r(a) = 0,
f(a) does not exist.)
+ The function I(s) is, of course, also expressible for s > 0 as a real linear combination of functions
of the form (9).
9 The Green’s Function 95
Indeed, since r(r) vanishes for r < a and I'(é — 7) vanishes for rt > t, the
integral (21a) can be written as
(21b) fo = f Té — r)r(r) dr
THEOREM 9. [fL is strictly stable, formulas (21a) remain valid for any bounded
continuous function r(t) defined for —00 <t < 00; for such a function, f (1) is the only
solution of the DE (13) which is bounded for —00 < t < co,
If the equation L[{u] = 0 is strictly stable, then the real parts of all \, are nega-
tive. Let —m be the largest of these real parts. Then —m < 0, and
Since Re {\,} + (m/2) < 0 for 1 <j < 1, the right side remains bounded for 0
<= 5s = ©, Let M be an upper bound for the right side. Then we obtain
We next show that the integrals in (21a) are well-defined for any bounded
continuous function r. The first integral can be rewritten in the form
rf
J =
~ Tt — v)r(r) dr
since '¢ — 7) = 0 for ¢ < 7. Using the foregoing bound for I'(s), and letting R
be an upper bound for |7(r)| on —00 < 7 < 00, we obtain
for all t. Hence the integral is well-defined and defines a bounded function /f().
To show that fis a solution of the DE, we can argue as in Theorem 10 of Ch. 2,
96 CHAPTER 3 Linear Equations with Constant Coefficients
provided we can carry out the differentiation under the integral sign. This can
indeed be justified}; instead, however, we shall give a direct argument.
Consider the sequence of functions 7,(#) defined by the formulas
ift
= —k
(23) r,t) = |rd)
0 ift< —k k=1,2,...
are solutions of the DE (14). We shall show that, for ¢ ranging over any interval
a <= ts Bb, the functions f,(, as well as their derivatives of orders up to n,
converge uniformly to the derivatives of the function f(é). This will also prove
that f(/) is a solution of the DE (14).
From the expression (22) for I(s) as a linear combination of functions of the
form s‘e', we see that all derivatives of I'(s) are also linear combinations.of func-
tions of the same form, for different 7, but with the same sequence of exponents
A, That is, for the derivative of order £, we have
where the p, are polynomials in the variable s, depending on the order of differ-
entiation ¢.} It follows, as before, that
PO = f ' TO — r)r,(r) dr
we find, for sufficiently large k and j, where k = j
= f- [TC — a)|ln@)I ar
and the last integral clearly tends to zero as j, k > ©, uniformly for a Si S b.
Therefore, [f/() — f@®| < ¢ for sufficiently large 2, j, uniformly for
a =i <b. This completes the proof of the fact that fis a solution of the DE.
Lastly, one can easily see, by the following argument, that f thus defined is
the only bounded solution. If f,; were another bounded solution, then f — f,
would be a bounded solution of the homogeneous DE. But, since the DE is
strictly stable as ¢ — ©, no nontrivial solution of the homogeneous DE can
remain bounded as tf > —© (cf. Theorem 4). Hence, f — f; = 0, and the proof
is complete.
10,
(24) h(x) = f =-
~ Se — dg() at
whenever the integral is finite. If the functions f and g are identically zero for
i < 0, this formula simplifies to
= ye
the initial conditions u(0) = u’(0) = - - (0) = 0 is the convolution
r *T of r and the Green’s function I for the same initial conditions, provided
that r= 0 fori
< 0.
Theorem 9 can also be interpreted in terms of the convolution operation. It
asserts that, for strictly stable L, the solution of the “input-output” problem
Liu]
—
=
EXERCISES E
In Exs. 1-6, construct the Green’s function for the initial value problem of the DE
indicated.
*9 Show that, if (19) is defined for all ¢ and the Green’s function G(t, 7) = Tt — 7),
then all coefficients p,(f) are constant.
*10 Show that, if u(¢ + 7) is a solution of (19) with r(@) = 0 whenever u(t) is, then the
coefficients p,(é) are all constants.
98 CHAPTER 3 Linear Equations with Constant Coefficients
11 . (a) Show that in Ex. D1, the values of p,(ék) traverse the parabola x = 4 —(y/5)?.
(b) Verify Theorem6in this case.
12 For u” + au’ + bu = ce“, make graphs of the gain function versus the dimension-
less frequency k/\V/b for the values 7 = 0,0 =1/2,0= 1/v2, ” = 2 of the param-
eter7 = a/2V/b (b > 0).
13 Show that, if p, (ik) = pik) = - > - = pith) = 0 but p(ék) ¥ 0, then
asolution
of L{u] = eis u(t) = (1/pi(anyt"e.
14 A DE (13) is stable at t —© when all solutions of the DE remain bounded as ¢
— —0o, Find necessary and sufficient conditions for stability at —00,
15 In Ex. 14, find necessary and sufficient conditions for strict stability at —o.
POWER SERIES
SOLUTIONS
1 INTRODUCTION
x
sin x do
« f
(1)
J dx and
V1 — #? sin
a,x" a3x°
fe + ayt + ast® ) dt
= C + agx + —
2
+ —
3
+
+ Cf. Dwight. A valuable collection of analogous explicit solutions of ordinary DEs may be found in
Kamke, pp. 293-660
99
100 CHAPTER 4 Power Series Solutions
3 5
sin & x
(2) Si(x) = dE =x — ——
~~ wee
187 600
2 yt
(Since E\(x) = —
This approach is especially easy to carry out for first-order linear DEs. For
example, we can use it to derive the key properties of the exponential function
E(x) (or e*) as the (unique) solution of the DE E’(x) = E(x) that satisfies the
initial condition E(@) = 1. For any a, f(x) = E(@a + x) must satisfy f(0) = E(a)
and f’(x) = E’(a + x) = E(a + x) = f(x). Since f(x) and E(x) are solutions of
the same first-order linear homogeneous DE, it follows that f(x) = f(O)E(x) =
ents
E(a)E(x), giving the formula
=
=
e“e*. In particular, E(a) can never vanish and
is always positive, together with E’(a) = Ea), E”’(a) = Ef@),.. , which shows
that E(x) is increasing and convex. Finally, its Maclaurin series is
In Ch. 2, §§6—7, we have seen how one can derive many oscillation and non-
oscillation properties of solutions of linear second-order DEs
can be expressed as sums of convergent power series, we will show in this chap-
ter how to find a basis of solutions of the DE (4) having the same form. The
functions so constructed include many of the special functions most commonly
used in applied mathematics.
Namely, by substituting a formal power series
we shall first show how to compute the unknown coefficients a, in (4’). We shall
then show that the radius of convergence of the resulting formal power series
(4’) is at least as great as the lesser of the radii of convergence of the series for
p(x) and g(x).
We will then give a similar construction for the solutions of normal first-order
DEs of the form y’ = F(x,y), where
2k+1
k=0
1)2(2k)!
Expanding the integrand in power series, we can obtain similarly the first few
terms of the series expansion for the elliptic integral of the first kind,
F(k,sin™!x) =
=
We now give some other applications of the same principle to some special
functions familiar in applied mathematics, defined as solutions of second-order
linear homogeneous DEs (4) with analytic coefficient-functions.
d
(6)
dx
[a - od + hu = 0
where A is a parameter.
Substituting the series (4’) into the linear DE (6), and equating to zero the
coefficients of 1 = x°, x, x2 ,-+., We get an infinite system of linear equations
kk +1)—A
(6’) +2 = (k + lk + 2) ay
This relation defines for each \ two linearly independent solutions, one consist-
ing of even powers of x, and the other of odd powers of x. These solutions are
power series whose radius of convergence is unity by the Ratio Test, unless \ =
n(n + 1) for some nonnegative integer n. When A = n(n + 1), the Legendre
DE has a polynomial solution which is an even function if n is even and an odd
function if m is odd. These polynomial solutions are the Legendre polynomials
P,,(x).
analytic near the origin. That is (cf. §1), they can be expanded into power series
1 Po(z)
P, (x)
P,(z)
P,(z)
-l
P,(z)
~i+
k=1
xepores
Gu= qody+ (Goa+ Gyaq)xTess + bp tite x +
na]
Substituting into (4) and equating the coefficients of 1, x, Xx ,-.. to zero,
we get successively
| >
k=0
(n—R)pQn—~+ > tie
(8) n=l
=
_—
ant) = >
n(n + 1)
(9) u” — QIxu’ + ru = 0
2k—2X
(9) Qn+2
(kh + DR +2) 7
which again gives, for each A, one power series in the even powers x** of x and
2k+1
another in the odd powers x . These power series are convergent for all x; if
\ = 2n is a nonnegative even integer, one series is a polynomial of degree n,
the Hermite polynomial H,(x).
Caution. We have not stated or proved that the formal power series (7’) con-
verges or that it represents an analytic function. To see the need for proving
this, consider the DE x?u’ = u — x, which has the everywhere divergent formal
power series solution
x + x? + (2Nx? + (3Nxt + - +n
— Vint +
For normal second-order linear DEs (4), the convergence of the power series
defined by (7’) to an analytic solution will be proved in §6. First we treat some
more special cases, in which convergence is easily verified.
EXERCISES A
1 (a) Prove that the Legendre DE has a polynomial solution if and only if
AX = n(n
+ 1).
(b) Prove that the radius of convergence of every nonpolynomial solution of the
Legendre DE is one.
Find a recurrence relation like (8’) for the DE (1 + x’)y” = y, and compute expan-
sions through terms in x!° for a basis of solutions.
(a) Find power series expansions for a basis of solutions of Airy’s DE u” + xu = 0.
(b) Prove that the radius of convergence of both solutions is infinite.
(c) Show that any solution of Airy’s DE vanishes infinitely often on (0,0), but at
most once on (—©,0).
(d) Reduce u” + (ax + Bu = 0 to d®u/dt®? + tu =
=
0 by a suitable change of inde-
pendent variable.
Show that F(r,0) = r"P(cos 6) satisfies V?F = 0 if and only if P(x) satisfies the
Legendre DE with A = n(n + 1).
*8 Show that, for any positive integer n, the polynomial d*[(x? — 1)"]/dx" is a solution
of the Legendre DE with A = n(n + 1). (This is Rodrigues’ formula.)
MORE EXAMPLES ‘
(10’) (xu’l + xu =
=
0
Substituting u
—
=
2
2 3
Jot)
=
| }+( } |
x
_
2
x
2-4
x
2-4-6
)+
(11)
2 4 6 2r
x x x
+ .--
1-—+—-~ +-+++(-17
4 64 2304 [2’@ry]?
This series defines the Bessel function of order zero. It is convergent for all x
(by the Ratio Test), and converges rapidly if |x| < 2.
Similar calculations give, for general n, the solution
2 4
cw(
(Din + n! |
whose coefficients 5,
=
=
For general n, one can verify, in the same way, the recursion formula
Clearly, formula (13) defines Ji, Jo, Js, . . . recursively from Jo.
where ) is a parameter.
If we set u = Lf.o a,x", substitute into Eq. (14), and collect the terms in x”,
we get
Jo(z)
J; (x)
ISS
"|
Figure 4.2. Bessel functions Jo(x) and J,(x) = — Jo(x).
4 Three First-Order DEs 107
For most values of \, this leads to two solutions of (10), one an even function
2
and the other odd. However, when X n is a square, we have instead a poly-
nomial solution
Even simpler to solve by power series is the Airy DE
(15) u” + xu=
0
One easily derives from (15) the recursion relation n(n — 1)a, + a,-3 for the
coefficients. A basis of power series solutions therefore consists of the functions
x? x® 9
A1(x)
=
=
1-> +757
180 12960
and
4 7 10
x x x
+
Bi(x) x-isteyt
12 504 45360
The three homogeneous, linear, second-order DEs just treated were quite sim-
ilar to each other. In this section, we will derive power series expansions for
three first-order DEs having much less in common
—* 2 3
é x x
Seet-1t5-g+g7:
9} 3!
x
k
OO x
~ dt
(16) Ey) —{ x 1+ @/x) x
Ja-2+2 pto0)
108 CHAPTER 4 Power Series Solutions
The final series of negative powers is divergent for all x: it is a so-called asymptotic
series (cf. Ch. 7, §'7). However, the partial sum of the first n terms hasa relative
error of less than 1%, and an absolute error less than 107° for x 2 10.
Example 6. Pearson’s DE is
and
The two sides of the preceding equations are equal if and only if
_ [Fay + (D — Ba,)]
a, = Dao, ao =
2
[E + D(l — B)}x?
y= 1+ Det
2
+ x a,x"
h=3
where the a, (k 2 3) are computed recursively from the last previously displayed
formula.
yy” = qd + yy
Qyy’ = Qy(1 + y?)
=
=
a ~%_
=1l+y=14+2 and z(0)
dt ~ dx
y =x + agx + asx?
+ ayx’ + agx® +
and so on. Solving recursively, we get the first few terms of the power series
expansion for tan x,
2x° 17
(**) tanx=x45
3
+384 HF
315
x? + =
62
2835
x?
+
The radius of convergence of this power series is 7/2; this follows from the for-
mula tan x
=
=
sin x/cos x and the results of §3
where
EXERCISES B
1. Derive formula (13) by comparing the coefficients of the appropriate power series.
2 (a) Show that the function (sin r)/r satisfies the DE u,, + (2/ryu, + u = 0 and the
initial conditions u(0) = 1, u’(0) = 0.
(b) Find another, linearly independent solution of this DE.
DE (8) of the text into ugg + (cos #)ug + Au = 0. What is the self-adjoint form of
this equation?
*5 Find conditions on the constants A, ..., F necessary and sufficient for the DE (Ax?
+ Bx + Chu” + (Dx + E)u’ + Fu = 0 to have a polynomial solution of degree n.
Show that if y’ = 1 + 9°, then y” = 2y(1 + 9%), y” = 2(1 + 99(1 + 3y%), and
y” = By(1 + y°)(2 + By’).
Show that any function that satisfies y’ = 1 + y* is an increasing function, and that
its graph is convex upward in the upper half-plane. [HinT: Use Ex. 6.]
Derive the coefficients 1/3, 2/15, 17/315, and 62/2835 of the series (**) of the text.
Show that, if y’ = 1 + y%, 9” = 89(2 + 15y? + 15y*)(1 + 9°).
5 ANALYTIC FUNCTIONS
in the open interval (x),x2) when, given any x in this interval (i.e., satisfying
x; <_X9 < X9), there exist coefficients p,p),po, and a positive number 6
such that
The numerical values of 6 and the coefficients p, will, of course, depend on xp.
Likewise, a real function F(x,y) is analytic in a domain of D of the real
xy-plane when, given (X9,yo) € D, there exist constants 6, (j,k = 0,1, 2,...)
and 6 > 0 such that
M M
(18) G(x,y) =
x
1-2
“£8 HK ayf
j=0 k=0
(
1-—
A I k
This series converges in the rectangle |x| < H, |y| < K and defines an analytic
function in this rectangle.
Analytic functions of three and more variables are defined similarly.
_ (anar*F
=
‘jk
dx/Ayjf (x0290)
If F and G(x,y) = Dye Gulx — xo¥(y — yo)* are any two power series expansions
about the same “center” (%o,jo), moreover, then their power series can be
added, subtracted, and multiplied termwise within the intersection of their
domains of convergence. Worth noting is Cauchy’s formula for the product h(x)
= f(x)g(x) of two analytic functions f(x) = Lf» a,x" and g(x) = Lio bx". This
formula is
The series diverges for all x with |x — x9| > R. The inierval of convergence of
(17) is the interval (x9 — R, x9 + R).
For functions of a complex variable, the radius of convergence of the series
(17) is still determined by Eq. (19). The series is convergent in the circle of con-
vergence |x — x9| < Rand divergent if |x — x9| > R; it defines a single-valued
analytic (or holomorphic) complex function inside its circle of convergence. When
R = 00, the power series (17) defines an analytic function for all x, real or com-
plex; such functions are called entire functions.
/
(20) u” + u 0, c>0
C—-x (C— x”
to Euler’s homogeneous DE
du A du B
(20) 0
dg? &dé Pe”
already discussed in Ch. 3, §2.
To solve (20), try the function u = & = (C — x)’. This satisfies (20) if and
only if v is a root of the indicial equation of (20’,
vpy— 1)-—Av+B=0
When B < 0, this indicial equation has one positive root and one negative root
—p. Hence, (20) has two linearly independent real solutions, given by the bino-
mial series
>
2 Cc
+ Courant and John, pp. 542-544, 555; Widder, pp. 303-306 and 318-320. For a more complete
discussion, see K. Knopp, Theory and Application of Infinite Series, Dover, 1956.
6 Method of Majorants 113
and a like series with vy replaced by —y. When » is a nonnegative integer, a poly-
nomial solution is obtained. Otherwise, the radius of convergence of the series
is C, the same as that of the power series expansions of the coefficient-functions
p(x) =
(C —x) “(OP +E 8]
and q(x) = B/(C ~ x)* of the DE (20).
6 METHOD OF MAJORANTS
If one keeps in mind the results of §4, one can show quite easily that the
formal power series solutions of (4), obtained by the Method of Undetermined
Coefficients of §2, have for all choices of a9 and a, radii of convergence at least
as large as the smaller of the radii of convergence of the coefficient functions.
To prove this, one uses an ingenious method due to Cauchy, the so-called
Method of Majorants.
A power series La,x* is said to be majorized by the series LA,x* if and only if
|a,| << A, for all k = 0, 1, 2, 3, .... By the Comparison Test, the radius of
convergence of Xa,x* is then at least as large as that of XA,x*, and all A, are
positive or zero. Therefore, we say that the DE
majorizes the DE (A) if and oniy if P, = |p,| and Q, = |q|, for all &.
In particular, the choice of coefficient-functions
LEMMA 1. Let the DE (22) majorize the DE (4), and let Df c,x* be the formal
power series solution of (22) whose first two coefficients are |agland |a,|. Then ¢, =
=
Proof. For the DE (22), the coefficients of formal power series solutions sat-
isfy, by (8) with p,
=
=
Pr Gh = —Qe
Hence,if ¢)
= |a,|,¢, = lai|,.-. 5 = Jal,
it follows that c,.; = |a,41|, as
stated. This is because a,,, is given for n = 1 by the display (8), like (*) above,
114 CHAPTER 4 Power Series Solutions
with each (positive) term replaced by one having at most as great an absolute
value. The lemma follows by induction on n.
Now let x, be any number whose absolute value |x,| = C is less than the
smaller of the radii of convergence of the two series (7). Then p,x} and q,x* are
uniformly bounded} in magnitude for all k, by some finite constant M. Hence
we have
This implies that the power series for p(x) and q(x) are both majorized by the
geometric series
we
x MC
»
k=0 (Cc
_—
~ (C—x)’
for some
M > 0,C
> 0
k
MC?
-uya+(2)
(C — x)
4
MC MC?
(23) =
uw+
(C —x) (C—
x
But, as in Example 8, one solution of this DE is the function
om =[1-( _
Cc IP
where —p is the negative root of the quadratic indicial equation of (23). Again
as in Example 3, this equation is
pe + 1)x?
(24) d(x) = ( 1--—
Cc
)oa+ 2C?
+
+ This is because, if a series is convergent, its n-th term tends to zero as n — ©0,
6 Method of Majorants 115
Now apply the foregoing lemma. Each solution of (4) is majorized by K times
the solution ¢(x) of (23), provided that
aC
(24’) k= max{ aol |
But K@(x) has the radius of convergence C. Hence, by the preceding lemma, the
radius of convergence of the series (6) is at least C = |x,|. This proves the
following result.
THEOREM 2. For any choice of ay and a), the radius of convergence of any power
series solution defined by the recursion formula (8) is at least as large as the smaller of
the radii of convergence for the series defining the coefficient functions in (4).
We now recall (§5) that power series can be added, multiplied together, and
differentiated term-by-term within their intervals (circles) of convergence. It fol-
lows from Theorem 2 that when applied to power series defined by (8), the three
equations displayed in §2 between formulas (7) and (8) are identities in the com-
mon interval of convergence specified. Hence, the power series defined by (8)
are solutions of (4), and we have proved the following local existence theorem.
EXERCISES C
1 Let Da,x* have the radius of convergence R. Show that, for any r < R, the series is
majorized by L(m/r*}x* for some m > 0.
Using Ex. 1, prove Cauchy’s formula (19).
Prove that, unless v is a nonnegative integer, the radius of convergence of the bino-
mial series (21) is C.
Using the symbols A, B, C to denote the series La,x*, Db,x*, Cex, and writing
A < B to express the statement that series A is majorized by series B, prove the
following results:
(a) A « Band B < Cimply A « C.
(b) A « Band B < Aimply A =B.
(c) If A « B, then the derivative series A’: Dka,x* and B’: Lkbyx* satisfy A’ « B’.
(d) IfA « BandC
« D, thenA + C «B+ Dand AC < BD.
(a) Obtain a recursion relation on the coefficients a, of power series solutions La,x*
of Pearson’s DE y’ = (D + Ex)y/(A + Bx + Cx, A # 0.
(b) What is the radius of convergence of the solution?
(c) Integrate this solution by quadratures, and compare.
*7 Extend the Method of Majorants of §6 to prove the convergence of the power series
solutions of the inhomogeneous DE u” + p(x)u’ + q(x)u = r(x), when the functions
116 CHAPTER 4 Power Series Solutions
p,q, 7 are all analytic. [HinT: Show that the DE is majorized by setting p(x) = —MC/
(C — x), q(x) = —MC/(C — x)’, r(x) = M/(C — x), for some finite M > 0, C > 0.]
*8 Let the coefficients of u
=
=
Lax" satisfy a recursion relation of the form
Gy+\/a, = P(k)/Q(k + 1), where P and Q are polynomials without common factors
and Q(0) # 0. Show also that « must satisfy a DE of the form
To illustrate the fact that properties of solutions of DEs can often be derived
from the DEs themselves, we will now study the trigonometric DE
(25) y’ ty =0
2 4 3 5
x x
First, by the chain rule for differentiating composite functions, the function
S(a + x) is also a solution of the DE (25). Therefore (Ch. 2, Theorem 2), this
function must be a linear combination of S(x) and C(x):
Furthermore, if we write f(x) = S(@ + x), then f(0) = S(a) and f’(0) = Cia).
But if we differentiate the right side of (26) and set x 0, we find that
=
=
4 0 Vvi-#
sin (5 + *) = —_—
Va
(sinx + cosx) = cos(; - *]
In particular, sin (7/2) = (2/V2)/ V2 = land, therefore, cos (4/2) = 0. Using
the addition formulas again, we get the formulas sin (7/2 + x) = sin (r/2 —
x), cos (x/2 + x) = —cos (#/2 — x), sin (m7 + x) = —sin x, cos (a + x) =
—cos x and, finally, the periodicity relations cos (24 + x) = cos x, sin (2x + x)
= sin x.
*8 BESSEL FUNCTIONS
mula (13) expressing J,(x) algebraically in terms of Jo(x) and its derivatives. In
this section, we shall derive many other useful facts and formulas involving Bes-
sel functions from the results proved in §3. We emphasize that all of these for-
mulas can be derived from their defining DEs (10), the fact that /,,(x) is analytic
at 0, and the choice of leading coefficient in formula (12).
Specifically, one can prove all the properties of the Bessel functions of inte-
gral order from (10) and the recursion relations (13). For example, one can
obtain such useful formulas as
fetdx = ZR + y= SUR+W
More generally, we can obtain useful expressions representing, in closed form,
integrals of arbitrary polynomial functions times Bessel functions and products
of Bessel functions. The basic formulas are (13) and
(27c) Qfx*Jo
Ji dx = —x*]J§ + kfx* YG dx
(27d) Sx*(J§ — Ji) dx = x'Jo
fy — kh — DSx* Yo fy dx
(27e) Sx*t& + JB + & — DUP de = x13 + JP)
—1
4n?
(28) vo”+ [1- 4x”
Jono
+ G. N. Watson, Bessel Functions, Chapter 8. Cambridge University Press 1926; A. N. Lowan et al.,
J: Math and Phys. 22 (1943).
8 Bessel Functions 119
The oscillatory behavior of nontrivial solutions of the Bessel DE (10), for large
x, can now be shown, using the Sturm Comparison Theorem (Ch. 2 §4). When
applied to (28), this result shows that, for large x, the distance between succes-
sive zeros of Jy(x) is inferior to + by a small quantity (at most 7/8x*), while that
between successive zeros x, and x,+, of J,(x) exceeds x by about n*1/2x° ifn =
1. Also, since J; = —Jo, there is a zero of J, between any two successive zeros
of Jo
Setting o(x) v? + vo? = x(Ji t+ Jr) + JnJn + J7/4x,it also followsfrom
(28) that o’(x) (4n® — 1)uv’/2x?. Since |2uv’| < v? + v” = a(x), there follows
Kote)
(28’) lo’(x)| = K,
=
=
2 a(x) > 0
Using the Comparison Theorem of Ch. 1, §11, we get from (28’)
The General Solution. The general solution of the Bessel DE of zero order
is, setting W(x)
=
=
e */* = 1/x in formula (13) of Ch. 2
2
(29) Zo(x=)Jo(x) [4 + af |
But a straightforward computation with power series gives
2
1 x
—+—-
5x4
1
4
Jo") 2 32
From this formula, substituting back into (19) and integrating the resulting
series term-by-term, we see that the general solution Zo(x) of the Bessel DE of
zero order is
5x*
Zo(x)=Jo(x) l4 +B (Ins +74 OH
(30)
128 ‘)
It follows that every solution not a constant multiple of J)(x) becomes logarith-
mically infinite as x | 0, since B # 0. For further information, see Ch. 9, §7
120 CHAPTER 4 Power Series Solutions
is called its generating function. When the series on the right converges in an
interval, this defines a function g(x) there; otherwise, the infinite series is just a
formal power series. In many cases, useful information can be obtained about a
sequence {a,} by studying its generating function.
Likewise, given a sequence of functions F,(r), the function defined by the
power series Lt"F,,(r) is also called the “generating function’’ of the sequence.
Thus, the generating function of the sequence of Legendre polynomials is
The same phrase is used when the sum is taken over all integers; for example,
the Bessel functions of integral order have the generating function
aw
EXERCISES D
1. Define E(x) as in §1, by the DE & = E and the initial condition E(0) = 1. Prove in
turn, justifying your arguments by referring to theorems, that
(a) E(x) = Exo x*/(k!) (b) Ela + x) = E(@E(x)
(c) E(—x) = 1/E(x) [Suggestion: Show that for any a, E(a + x)/E(a) satisfies the
conditions defining E(x).]
Using methods like those of §7, establish the following formulas (cf. Ex. 4):
(a) cosh? x — sinh? x = 1 (b) cosh (—x) = cosh x
(c) sinh (—x) = —sinh x (d) sinh (x + y) = sinh x cosh y + cosh x sinh y
(a) Show that sinh x + cosh x satisfies the conditions used to define E(x) in Ex. 1.
*(b) Using this result, and the formulas of Ex. 3, show sinh !(x) = In(e +
Vx"? + 1).
Prove formulas (27a) and (27b) in detail, expanding on the remarks in the text.
Prove formulas (27c)~—(27e) similarly.
9 First-Order Nonlinear DEs 121
Establish the identities for Bessel functions of integral order in Exs. 7-10.
k
12 Show that (sin 7)/Vr= Vr Jo”) is a constant multiple ofJ;,.(7). (Cf. Ex. A.)
*13. (a) Show that the real and imaginary parts cos(r sin 6) and sini sin 6) of
e? = e808 satisfy Vu + u = 0.
(b) Show that e? = e"""2,t = oy = rsin 0.
(c) Show that the functions F,(r) in the Laurent series expansion
gt? y OF,(7)
a(a + 1) x?
M(absx) = 1+ 2x +
b b(6 + 1) 2!
+
aa+1)---@tn)x"*
+
bb+1)---O+n)
x!
*15, (a) Show that u(r,0,a) = cos [r cos (0 — a@)] satisfies V7u + u =
=
0 for all a, and
(32)
®
dx
= F(x,y) = > ooney
y=0 &=0
The DE
y 1 + y* of Example 7 is one of the simplest nonlinear such DEs; we
refer the reader back to §4 for a preliminary discussion of how to solve this
particular DE
In this section we will explain how to solve a general DE of the form (32) by
the same method. Namely, we substitute into the DE (32) the formal power
series
assuming that we are looking for the solution of (32) satisfying the initial con-
dition y(0) = 0, which we can always do bya translation of coordinates. Accord-
ingly, setting
and so on. The expression on the right side of each of these equations is a poly-
nomial with positive integral coefficients. Equations (34) can be solved recur-
sively, giving the formulas
(10 + 5o0b01)
a= boo, an
=
2
and so on. When we substitute the series (33) for y into the series (32) for F(x,y)
the coefficient of x” is a sum of products of factors b, (with j + k < h) times
polynomials of degree k in the a, obtained by raising the series (33) to the Ath
power. The coefficient of x" on the left side of (32) is, however, (h + 1)a,,1, by
(33’). Equating coefficients of like powers of x, we have, therefore,
where the coefficients of q, are positive integers. Substituting for a,, >» &,
already available formulas, we obtain
THEOREM 4. There exists a power series (33) which formally satisfies any analytic
first-order DE (32). The coefficients of this formal power series are polynomial functions
of the bj, with positive rational coefficients.
The preceding formulas can also be obtained in another way. Let y = f(x) be
the graph in the (x,y)-plane of any solution of the DE y’ = F(x,y), and let u(x,y)
be any analytic function in a domain containing this graph. Differentiating with
respect to x along the graph, we get the formula
du Ou ou _ du du
ay
—_—_—a=
dx ax ( N dy dx Ox dy
(35)
a
= (2 + F(x,y)2)
dx”
d”*ly 0
+F 2) [F(x.y)]
-(
(a+)
(35’)
—_—
J dxt*!
Ox
EXERCISES E
In Exs. 1-7, calculate the first four nonzero terms of the power series expansion of the
solutions of the DE indicated, for the initial value (0) = 0.
9, y =] + xy?
yuxty
y = glx) = Ldyx* 4. y = gy) = xd,9"
yf = xy? + yx? +1 6. ¥ =1+%
y = cos
Vy +1
Calculate explicitly the polynomial p, = a, of Theorem 4.
Compute the first five polynomials p,, of Theorem 4 when F(x,y) = 5(x)y + c(x) in
(30).
124 CHAPTER 4 Power Series Solutions
10. Apply the same question for the Riccati DE y’ + y? = b(x)y + e(x).
11. Show that the DE y’ = x” + y” has a solution of the form LP a,x**~! with all a, > 0.
For a, = 1, compute ag, as, a4.
12. (a) From the DE y’ = 1 + xy, prove that the coefficient a,,, in the expansion tan x
= La,x", ag = 0, satisfies (n + 1)a,4, = ORZ} aa,-1. [Hint: Differentiate y* using
the binomial expansion of (uv)™.]
(b) Compute the first five nonzero coefficients, and compare with those obtained by
solving for y recursively from x = y — y°/3 + 9°/5 — ..., the series for x =
=
arctan y.
Ww
y
=
=
14 For the DE y’ = 2y/x and the initial condition y(1) = 1, calculate the first four terms
of the Taylor series of the solution.
10 RADIUS OF CONVERGENCE
(36) =
is satisfied by the ratio v(x) = u’(x)/u(x), if u(x) is any nontrivial solution of the
second-order linear DE (4). Conversely, if v(x) is any solution of the Riccati DE
(36), then the function u
=
=
A-1
h-1 h
This is, of course, just the special case of formula (34’) corresponding to setting
F(x,y) = —y? — XP party — LP yx’
We shall now return to the general case. Let F be any function of x and y
analytic in some neighborhood of (0,0). This means that F can be expanded into
a double power series
where b, are given real numbers, and the series is convergent for sufficiently
small x and y. We shall show that the series (33) referred to in Theorem 4, and
defined by formulas (34)-(34’)-(34”) has a positive radius of convergence.
to the terms of the series (37). Comparing with the double geometric series men-
tioned in §6
M
“o> ez
(38) G(x,y) =
J j=0 k=0 (
HK*
(
1-—
It
126 CHAPTER 4 Power Series Solutions
and applying the Comparison Test, we see that the series (37) is absolutely con-
vergent in the open rectangle |x| < H, |y| < K, and can be differentiated
there, term-by-term, any number of times.
The preceding remarks have the following immediate consequence.
COROLLARY. [If the power series (33) of Theorem 4 has a positive radius of con-
vergence, then the function which it defines is an analytic solution of the DE (30) for
the initial condition (0) = 0.
We will now complete the proof of an existence theorem for analytic (normal)
first-order DEs by showing that the series (33) has a positive radius of conver-
gence. This is again shown by the Method of Majorants, which we now extend
to functions of two variables.
Consider the power series
ys: GX — Ax?
+ age + ---
one says that the formal power series G majorizes the formal power series F when
(39) F<G.
LEMMA 1. Let F, G, H be any three formal power series. Then F < Gand G &
F imply F =G, and F < Gand G « H imply F « H.
11 Method of Majorants, II 127
LEMMA 2. If F and G are formal power series and if F < G, then F converges
absolutely at any point (x,y) if G converges at (|x|, |y|).
The crucial result for the proof of convergence of the formal power series
(33), obtained from Theorem 4, is the following.
LEMMA 3. Let F < G, and let fand g be the formal power series (without constant
terms) obtained by solving y’ = F (x,y) and y’ = G (x,y) formally, as in Theorem 4,
for the initial condition y(0) = 0. Then g majorizes f (that is, f < g).
for all h; hence, the absolute value of each coefficient a, is less than or equal to
the corresponding coefficient of the formal power series g, q.e.d.
THEOREM 5. Let F (x,y) be analytic in the closed rectangle |x| =< H, |y| = K,
where H and K are positive. Then the formal power series solution (33) of the DE (32)
has a positive radius of convergence.
Proof. The power series for the function F is convergent at (H,K); as in §9,
it follows that for some finite M = max |b,H'K"|,
That is, the formal power series F is majorized by the double geometric series
(38):
cw
K
=
=
>
pk=0
HK
xy
k
.
This series is the product of two geometric series, each absolutely convergent if
|x| <H, |y| < K. Therefore, it is also absolutely convergent in this rectangular
domain, and defines an analytic function there.
Furthermore, the DE y’ = G(x,y) can be solved in closed form by separation
128 CHAPTER 4 Power Series Solutions
where the principal values of the logarithmic and square root functions are
taken, corresponding to the usual expansions of the functions 1+
¢ and
In (1 + 4 in power series with center at ¢ = 0. The radius of convergence is
given by the equation (2MH/K) In [1 ~— x/H] = —1, or
(41) R= HU — e“ KM)
since the binomial series for the radicand in (40) converges so long as
(2MH/K)|In [1 — («/H)]| < 1. This completes the proof.
By Theorem 6 of Ch. 1, whose hypotheses are satisfied since F is continuously
differentiable, the solution satisfying the initial condition f(0) is unique. This
proves the following result.
=
u + iv and
z
—
=
x + iy refer to dependent and independent complex
variables.
Similarly, let
x
rectangle, but the four-dimensional product of two discs. Though this domain
is harder to visualize than a rectangle, it has the advantage that Cauchy’s integral
formulas hold on it: the constant M in (41) is given explicitly byt
M = sup |Fi(z,w)|.
obtained has coefficients ¥,,,, which are polynomials in the b,, (because each b,,
only affects the 6, with k = h) with positive coefficients. This series formally
satisfies (43).
As in the proof of Theorem 5, this series is majorized by the power series
solution without constant term of the DE
Integrating, we get
(44)
1-—
z——
2K
= -—-MHI|n
(3)
t See, for example, Picard, Vol. 2, p. 259.
130 CHAPTER 4 Power Series Solutions
Zz
=
=
EXERCISES F
For the power series expansion of each function defined in Exs. 1-4, determine the
domain of convergence:
8 Iff <GandP<«Q,thenF+P<G+Q,.
11 If F < G, then 0F/dx « dG/dx (interpret the derivatives formally). Is the converse
true?
12 Obtain even and odd power series solutions of the DE w” + iw = 0, and interpret
the solutions.
14 Do the results of Exs. 8~11 hold for complex powerseries? Justify your answer.
CHAPTER 5
PLANE AUTONOMOUS
SYSTEMS
1 AUTONOMOUS SYSTEMS
This and the next three chapters will be concerned with systems of first-order
ordinary DEs in normal form. By this is meant a set of equations
dx,
= Xy(x), . » Xp} 0)
dt
(1)
ax,
= X01, +» Xq; b)
di
i
dx
(2) — = X(x,Z)
d
dx,
(3) =~ = X,(x1, Xn), 2=1,...,n
d
The characteristic property of autonomous systems is the fact that the functions
X, do not depend on the independent variable ¢. When this variable is thought
of as representing time, autonomous systems are thus time-independent or
stationary.
In vector notation, the autonomous system (3) reduces to
dx
Co) — = X(x)
d
To every autonomous system (3) there thus corresponds a unique vector field
X(x) in Euclidean n-space, and conversely. Throughout this chapter, we will con-
sider only vector fields that are of class @', and hence satisfy a Lipschitz condi-
tion in every compact domain. As will be shown in Chapter 6, this implies that
one and only one solution x(é,c) of the autonomous system (3) satisfies the initial
condition x(0)
~
=
c, and that this solution depends continuously on c.
When n = 3, the autonomous system (3) can be imagined as representing the
steady flow of a fluid in space: at each point x in a region of space, the vector
X(x) expresses the velocity of the fluid at that point in magnitude and direction.
The flow is called steady because its velocity depends only on position and does
not vary with time. The solution x(é,c) of the autonomous system (3) for the
initial “value” ¢ then has a simple physical interpretation: it is the trajectory (path,
orbit, or streamline) of a moving fluid particle, whose position (initially at c) is
given as a function of the time ¢.
When the preceding path x(i,c) is considered as a set of points (that is, as a
geometric curve), without reference to its parametric representation, it is also
called a solution curve of the autonomous system (3), or of the associated vector
field X(x). If x(t,c) is a solution of the autonomous system (3), then so is
x(t + a,c) for any constant a; this can also be interpreted as the path of a particle
that passed through the point ¢ at time ¢ = a.
simpler notation as
dx 2
=
(x,9); = Y(x,9)
— =
dt d
dx ]
(4) =
dt 3
d
are evidently x
=
=
2,7
=
=
Note that the solution curves (“trajectories”) of any autonomous system are
endowed with a natural sense or orientation, the direction of increasing ¢. This is
indicated in drawings of solution curves by marking on them arrowheads point-
ing in this direction. See Fig. 5.1, which depicts sample (oriented) solution
curves of the system (4).
In Fig. 5.1, the origin (0,0) is evidently a very special point: integral curves
emanate from it both horizontally and vertically. This is possible only because
the vector field (X,Y) = (mx,ny) reduces there to the null vector 0 = (0,0), whose
direction is indeterminate. Such points are of particular importance for the
study of autonomous systems; they are called critical points.
DEFINITION. A point x = (x, »X,) where all the functions X, are equal
to zero is called a critical point of the autonomous system (3) and of the associ-
ated vector field X(x).
Ifx = c isa critical point of (3), then the functions x,(!) = ¢c, defineatrivial
solution x(é)
=
=
c of (3), which describes not a curve but just a point. In the
2 =
dx
(5) —_—
= X(x,y),
—_
= Y(x,9)
dt di
We then speak of a plane autonomous system. The plane autonomous system (5)
is evidently equivalent to the first-order DE
dy _ Y(x,9)
(5’)
dx X(x,9)
wherever X(x,y) # 0. The main advantage of the parametric form (5) is that
points X(x,y) = 0 of vertical tangency of the solutions of the DE (5’) are no
longer singular points of the corresponding plane autonomous system (5). Like-
wise, the solution curves of (5) are just the integral curves of the quasilinear DE
dx dy
(6) —_— = —
= %
da di
whose solutions are the function-pairs x = r cos(é + ¢), y = rsin(¢ + ¢), where
y and ¢ are arbitrary constants. The graphs of these solutions are concentric
2 Plane Autonomous Systems 135
circles, with center at the origin. The solutions of the corresponding first-order
DE
(6/) 9. _=
x
dx J
are the functions y = + Wr? — x®, which are defined only for |x| < |r|.
Whereas the function —x/y is undefined where y = 0, the functions X(x,y) =
—y and Y(x,y) = x in the system (6) are defined throughout the plane. This gives
the system (6) an obvious advantage over the DE (6’).
Referring to the definition of Ch. 1, §12, we see that the circles x? + yx =f
form a regular} curve family in the ‘‘punctured” xy-plane, the critical point of
(6) at the origin (0,0) being deleted. In Ch. 6, §11, it will be shown that this is
true of plane autonomous systems in general.
Plane autonomous systems have the following interesting relation to level
curves (cf. Ch. 1, §5).
dx ov
(7) =. 5)
dy — —
(x,y)
dt dt
dV_aV dx AV dy dV aV av av
=. eee
ee ee
dt Ox dt dy dt ax oy ov Ox
and so V[x(/),y(] = constant. Observe that the associated steady flow is diver-
gence-free or area conserving, because
ov -9d Vv _ &Vv a
aiv(Oy’ —
Ox ~ Axdy 7 OyOx ~
In fluid mechanics, such a steady flow (7) is called incompressible, and Vis called
its stream function.
The representation (7) also reveals the level curves of Vas the solution curves
of dx,/dt = 0V/dx,,
7 = 1,. »n—that is, in vector notation, of dx/dt = grad
+ Note that this regular curve family is not normal in the plane, whereas the graphs of the function
=
— x* form a normal curve family in the upper half-plane y > 0.
136 CHAPTER 5 Plane Autonomous Systems
The main advantage of the parametric representation (7) over the normal
form
ay __ OV/dx
(7/)
dx OV/dy
considered in Ch. 1, §6, is the following. Whereas the solution curves of (7’)
terminate wherever 0V/dy vanishes, those of (7) terminate only where the func-
tion Vhasacritical point (maximum, minimum, or saddle-point) in the sense that
grad V = 0. This happens exactly where the autonomous system (7) has critical
points.
If we set V = ~(x? + °)/2 in Theorem 1, we get the system (6) of Example
2, having circular streamlines. 1f y(x,y) is nonvanishing, then the system dx/di
= —yp, dy/dt = xp also has circles for solution curves, and we can construct a
wide variety of autonomous systems having the same solution curves in this way,
as has been noted before.
Another illustration of Theorem 1 is obtained by setting
and letting u(x,y) = x*y*. Evaluating (7), we get the following example.
d. y
(8) —_ =
=
dy - — (2x? — y) _
dx x(2y> — x9) ()
is homogeneous of degree zero (see the end of Ch. 1, §7).
We have already defined the ‘‘phase plane” in Ch. 2, $7, as a way of visual-
izing the behavior of solutions of (normal) second-order DEs ¥ = F(x,x). By
3 The Phase Plane, II 137
ye z
F(x,u).
Plane autonomous systems of this special form arise naturally from dynamical
systems with one degree of freedom. Let a particle be constrained to move on a
straight line (or other curve) and let its acceleration x be determined by New-
ton’s Second Law of Motion as a function of its instantaneous position x and
velocity %. Then
(9) x = F(x,%)
xv-plane the phase plane. Since the variables x and mv are conjugate position and
momentum variables, the phase plane is a special instance of the more general
concept of phase space in classical dynamics.
Since (9) is time-independent, it is called an autonomous second-order DE. In
the xv-plane, the second-order autonomous DE (9) is equivalent to the first-
order plane autonomous system
dx _ du dv
(9) — =v— = F(x,v)
a di d.
The integral curves of this autonomous system in the Poincaré phase plane
depict graphically the types of motions determined by the DE (9). Note that the
solution curves point to the right, to the left, or are vertical according as x > 0
(upper half-plane), x < 0 (ower half-plane), or x 0 (x-axis). This is because
=
=
(10) K+ px + qx 0,
=
=
p, q constant
discussed in Ch. 2, §2. The associated autonomous system in the Poincaré phase
plane is
dx dv
(10’) —_>
a= —
po — qx
de” di
a0 pas
(11)
— =
Rk’ sin 6,
dat? £
Here 6 is the (counterclockwise) angle made by the pendulum with the vertical.
The corresponding plane autonomous system in the phase plane (the v@-plane)
is
Since the function sin @ is periodic, the trajectories form a periodic pattern in
the sense that, if [v(),6(/)] is a solution of (114, so is [v(é),0() + 22]. The case
k? =1 is sketched in Figure 5.4.
The solutions of (11’) correspond to the states of constant energy E: v? — 2
cos 6 = 2E (when k? 1). There are two “critical points” in the v6-plane: the
=
=
points (0,0) and (0,7), corresponding to stable and unstable equilibrium, respec-
tively. Near the “vortex point” (0,0), the pendulum oscillates back and forth;
the corresponding trajectories are closed curves, roughly elliptical in shape.
The point (0,7) is a saddle-point; the trajectories v? = 2(1 + cos 6) or v =
=
+2 cos(@/2) that terminate there are called ‘‘separatrices,”’ because they sepa-
rate the closed trajectories from the wavy trajectories v? = 2(E + cos 6) with E
> 1, which correspond to whirling the pendulum in circles.
When the amplitude is small, the DE (11) can be approximated well by
6+ 0
= 0. In this case, the period of oscillation is independent of the amplitude. For
exact solutions, the period increases with the amplitude. We will discuss this
phenomenon in §10 below.
EXERCISES A
1. Find and describe geometrically the solution curves of the following vector fields:
(a) (x, » 2) (b) (ax, by, cz) (c) (y, x, 1) (d) (y, 2, x)
2. Show that the solution curves of the autonomous system
dx d
S~=e-1, =<
dt di
LL
3. Show that the functions xyz and x* + y* are integrals of the system
dx y
dt
= xy", a= —
4. (a) Show that the orthogonal trajectories of the level curves of V = x/(x? + 9°) are
another family of circles. Draw a sketch that displays both families of circles
(b) Same question for V = [(x — 1)? + 9*J/[(@ + 1)? + 97]
The gradient field of a scalar function V(x) is defined as the vector field
The gradient lines of V are the solution curves of the autonomous system dx,/dt =
=
OV/dx
5 6. Vex ty
— 2z 7, V=In[@&e — oa? + ¥/[(x + o)? + 71)
8 Show that a function (x), , X,) of class @! is an integral of the system (3) if and
only if it satisfies the partial DE X, 0¢/dx, + X,, 86/8x, =
=
Show that if 0X/dx = dy/dy and 6X/dy = —dy/dx, the plane autonomous system
(5) is the real form of a single first-order complex analytic DE, and conversely
= [@ — a)? + + 2717?
y > é, arccos a,
= (x
)/ and 6 = arctan (z/y)
are integrals of %
=
=
dV/dx, OV/dy, 0V/dz. Express the integral curves as
intersections of the surfaces defined by the preceding equations
Exercises 11-14 derive some of the main properties of elliptic functions by the methods
of Ch. 4, §7, and give an application to Example 5
11. The elliptic functions u = sn t, v =
=
cnt, w = dnt, may be defined as the solutions
of the autonomous system
du dv
(*) —— = vw, —wu, — = kup
dt
t=K= f FOS
4 Linear Autonomous Systems 141
(b) Prove the following addition formulas, valid with k”’ = V1 — Re: sn(t + K) =
cn t/dn t, en(t + K) = —Rk’ sni/dnt, dnt + K) = k’/dnt. [Hint: Show that the
vector-valued function v/w, —k’u/w, k’/w satisfies (*), and that this vector
reduces ati = K to (1,0,k’.]
(c) Prove that sn(—t) = —sn i, cn(—1t) = ent, sn(t + 2K) = —snt, en(t + 2K) =
—cnt, dn(t + 2K) = dnt.
(d) Show that sn ¢ and cn ¢ have infinitely many zeros, and that the zeros of sn t
separate those of cn t.
13 (a) From the assumptions of Ex. 11, show that the function u = sn ¢ satisfies the
second-order DE
(**) a+ (1 + kyu — 2k = 0
An autonomous system (3) is called linear when all the functions X, are linear
homogeneous functions of the x, so that
d
(12) = ay + ++ + AyXy; i=1, ,n
d
X(x) = Ax
Initial Value Problems. For any autonomous system x’(é) = X(x), the “‘ini-
tial value problem” consists in determining, for each ¢c in the domain of the
vector field X(x), the solution x(f) of the DE that satisfies the ‘‘initial condition”
x(0) = e. We will now show how to solve this problem for any linear plane auton-
omous system (i.e., in the case n = 2).
Any such system has the form
x
2.
(13) + by,
ax cx
+ dy
=
— —_
= =
d d
critical point of the system (13). Since the simultaneous linear equations ax +
by = cx + dy = 0 have no solution except x = y = 0 unless A is singular, we
see that the origin is the only critical point of the system (13), unless ad = bc
(the degenerate case |A| = 0)
To solve the initial value problem for the system (13), it is convenient to intro-
duce a new concept.
THEOREM 2. [f (x(é), y(¢)) is any solution of the plane autonomous system (13),
then x(t) and y(t) are solutions of the secular equation (14) of (13).
Proof. We shall prove that x(é) is a solution of (14); the proof for y(é) is the
same, replacing a with d and 6 with c. The first equation (13) implies x — ax =
by, which implies ¥ — ax = by. From the second equation it follows that
The first equation of (13) will be automatically satisfied, whereas the second will
be equivalent to
which holds by (14). In the same way, if6 = 0 but c ¥ 0 in (13), set x = u —
du and y = cu; (13) follows similarly. In both cases, a second solution can be
constructed from v(é).
In the remaining case that b = c = 0, the obvious formula
The preceding recipes are effective computationally. Thus, to solve the initial
value problem
solution
Moreover, the initial condition x(0) = 1 implies that A = 1, while (0) = x(0)
— x(0) implies that B = 0. The solution of the initial value problem stated is
therefore x
=
=
THEOREM 3. If x(é) is any solution of (12), then every component x,(i) of x()
satisfies the secular equation (16) of (12).
The proof of this result depends on theorems about matrices and so will be
deferred until Appendix A.
space.” But for real linear plane autonomous systems, they arise only when 0 is
a saddle-point. For example, since
—-3 2 2 2 —-3 2 1 1
( —2 2 N Je } wm [
1 1 —2 2 N }=(
2 2
—2 2
way to solve the initial value problem x’({) = Ax. Namely, expand the initial data
x(0) = c into a linear combination of eigenvectors of A: c = La,¢,. Then the
solution of the system x’(t) = Ax for these initial data is
5 LINEAR EQUIVALENCE
The secular equation (14) of a linear plane autonomous system (13) estab-
lishes a clear connection between its solutions and those of an associated (linear)
constant-coefficient second-order DE. As we shall now show, it also throws light on
the rough classification, made in Ch. 2, $7, into “focal,” “nodal” and ‘‘saddle”
points of the critical points of such DEs.
Consider first the case of focal points. Anticipating what will soon be proved,
we begin by considering system of the special form
(18) % = ax
— by, 3 = bx + ay
(18’) ui — 2at + (+ Wu =0
Here —2a is clearly arbitrary, while the discriminant A = 4a? — 4(a2 + 04) =
—46 can be any negative number. Hence (cf. Ch. 2, §7, Case A), all secular
equations of focal point type can be obtained from linear plane autonomous
systems of the special form (18).
In polar coordinates, on the other hand, one easily verifies that (18) reduces to
(i) x
=
=
x =
In the first case, the origin is said to be a star point of (18); in the second, it is
said to be a vortex point of (18). It should be noted that these two “‘degenerate”’
cases (occurring when q = 0 resp. A = 0) were explicitly omitted in the discus-
sion of Ch. 2, §7.
Clearly the phase-plane representation of (18), which is
is less attractive than (18); cf. Ex. 1. Yet the two are linearly equivalent in the
following sense.
dx
(21)
du
_ — = KAx = (KAK™u
di di
under the change of basis associated with the nonsingular matrix K. Thus, in alge-
braic language, linearly equivalent linear autonomous systems are associated
with similart matrices A and KAK™'.
Therefore, the reduction of linear autonomous systems to a standard simpli-
fied (or ‘“‘canonical”) form under linear equivalence amounts to reducing matri-
ces to canonical form under “‘similarity.” We will treat this problem here only
for linear plane autonomous systems
(22)
d
—_ =
ax +by, Om ox + dy
Its solution will throw considerable light onto the classification of critical points
of linear and nonlinear plane autonomous systems into those of focal, nodal,
and saddle-point type.
LEMMA. Linearly equivalent linear plane autonomous systems have the same sec-
ular equation.
Proof. This result follows immediately from (14’), (21), and general identities
of linear algebra. If B = KAK™', then?
|B — dll =
=
|KAK~! — dI| = |K(A — NK"!
|
=
} Birkhoff and MacLane, p. 264. For reduction to diagonal form and Jordan canonical form, see
ibid., pp. 294, 354, For the companion matrix form| _ ___ |, see 1bid., p. 338.
p
146 CHAPTER 5 Plane Autonomous Systems
(The lemma is also a corollary of Theorem 3 unless there are two linearly inde-
pendent equations ui + pu + qu = 0satisfied by both components of all solu-
tions of dx/dt = Ax.)
du d
(23) v,
—_—_— =
qu
— po, =—(a@+d, q = (ad
—bc)
di dt
KAK"! = ( 0 1
~4q —?p
b=axt+djy=axtd 2.y
(25) x = ax,
jy = ay
0,
just as it is for the system
but (25) and (25’ are not linearly equivalent. Every component of every solution
of (25) satisfies 2 = au, but this is not true of the solution (e, te“) of (25/.
The exceptional case arises when the characteristic polynomial of the secular
equation (14) has equal roots, that is, when its discriminant A = p* ~ 4q =
(a — d)? + 4bc vanishes. This gives us the following corollary of Theorem 4.
COROLLARY. Unless the discriminant (a — d)* + 4bc vanishes, two linear plane
autonomous systems are linearly equivalent if and only if they have the same secular
equation.
A. Focal Points. Suppose A < 0, so that the characteristic equation (26) has
distinct complex roots \, = » + iv (v # 0). This is the case g = (uv? + v?) > 0
and 0 <= p® = 4y” < 4q of a harmonic oscillator. We choose the canonical form
(see the Corollary of Theorem 4 and Exs. B1—B3)
for (26), whose solutions are the equiangular spirals r = pe“, 0 = vt + 7 in polar
coordinates, where p = 0 and 7 are arbitrary constants. When p > 0, the spirals
approach the origin (stable focal point); when p < 0, they diverge from it (unsta-
148 CHAPTER 5 Plane Autonomous Systems
ble focal point); when p = 0, they are closed curves representing periodic oscil-
lations (neutrally stable vortex points). See Figures 5.5a and 5.5b.
B. Nodal Points. Suppose that A > 0 and q > 0, so that the roots A = yy,
Hg of the characteristic equation (26) are real, distinct, and of the same sign. We
choose, as the linearly equivalent canonical form,
whose general solution is (ae"", be"). The system is stable when pu, and py are
negative and unstable when they are positive (the two subcases are related by
the transformation t > —t of time reversal). Geometrically, the integral curves
y = cx™ m Ho/Hy, look like a sheaf of parabolas, tangent at the origin, as in
=
=
Figure 5.5c.
C. Saddle-Points. Suppose that A > 0 but g < 0, so that the roots of the
characteristic equation (26) are real and of opposite sign. We again have the
canonical form (26b). But since y, and yy have opposite signs, the integral curves
x"y = c,m = — p/p, > 0 look like a family of similar hyperbolas having given
asymptotes, as in Figure 5.5d. A saddle-point is always unstable.
There remain various degenerate cases and subcases, in which A = 0 or
q = 0. The simplest such case is the exceptional case (25), in which A = 0,
= a? > 0. The integral curves consist of the straight lines through the origin,
and the configuration formed by them is calleda star, as in Figure 5.7e. In the
nonexceptional subcase, we have the canonical form of (25’)
whose integral curves have the appearnce of Figure 5.5f. Such a point is also
called a nodal point, and it is stable or unstable according as a < 0 ora > 0.
The case q = 0, A # 0 corresponds to the phase plane representation of the
second-order DE ¥ + px = 0, p # 0. This corresponds to a rowboat ‘‘coasting”’
on a lake, with no wind and its oars shipped. The boat comes to rest at a finite
distance, in infinite time. The integral curves form a family of parallel straight
lines X + px = constant, as in Figure 5.5g. The origin is a stable (but not strictly
stable) critical point if p > 0, unstable if p < 0.
Finally, the case g = A = 0 reduces to x =
=
y = 0 in the exceptional case
(25) and to the phase plane representation of ¥ = 0 otherwise. The former case
is (neutrally) stable; the latter case is unstable.
EXERCISES B
x=
ax — By, J = Bx
+ ay (a, 6 real)
ix
(e) Star point
Ay]
(f) Nodal point
Figure 5.5
150 CHAPTER 5 Plane Autonomous Systems
x = ax,
I= by, a #0, are y= Cx, Yr"
@ a) -(
2
3 —2
1
I
x
y
) [x(0),9(0)] = (1,0)
0 s0)-(
1
2
2
1
I
x
y
) x0) = 1, 90) = 0
(a) Solve the initial value problem for the DE
1
forgeneral ((0)
d 2 x
x
)
( ( I (
=
=
dt 2 1 y 9(0) b
y
(b) Show that the trajectories are hyperbolas or straight lines through the origin.
(a) Show that the characteristic equation of the system
Show that any linear plane autonomous system (13) with zero discriminant is equiva-
lent to dx/dt = ax, dy/dt = ay, or to dx/dt = ax, dy/dt = x + ay. Describe the asso-
ciated flows geometrically.
dy _xtatf
dx ax + by te
dx _ 3x4 — 12x25? + y4
(x? + yy?
(27)
dt
dy _ 6x°y ~ 10xy?
dt (x? + 9°3/2
+
In this form, one sees at a glance that the rays @ = n/3 are integral curves,
for n = 0,...,5. Other integral curves are sketched in Figure 5.6.
We shall study below how far we can simplify linear autonomous systems by
such coordinate transformations. Our study will be based on a general concept
of equivalence under diffeomorphism, which we now define precisely. Let
Bua
VW.
Figure 5.6 Solution curves at (18).
152 CHAPTER 5 Plane Autonomous Systems
dx; =
(29)
—_
and conversely. In this sense, the autonomous systems (29) and (31) are equiv-
alent. We formalize the preceding discussion in a definition.
are concentric circles, as in Example 2 (x —y, § = x). Yet the two systems are
=
=
+ See Ch. 1, §5. The Jacobian of (28) is the determinant of the square matrix|| 0/,/dx,|| of first partial
derivatives. See Widder, p. 28 ff. Transformations with the properties stated are called
“diffeomorphisms.”
2a, whereas the periods of the solutions of (32) vary like 1/r? with distance from
the origin.
7 STABILITY
The concepts of stability and strict stability, already defined for linear DEs
with constant coefficients in Ch. 2, §3, apply to the critical points of any auto-
nomous system. Loosely speaking, a critical point P is stable when the solution
curves originating near P stay uniformly near it at all later times; P is strictly
stable if, in addition, each such individual solution curve gets and stays arbitrar-
ily near P as ¢ increases without limit. In vector notation, the precise definitions
are as follows.
(i) stable when, given ¢ > 0, there exists a 6 > 0 so small that, if |x(0) — a] <
6, then |x() — al < ¢forallt > 0
(ii) attractive when, for some 6 > 0,
(iii) strictly stable when it is stable and attractive. A stable critical point which is
not attractive is called neutrally stable; a critical point which is not stable is
called unstable.
Proof. If X(x,) = 0 for some x, with 0 < |x,| < 4, then x(#) x, is a solu-
=
=
tion, violating (ii) above. In the same way, if x,X(x,) > 0 [that is, if x, and X(x,)
have the same sign], then the solution with initial value x(0) = (x, + 6 sgn x)/
2 could never cross x,; hence it would also violate (ii). Therefore, condition (ii’)
is necessary for being “‘attractive.”’ It is sufficient since, if 0 < x(0) < 6, then
SUP/5;,xc0}X(*) = —a, < O for all 6, € [0,x(0)]. Hence, we have 0 < x(#) < 6, for
all t = x(0)/a,, proving (30); we omit the details. A similar argument covers the
case —6 < x(0) < 0.
a
Proof. If some eigenvalue A of A has a nonnegative real part, then (12) has a
solution of the form x() = e“f (an “eigensolution’”’ or “norma! mode”’), where
the initial eigenvector x(0) = f can have arbitrarily small length. Conversely,
note that by Theorem 3, every component x,(t) of every solution x(é) of (12)
satisfies the secular equation P,(D)x,(@) = 0, where the roots of the polynomial
equation P,(\) = 0 are just the eigenvalues A, of A. From Theorem 4 of Ch. 3,
we know that every x,(¢) F> 0 if these A, all have negative real parts, which implies
(33). This proves the first statement of Theorem 5’.
To prove the second statement, we extend the concept of solution basis to
vector DEs with constant coefficients. If x(0) = 0 and x’() = Ax, then by
repeated differentiations, we have x0) = 0 for all n. Hence, in Theorem 2,
every x(0) = 0 and, by the crucial Lemma of Ch. 3, §4, every x,() = 0. It
follows from this that the vector form x(é) = Ax of (12) can have, at most n
linearly independent solutions (it will be proved in Ch. 6 that it has exactly n of
them). Calling them u'(#), . , u(t), we see that the general solution of (12) is
x(t) = Lcu'(t) > 0, where |x(é)| < € for all sufficiently large t, uniformly, pro-
vided only that Xc, < 6, some sufficiently small number. The stability condition
(i) above follows.
It follows from Ch. 3, §5, that the conditions for strict stability of the second-
order system dx/di = ax + by, dy/dt cx + dy, are p = — (a + d) > 0,
—
=
q = ad — bc > 0 or equivalently
a + d < 0, ad > bc.
Caution. One should not conclude from Theorems 5 and 5’ that attractive-
ness implies strict stability for all autonomous systems. Indeed, we now con-
struct an attractive critical point of a nonlinear plane autonomous system that
is unstable.t Figure 5.7 depicts sample solution curves.
¢ The authors are indebted to Dr. Thomas Brown for constructing Example 6.
7 Stability 155
2xy on D, U Dy
U Dg
| 2xy/[3 — (41x1)] on Dg
yg — x?
on D, UDy
A|x| — y? — 3x? on Ds
(4lx| — 9? ~ 3x?)/[3 — 4/|xI] on Dg
aq,
(34) df? = F(a 41) =F,q,p) t=1,...,m
°
and conjugate velocity variables p, = dq,/dt are introduced, then (34) defines an
autonomous system of first-order DEs
dq, dp,
(34’) = F(q, p)
a7 Pe di
are closed curves (loops) surrounding the origin or any one of the critical points
for 6 = +2nm. As c > —2, these loops tend to the origin. Consequently, the
origin and its translates 6 = +2nzx,v = 0 are neutrally stable. For c = 2, we
have the separatrix curve defined by v = +2 cos (0/2). From the first of equa-
tions (11%, it is seen that the direction of the motion is from —7 to x for v >
0 and from a to —a for v < 0. Therefore, the critical points v = 0,6 = +(2n
+ 1)x are unstable. These unstable critical points occur when the pendulum is
balanced vertically above the point of support.
EXERCISES C
1 For the following DEs, determine the stability and type of the solution curves in the
phase plane, sketching typical curves in each case:
(a) ét+u=0 (b) i+ a + u =
=
0 (c) i-u+tu=0
dd) @+t+u-—u=0 (c) @+2%+u=0 @ a+ 44+ u=0.
2 2 3 = 3
x= x’,
=
—x*,
x= x*, x =
x =
—x
Show that x = X(x) hasastrictly stable critical point at x = 0 if X(0) = 0 and X’(0)
< 0. [ Hint: x? is a Liapunov function.]
Show that the plane autonomous system ¥ = x — y, § = 4x? + 2y* —6 has critical
points at (1, 1) and (—1, —1), both of them unstable.
Show that the system dx/dt = In (1 + x + 2y), dy/dt = (x/2) — y + (x?/2) has an
unstable critical point at the origin.
*7 Is the system (*) of Ex. B5 strictly stable, neutrally stable, or unstable at (0, 0, 0)?
+ Courantand John, Vol. 2, p. 326. Points where the value of a function is stationary are also often
called “critical points.”
8 Method of Liapunov 157
( Je sin89+ 2sinxcosx+e¢—1
dx
— = -
dt 2
2)
—_=
sin (2x + 3y), = = tan (2x + z)
di di
x? + x = +2ax in D, U Dy
8 METHOD OF LIAPUNOV
where the R; are infinitesimals of the second order. This suggests that the behav-
ior of solutions near the critical point will be like that of solutions of the asso-
ciated linearized system (12). We now show, at least for n = 2, that this is indeed
true as regards strict stability.
THEOREM 6. [If the critical point (0, 0) of the linear plane autonomous system
(12) is strictly stable, then so is that of the perturbed system
+ The symbol O(x? + 9°) stands for a function bounded by M(x” + y°) for some constant M and all
sufficiently small x,y.
158 CHAPTER 5 Plane Autonomous Systems
Idea of Proof. The proof is based on a simple geometrical idea due to Liapu-
nov. Let E(x,y) be any function having a strict local minimum at the origin. For
a small positive C, the level curves E(x,y) = E(0, 0) + C constitute a family of
small concentric closed loops, roughly elliptical in shape, enclosing the origin.
Now, examine the direction of the vector field defined by (35) on these small
loops. Intuition suggests that the critical point will be strictly stable whenever,
for all small enough loops, the vector field points inward. For this implies that
any trajectory which once crosses a loop is forever trapped inside it, because, to
get outside, the trajectory would have to cross the loop in an outward direction.
At such a crossing point, the vector could not point inward, giving a
contradiction.
To make the preceding intuitive argument precise, we define a Liapunov func-
tion for a critical point a of an autonomous system x(t) = X(x) to be a function
E(x) that assumes its minimum value at a and satisfies
E = 2(xx + yy) = QE
In the strictly stable case, » < 0. In (26b), the same Liapunov function satisfies
E = 2(xx + yy) = Qux® + Quey® < 2u,£, in the strictly stable case 0 > pw, =
fy. (By allowing equality, we also take care of the exceptional case of a star
point.) In (26c), the Liapunov function E = x® + ay? satisfies, in the strictly
stable case a < 0
Hence, in the three possible cases of strict linear stability, we have E = 2uE, E
< 2, E, or E S aE, where the coefficient on the right side is negative. It follows
that E < —kE for some k > 0, in every case. Since the quadratic function E(x,
9) is positive definite, we conclude that E(x, y) = K(x? + 9°) for some constant
K> 0.
9 Undamped Nonlinear Oscillations 159
(36) ¥ + g(x)
=0
The key to the analysis of systems (36) is the potential energy integral
THEOREM 7. If q ¢ @! and if xq(x) > 0 for small nonzero x, then the critical
point (0, 0) of the system (36) is a vortex point.t
Proof. For any given positive constant E, the locus v?/2 + V(x) = E is an
integral curve, where V(0) = 0 and V(x) increases with |x|, on both sides of x
= 0. These curves are symmetric under reflection (x, v) F> (x, —v) in the x-axis;
they slope down with slope — q(x)/v in the first and third quadrants and up in
the second and fourth quadrants. For any given small value of EF, the function
E — V(x) has a maximum at x =
=
0 and decreases monotonically on both sides,
crossing zero at points x —B and x = A, where B and A are small and posi-
=
=
tive. Hence each locus v? = 2[E — V(x)] is a simple closed curve, symmetric
about the x-axis.
As the energy parameter E decreases, so does |v| = V2[E — V(x)]; thus, the
simple closed curves defined by the trajectories of (32) shrink monotonically
toward the origin as E | 0. In fact, consider the new coordinates (u, v), defined
Vata
by u + V2V(x), according as x is positive or negative. The transformation (x,
=
=
v) > (u, v) is of class @' with a nonvanishing Jacobian near (0, 0), if ¢(0) exists
and is positive. Hence the integral curves of (37) resemble a distorted family of
circles u? + v? = QE.
The most familiar special case of (36) is the undamped linear oscillator
(40) % + qx
= 0, q=kh>0
+ As in the linear case, a critical point of a plane autonomous system is called a vortex point when
nearby solution curves are concentric simple closed curves.
10 Soft and Hard Springs 161
for which q(x) = kx. The general solution of (40) is the function x = A cos [A(é
— &)], representing an oscillation of amplitude A, frequency k/2m (period 27/h),
and phase to.
In other cases, the DE (36) can be imagined as determining the motion of a
unit mass, attached to an elastic spring that opposes a displacement x by a force
q(x), independent of the velocity ¥. The ratio h(x) = q(x)/x is called the stiffness
of the spring; it is bounded for bounded x if g € @! and q(0) = 0. The case (40)
of a linear spring is the case of constant stiffness (Hooke’s Law). For linear
springs, the formulas of the last paragraph show that the frequency f = k/27 is
proportional to the square root of the stiffness k® and is independent of the
amplitude. We will now show that, for nonlinear springs, the frequency f still
increases with the stiffness but is amplitude-depenValeo
dent in Von
general.
Indeed, the force law (36) implies that x = v = 2(E — V(x)]. Hence, if the
limits of oscillation [i.e., the smallest negative and positive roots of the equation
V(x) = #] are x —Band x = A, the period T of the complete oscillation is
=
=
A
dx
T=
(41)
2 Jos V2[E — Vix)]
The integral (41) is improper, but it converges, provided that q(x) does not van-
ish at —B or A; hence it converges for all sufficiently small amplitudes if g € @'
in the stable case h(0) > 0.
We now compare the periods T and T, of the oscillation of two springs, hav-
ing stiffness h(x) and h,(x) = h(x), and the same limits of oscillation —B and A.
By (39), E = J q(u) du = V(A); hence, E — V(x) = 4 q(u) du. From the stiffness
inequality h,(x) = h(x) assumed, therefore, we obtain,
THEOREM 8. For any two oscillations having the same span [—B, A], the period
becomes shorter and the frequency greater as the stiffness q(x)x increases in (36).
Springs for which h(x) = h(—x) are called symmetric; this makes g(—x) =
—q(x) and V(—x) = V(x), so that B = A in the preceding formulas: symmetric
springs oscillate symmetrically about their equilibrium position. Hence, for sym-
metric springs, the phrase “span [—B, A]’”’ in Theorem 8 can be replaced by
“amplitude A.”
For any symmetric spring, h’(0) = 0; if h”(0) is positive, so that h(x) increases
with |x|, the spring is said to be “hard”; if h”(0) is negative, so that h(x)
decreases as |x| increases, it is said to be ‘‘soft.” Thus the simple pendulum of
Example 5, §3, acts as a “‘soft” spring. We now show that the period of oscilla-
tion is amplitude-dependent, at least for symmetric hard and soft springs.
162 CHAPTER 5 Plane Autonomous Systems
THEOREM 9. The period of a hard symmetric spring decreases, whereas the period
of a soft symmetric spring increases as the amplitude of oscillation increases.
Proof. The period is given by (41); it suffices to compare the periods of quar-
ter-oscillations, say from 0 to A and from 0 to Aj, with A, > A. We write A;
=
=
where h(x) = q(x)/x. The oscillation of amplitude cA for (36) corresponds to the
oscillation of amplitude A for (42); and, since the independent variable ¢ is
unchanged, the periods of oscillation are the same for both. Therefore, it suf-
fices to compare the quarter periods p and p, for amplitude A for the two
springs (36) and (42), respectively. Using Theorem 8, we find that, for y > 0, if
yh(cy) = q(y) = yh(y), that is, if h(cy) & h(y) for c > 1 (hard spring), then we
have p, = p, and so the period decreases as the amplitude increases. Soft springs
can be treated similarly.
EXERCISES D
1 (a) Show that the integral curves of ¢ — x + x® = 0 in the phase plane are the
curves v? — x? + x*/2 =C,
(b) Sketch these curves.
(c) Show that the autonomous system defining these curves has a saddle point at
(0, 0) and vortex points at (+1, 0).
Duffing’s equation without forcing term is ¥ + gx + rx* = 0. Show that, for oscil-
lations of small but finite half-amplitude L, the periodT is
r=avaf™
V2q+ a + sin®6)
Verify Theorem 9 in this special case as a corollary.
(a) Draw sample trajectories of the DE # = 2x° in the phase plane, including ¢ =
+x.
(b) Show that x = 0 is the only solution of this DE defined for all ¢ € (—00, 0).
*4 Show that if (0) = 0 and q’(0) < 0 in (36), the origin is a saddle-point in the phase
plane.
Discuss the dependence on the sign of the constant y, of the critical point at the
origin of the system
w= —v
+ pu’, v=ut
pw
(a) Show that the trajectories of + q(x) = 0 in the phase plane are convex closed
curves if q(x) is an increasing function with g(0) = 0.
(b) Is the converse true?
where ad # bc and A # 0, then the system X = X(x, y), 3 = Y(x, y) is also unstable
there.
_ gels
*=8 2
bv
where the constant v is the “terminal velocity.”’ Sketch the integral curves of this
DE in the phase plane, and interpret them physically.
Show that the plane autonomous system x = y — x°, ¥ = —x? is stable, though its
linearization is unstable. [Hrnt: Show that x* + 2y? is a Liapunov function.]
*10. Show that for the analytic plane autonomous system
the origin is an unstable critical point that is asymptotically stable. [HinT: Study
2
Example 6. To prove instability, show that the ellipse 4y* =
=
x—xX cannot be
crossed from the left in the first quadrant.]
When A(0) is positive, the equilibrium point is called statically stable, because the
restoring force tends to restore equilibrium under static conditions (when
v = 0). The conservative system (36) obtained from any statically stable system
(43) by omitting the friction term xp(x, %), is neutrally stable by Theorem 7.
The differential equation (43) has a very simple interpretation in the phase
plane, as
d
(44) dx {S| =we
=
vp(x, v) — q(x)
ds di
The critical points of the system (44) are all on the x-axis, where x =
=
Vv
=
=
0;
they are the equilibrium points (x, 0) where g(x) = 0 in (43). Since in (43), (0)
= 0h(0) = 0 the origin is always a critical point of (44); unless h(0) changes sign,
there is no other.
We shall consider only the case h(x) > 0 of static stability in the large, which
is the case of greatest interest for applications. For simplicity, we will also
assume that p(0, 0) # 0.
Under these assumptions, the origin is the only critical point of (44). More-
164 CHAPTER 5 Plane Autonomous Systems
d q(x)
— p(x,v) —
— = r(x, v) = —
d v
points to the right in the upper half-plane, where ¥ = v > 0, and to the left in
the lower half-plane. On the x-axis, the solution curves have finite curvature
q(x); they cut it vertically downward on the positive x-axis and vertically upward
on the negative x-axis. Thus, the solution curves have a general clockwise
orientation.
Conversely, any continuous oriented direction field with the properties spec-
ified represents a DE of the form (43) in the phase plane. From this it is clear
that the behavior of the solutions of the DEs of the form (43) can be extremely
varied in the large (see §13). However, the local possibilities are limited.
THEOREM 10. [If p and h are of class @' in (43), and if p (0, 0) and h(0) are
positive, then the origin is a strictly stable critical point of (44).
Proof. Under the hypotheses of Theorem 10, we have that p > 0 near the
critical point (0, 0), and we can write (44) as
An easy computation shows that the linearization of the system (44) has the sec-
ular equation A? + por + h(0) = 0. Since this quadratic polynomial has positive
coefficients, it is of stable type (Ch. 3, §5). Hence, by Theorem 6, the origin is a
strictly stable critical point of (43) when the damping factor p(0, 0) is positive.
The equilibrium point x —
=
d’x
(45)
d(—t)?
+ p(s*d(-t —dx dx
d(—t)
+ xh(x) = 0
we see that the substitutions i > —i, x ~ x, v > —v, of time reversal, reverse
the sign of p(x, %) but do not affect (43) otherwise. Hence, if p(0, 0) < 0, all
solution curves of (44) spiral outward near the origin.
commonly tends to a finite limit. The limiting periodic oscillation of finite ampli-
tude so approached is called a limit cycle.
The simplest DE that gives rise to a limit cycle is the Rayleigh equation,
(46) &— wl — xe
+x =
=
0, up>od
The characteristic feature of this DE is the fact that the damping is negative for
small x and positive for large x. Hence it tends to increase the amplitude of small
oscillations and to decrease the amplitude of large oscillations. Between these
two types of motions, there is an oscillation of contant amplitude, a limit cycle.
If we differentiate the Rayleigh equation (46) and set y = ¥V/3, we obtain
the van der Pol equation
(47) 5 — wl — yy
+ y =0, u>od
This DE arises in the study of vacuum tubes. The sign of the damping term
depends on the magnitude of the displacement y. The remarks about the Ray-
leigh equation made above apply also to the van der Pol equation.
As stated in §11, negatively damped nonlinear oscillators can give rise to a
great variety of qualitatively different solution curve configurations in the phase
plane. For any particular DE of the form (43), such as the Rayleigh or van der
Pol equation with given yw, one can usually determine the qualitative behavior of
solutions by integrating the DE
v do + [vp(x,v) + g(x)] dx = 0
graphically (Ch. 1, §8). More accurate results can be had by use of numerical
integration. With modern computing machines, using the techniques to be
described in Ch. 8, it is a routine operation to obtain such a family of solution
curves. Figure 5.8 depicts sample integral curves for the van der Pol equation
with
» = 0.1, w = 1, and
» = 10 so obtained.
Liénard Equation. General criteria are also available which determine the
qualitative behavior of the oscillations directly from that of the coefficient-func-
tions. Such criteria are especially useful for DFs depending on parameters,
because graphical integration then becomes very tedious. They are available for
DEs of the form
The van der Pol equation is a Liénard equation; moreover, it is symmetric in the
sense that —x(?) is a solution if x(/) is a solution. This holds whenever q(—x) =
—qg(x) is odd and f(—x) = f(x) is even.
One can prove the existence of limit cycles for a wide class of Liénard equa-
tions; we can even prove that every nontrivial solution is either a limit cycle, or
a spiral that tends toward a limit cycle as t > +00. This is true if: (i) xq(x) > 0
166 CHAPTER 5 Plane Autonomous Systems
15
F
‘1 | 10
Q2t
2h
=
A) 5h
eX
—1b —5r
—2+
—2e
—10F-
~4b
1 —-3Fr
—4 -2 —15E
-2 -1 0
for x # 0; (ii) f(x) in (48) is negative in an interval a < x < 6 containing the
origin and positive outside this interval, and
0
(49) [roa f =_
wf!) dx = +00
(50) @+f) +2 = 0 if v0
It follows, since xq(x) > 0, that they can cross the x-axis only downward if
x > 0, and upward if x < 0. Also by (50), between successive crossings of
the x-axis, v(x) is a bounded single-valued function, decreasing in magnitude if
x > bin the upper half-plane, and if x < a in the lower half-plane.
Now, consider the Liénard function
(51)
E(x, v) = afv + FQ)? + Ux) F(x) = f *fle) dx
U(x) = f g(x) dx = 0
+ For a complete proof, see Lefschetz, p. 267, or Stoker, Appendices III and IV.
12 Limit Cycles 167
and the definite integral is negative for sufficiently large oscillations since xq(x)
> 0 and F(to0) = +00, by (49). Hence, every solution curve sufficiently far
from the origin must spiral inward. Therefore, every oscillation of sufficiently
large initial amplitude must spiral inward toward a limit cycle of maximum
amplitude. Similarly, every oscillation of sufficiently small initial amplitude must
spiral outward to a smallest limit cycle.
For the Rayleigh and van der Pol equations, these limit cycles are the same.
Therefore, every nontrivial solution tends to a unique limit circle, which is sta-
ble. The preceding result holds under much more general conditions. We quote
one set of such conditions without proof.
LEVINSON--SMITH THEOREM. In (48), let q(x) = xh (x), where h(x) > 0 and
let f(x) be negative in an interval (a, b) containing the origin and positive outside this
interval. Let q(—x) = —4q(x), f(—x) = f(x), and let (49) hold. Then (48) has a
unique stable limit cycle in the phase plane, toward which every nontrivial integral
curve tends.
EXERCISES E
¥ + (px? — gx + rx = 0
where g and r are positive constants, can be reduced to the van der Pol DE by a change
of dependent and independent variables.
3 2,
uw 2
a@=u-v—w— »
v=utv- v Uv
has a unique critical point, which is unstable, and a unique limit cycle.
(b) Discuss the stability of the related system
wu’,
“a= —u-—vtwt v=eu-vtvtuy
with special reference to oscillations of very small and very large amplitude
168 CHAPTER 5 Plane Autonomous Systems
In the DE ¥ + q(x) = 0, let Vix) = 6 q(u) du, and let g be a continuous function
satisfying a Lipschitz condition. Show that, if V(x,) = Vixg) and Vix) < V(x,) for
x, <x < Xx9, the equivalent autonomous system (44) has a periodic solution passing
through the points (x,, 0) and (x, 0).
_ rn
>
6=1 (polar coordinates)
100
Tr 1
| )a( Q 1
_
=
=
100 Tr
Prove in detail that, for 4 = 1 and the initial condition x(0) = 10, %(0) = 0, the
amplitude of successive oscillations decreases in the Rayleigh DE (46).
Answer the same question for the van der Pol DE, if y(0) = 10, y(0) = 0.
Sketch the integral curves of the van der Pol DE in the phase plane for » = 100.
[Hint: Most of the time, the integral curves are “relaxation oscillations,” near y = 0
ory = 1]
ADDITIONAL EXERCISES
Show that, in the punctured plane x” + y? > 0, two linear systems (13) are equivalent,
provided that their discriminants A, A’ are not zero, and they both have either (a)
stable focal points, (b) unstable focal points, (c) vortex points, (d) stable nodal points,
(e) unstable nodal points, or (f) saddle-points.
=
ps
i Cg Cg
12 Limit Cycles 169
(b) Show that the secular equation (¢) is invariant under any nonsingular linear trans-
formation of the variables x, y, z
(c) Conversely, show that, if the polynomial A® = p,A? — pod + ps has distinct real
roots dj, Ag, As, then the given DE is linearly equivalent to & = 6G = 1,2,3).
*(d) Work outa set of real canonical forms for the given DE, with respect to linear
equivalence, in the general case.
(a)
¥ txt =0 (b) # + x/[%] = 0 (Coulomb friction)
x—@
Let q(x) be an increasing function, with (0) = 0 and q(—L) = —q(L); let Q(x) =
[g(x) — 9(—x)]/2. Show that the period of oscillation of half-amplitude L for ¥ +
Q(x) = 0 is less than that for ¥ + g(x) = 0 unless g(—x) = —g(x) for all x € [0, LZ}.
Show that the DEs ¥ = 2x + sin x and ¥ = 2x are equivalent on (—00, +00) but
that x =
=
x + x® is not equivalent to ¥ =
=
x. [Hint: Consider the escape time.]
10 Showthat,
if a) < ag << +++ <a,andb <by<+++ <b, the DEs % =
=
Il(x —
a,) and % = I(x — 6,) are equivalent.
CHAPTER 6
EXISTENCE
AND UNIQUENESS
THEOREMS
1 INTRODUCTION
dx,
= X1(x1... 5 Xa t)
dt
(1)
dx,
=
n(Xy5 » Xq} f)
dt
For the most part, we shall restrict attention to the existence and uniqueness of
such solutions. But in the later sections, we shall consider more sophisticated
questions, such as the analyticity of solutions and their dependence on the initial
value vector c (1, , ¢,). We shall prove that, as one might expect, this
=
=
The curve x(é) in # defined by any solution of (1) will be called a solution curve
of the system (1).
Note that we can trivially inflate any normal system (1) of n first-order DEs
to a normal autonomous system of n + 1 DEs by the simple device of writing
t = X,+4,- This gives the equivalent system
dx,
—_— =
where X,,,, is the function 1. However, this does not help to prove the theorems
of major interest.
As in Chapter 5, §1, one can view any autonomous system x’(i) = X(x) as
defining a steady flow in the appropriate region # of x-space. Although this
does not help to prove the theorems of major interest, it does make it easier to
visualize their meaning.
Thus, a continuously differentiable function U(x) is called an invariant of the
autonomous system x’(!) = X(x) when U(x(t) is constant for every solution x(t)
of this vector DE—i-e., when & X;(x) 9U/dx; = 0. This means that each solution
curve of (2’) stays on a single level surface U = const., and thus generalizes the
concept of “integral” defined in Chapter 1.
dx dy dz
= —Yz, _ =
— = Xz,
“oy
di
=
x(t) and y = y(é) are solutions, the two
functions
As a familiar special case, consider also the plane autonomous system dx/dt
= N(x,y), dy/dt = —M(x,y) associated with the DE M(x,y) + N(x,y)y’ = 0, The
function U(x,y) is an integral of this system, associated as in Ch. 1, $5, with the
integrating factor p(x,y), if and only if OU/dx = uM and dU/dy = BN.
First-order normal systems (1) provide a standard form to which all normal
ordinary DEs and normal systems of DEs can be reduced. For example, one can
reduce the solution of a normal nth-order DE to the solution of a system of n
first-order normal DEs as follows. Let u(é) be any solution of the given nth order
172 CHAPTER 6 Existence and Uniqueness Theorems
DE,
du du du d’'y
(
=
dx, dx,
= Xe+1> kR=1,...,7 1; = F(x), %2, » Xai)
dt di
Conversely, given any solution of the preceding first-order system, the first com-
ponent x,(¢) will have the other components xo, » %, aS its derivatives of
orders 1,...,” — 1. Hence, substituting back, x,(¢) will satisfy the given nth-
order equation.
In the presertt chapter, we shall use this standard form to develop a unified
theory for the existence and uniqueness of solutions of DEs and systems of DEs
of all orders.
2 LIPSCHITZ CONDITION
In order to make use of vector notation for systems of DEs, we recall a few
facts about vectors in n-dimensional Euclidean spaces. Addition of two vectors
and multiplication of vectors by scalars are defined component-wise, as in the
plane and in space. The length of a vector x = (x), 9, . ; X,) is defined as
Ix
+ yl = |x| + lyl
KY SH xXyy tes
+ xn
is the vector function x’(i) = (xj}(i), x4(é), , x/(#)). The integral f° x(¢) dt is the
vector with components f° x, £) dt, J2 xo(t) dt,..., J x,(é) dt. We shall often make
use of the fundamental inequalityt
Note that both terms on the left side of (4) involve the same value of #.
It can be obtained from this inequality by recalling the definition of the integral {° x(t) dt as a limit
of Riemann sums, using the triangle inequality for each of the Riemann sums, and passing to the
limit on both sides,
174 CHAPTER 6 Existence and Uniqueness Theorems
Proof. Let M be the maximum of all partial derivatives |dX,/dx,| in the closed
domain D. For each component X, we have, for fixed x, y, ¢ and for variable s
Hence, by the mean-value theorem applied to the function X,(x + sy, #) of the
variable s on the interval 0 = s = 1, we have
for some o, between 0 and 1. Squaring, and applying the Schwarz inequality to
the right side, we obtain
Ox;
) (>: inl") < nM’ly|?
Consequently, summing over all components i, we have
Taking square roots, the Lipschitz condition follows with Lipschitz constant nM.
3 WELL-POSED PROBLEMS
+ A set S in n-space is convex when the segment joining any two points of the set S lies entirely within
S. This definition applies both to closed domains D and open regions &.
3 Well-Posed Problems 175
Differentiating o(/), and using the fact that x and y are solutions of the normal
system (2), we get
By the result of Lemma 2 of Ch. 1, §10, it follows that if x (2) = y(a), that is, if
o(a) = 0, then o(f) = 0 [that is, |x(t) — y()|? = 0] for all t = a.
A similar argument works for ¢ < a: replacing ¢ by —?, as in proving Theorem
6 of Ch. 1, we obtain
— = = |o(| = 2Lo()
d(—t
THEOREM 2 (CONTINUITY THEOREM). Let x(t) and y(t) be any two solutions
of the vector DE (2), where X(x, t) ts continuous and satisfies the Lipschitz condition
(4). Then
Applying Lemma 2 of Ch. 1, §10 to o(@), we get o(a + h) < o(a)e*™. Taking the
square root of both sides, we get the desired result.
176 CHAPTER 6 Existence and Uniqueness Theorems
From Theorem 2 we can easily infer the following important property of the
solutions of the DE (2).
COROLLARY. Let x(t, c) be the solution of the DE (2) satisfying the initial con-
dition x(a, c) = c. Let the hypotheses of Theorem 2 be satisfied, and let the functions
x(t, c) bedefined for |c ~ c°| < Kand |t — a| ST. Then:
(a) x(t, c) is a continuous function of both variables:
(b) ife > c°, then x(t, c) > x(t, c°) uniformly for |t — a| < T.
Both properties follow from the inequality (5).
In view of the preceding results, it remains only to prove an existence theorem,
in order to show that the initial value problem is well-set for normal first-order
systems (1). This will be done in Theorems 6~8 later.
EXERCISES A
1 Showthat u = x +» + zandv =
=
x? + y? + 2? are integrals of the linear system
dx/dt = y — z, dy/dt = z — x, dz/dt = x — y. Check that the solution curves are
circles having the line (, ¢, #) as the axis of symmetry.
Reduce each of the following DEs to an equivalent first-order system, and determine
in which domain or domains (e.g., entire plane, any bounded region, a half-plane,
etc.) the resulting system satisfies a Lipschitz condition:
(a) d®x/dt? + x? = 1 (b) ax/dt = xP, (c) dx/d = [1 + @?x/at?)?]'”
Reduce the following system to normal form, and determine in which domains a
Lipschitz condition is satisfied:
dv Qdu 3du
du —_—=
u? + v’, = Quu
dt dt dt dt
Show that the vector-valued function (¢ + be“, —e*/ab) satisfies the DE (2) with
X = [1 — (1/x9), 1/(x, — 4], for any nonzero constants a, 6
State and prove a uniqueness theorem for the DE y” = F(x, y, 9’), with Fe é}” (Hint:
Reduce to a first-order system, and use Theorem 1.]
(a) Show that any solution of the linear system dx/dt = y, dy/dt = z, dz/dt = x
satisfies the vector DE ax/dt = x, where x = (x, , 2).
(b) Show that every solution of the preceding system can be written x = da +
eb cosV3t/2 + csinV3t/2], for suitable constant vectors a, b, and c.
(c) Express a, b, and ¢ in terms of x(0), x’(0), and x” (0).
(a) Find a system of first-order DEs satisfied by all curves orthogonal to the spheres
x? + y? + 22 = Qax — a’.
(b) By integrating the preceding system, find the orthogonal trajectories in question.
Describe the solution curves geometrically.
10 (a) In what sense is the following statment inexact? “The general solution of the DE
4 Continuity 177
s-[eratt}e+[@—wm a+]. =
=
0
12 For which values of a, 8 does the function x%P satisfy a Lipschitz condition: (a) in
the open square 0 < x, y < 1, (b) in the quadrant 0 < x, y < +0, (c) in the part
of the quadrant of (b) exterior to the square of (a)?
13 For each of the following scalar-valued functions of a vector x and each of the fol-
lowing domains, state whether a Lipschitz condition is satisfied or not:
(a)x, +x tess
+ x, (b) x1%9 x, tt (c) y/(e? + 9°) (@) {x in @
|x| <1, (i) —00 < x, < ©, (iii) —00 < x, <0, }x,]| <1, k= 2.
14 Let X(x, é) = (X,(x, 4), . . . , X,(x, 2) be a one-parameter family of vector fields. Show
that X satisfies a Lipschitz condition if and only if each scalar-valued component X,
satisfies a Lipschitz condition, and relate the Lipschitz constant of X to those of the
x,
CONTINUITY
respectively, on a <= t < b, Further, let thefunctions X and Y be defined and continuous
in a common domain D, and let
We now apply the triangle inequality to the right side, and then the Schwarz
inequality to each of the two terms of the last expression. This gives the
inequality
To the first term on the right side we now apply the Lipschitz condition that X
satisfies, to the second term, we apply (6). This gives the following differential
inequality for o:
LEMMA. Let o(f) = 0,a <1 <b be a differentiable function satisfying the dif-
ferential inequality (8). Then
du
(9’) _
%Vu + 2Lu, u=O0
dt
which satisfies the initial condition u(a) = o(a), will have a nonnegative deriva-
tive, and therefore will remain, for ¢ > a, within the half-plane u = o(a).
The DE (9’) is a Bernoulli DE (Ch. 1, Ex. C7). To find the solution satisfying
4 Continuity 179
u(@) = a(a), make the substitution v() = Vu(é). (The square root is well-defined
because u(é) = o(a) > 0.) This gives the equivalent DE
If u(a) > 0, it follows that u(t) > 0 for all later ¢, since the derivative of u is
positive. This gives v(i) > 0, and so we can divide both sides of this DE by v.
The resulting DE is v’ ~ Lv = e, an inhomogeneous linear DE whose solution
satisfying the initial condition v(a) = Vu(a) is the function
for all n > 0. Letting n — 00, we obtain the inequality (9) also in this case.
EXERCISES B
1. Let X and Y be as in Theorem 3, and let x(a) — y(a). Show that |x() — y(@j/{t —
a| remains bounded as t > a.
2. Show that if |OF/dy| < L(x), then any two solutions of u’ = F(x, u) and vo’ = Fix, v)
satisfy |u(x) — u(x)| < |u(0) — v(0)|e,
For the pairs of DEs in Exs. 3-5, bound the differences on [0, 1] between solutions hav-
ing the same initial value (0) = c.
na
3. yy =PandY =1l+tytes- + 2
oa
!
180 CHAPTER 6 Existence and Uniqueness Theorems
a’
4,.—= 6 and at? =
=
—sin 0.
dt®
*7 Show that the conclusion of Theorem 3 holds if only a one-sided Lipschitz condition
(x — y) - (X(x, ) — Xfy, )) S Llx — y|? is assumed for X.
Let X(x, ¢, s) be continuous for |x — c] < K, |t — a| < T, and |s — so] < S, and
let it satisfy | X(x, t, s) — X(y, t, s}| == L|x — y|. Show that the solution x(, s) of x’
= X(x, i, s) satisfying x(a, s) = ¢ is a continuous function of s.
*5 NORMAL SYSTEMS
qr é,
dt
..3 &,
d&>
dt
3 Sn a
dé
n
)
k
=
=
1, , m, in which for each & only derivatives d’é,/df of any &, of orders
p < n(j) occur on the right side.
by the formulas
dé ae, _ a's
1 = &, x =
——
»%3 = >
n]
di di?” at"?
In terms of these new variables, the system (1) assumes the form
—_—=
alg
dé ae dle
&, re oe di} » $2»
* aie)?
s
di" 1
COROLLARY. [If the functions F, of the normal system (10) satisfy Lipschitz con-
ditions in a domain D, then the system has at most one solution in D satisfying given
initial conditions.
Example 2 (the n-body problem). Let n mass points with masses m,; attract each
J
other according to an inverse ath power law of attraction. Then, in suitable
units, their position coordinates satisfy a normal system of 3n second-order dif-
ferential equations of the form
d°x; — *)
mj(%
di? =) jFe
nen
In this notation, the system (10) is equivalent to a first order normal system of
the form (1):
En+3n h=1,...,3n
dé,
= F,(é) = > m(E,—3n ~ Ex-30)/rhton h=3n+1,...,6n
—
dt
Jj
da ap
~—>$ See
Y ay 8
R(s) Ris) Ts)’ ds ~ Ts)
where a, 8, and y = a@ X @ are three-dimensionalf vectors: the unit tangent,
normal, and binormal vectors to a space curve. The curvature «(s) = 1/R(s) and
torsion 7(s) = 1/7(s) are functions of the arc length s; @ = dx/ds is the deriva-
tive of vector position with respect to arc length.
If we let n(s) be the nine-dimensional vector (@,, >, 03, 81, Bos B3, Yi» Yo. Y3)s
the system can be written as the first-order vector DE
dn/ds = Y(n; 5)
K(S)Mis h=1,2,3
Y,(7; 5) = — K(s)m—3 + 7(S)ti+3 h=4,5,6
— 7(5)m—s h = 7, 8,9
If x(s) and 7(s) are bounded, the vector fields Y¥(n; s) satisfy a Lipschitz condition
(4) with L = sup{|x(s)| + |7(s)|}; hence, for given initial tangent direction a(0),
normal direction 8(0) perpendicular to (0), and binormal direction (0) = a(0)
X @(0), there is only one set of directions satisfying the Frenet-Serret formulas.
This proves that @ curve with nonvanishing curvature is determined up to a rigid
motion by its curvature and torsion.t}
+ Widder, p. 101.
ta X B denotes the cross product of the vectors @ and @.
+t This theorem of differential geometry can fail when «(s) is zero, because 8 = x”(s)/|x”(s)| is then
geometrically undefined, so that the Frenet-Serret formulas do not necessarily hold.
6 Equivalent Integral Equation 183
EXERCISES C
2,
dx dy d?x
—7 +sy
0, +xt+y=0
=
—
=—_ =
y
“Ue at* #2
di at?
= > BngXyXp
2
x
=
=
Let a = 2 in the n-body problem (Newton’s law of gravitation), and define the poten-
tial energy as V = — 2,2, mm,/rTy.
(a) Show that the n-body problem is defined by the system
m,a°x, av
=
dt? Ox
(b) Show that the total energy Z m,x2/2 + V(x) is an integral of the system.
(c) Show that the components & m,x,, etc., of linear momentum are integrals.
(d) Do the same as in (c) for the components Z m,(y,z zy,), etc., of angular
momentum.
Show that the general solution of the vector DE d*x/dt® = dx/dt is a + be! + ce,
where a, b, c are arbitrary vectors.
6 Show that, if a(s), B(s) , y(s) are orthogonal vectors of length one when s = 0, this is
true for all s, provided they satisfy the Frenet-Serret formulas.
Show that if 1/7(s) = 0, and dx/ds = a, the curve x(s) lies in a plane. [HiNT: Consider
the dot product ¥ - x.]
is a solution of the vector DE (2) that satisfies the initial condition x (a) = ¢, and
conversely.
Proof. If x(d) satisfies the integral equation (11), then x(a) = ¢ and, by the
Fundamental Theorem of the Calculus, x/() = X,(x(@); ) fork = 1,..., 7, so
that x(é) also satisfies the system (2). Conversely, the Fundamental Theorem of
the Calculus shows that x,(é) = x,(a) + Ji, x{(s) ds for all continuously differen-
tiable functions x(t). If x(é) satisfies the normal system of DEs (2), then
x(Z)
=
=
x(a) + J! X(x(s); 5) ds; if, in addition, x(a) =
=
COROLLARY. The DE (2) has a solution satisfying x(a) = c if and only if the
mapping U of (12) has a fixpoint in @{a, 5).
However, if X(x, ¢) is not defined for all x, the domain of the operator U has
to be determined with care. This will be done in Theorem 8 below.
7 SUCCESSIVE APPROXIMATION
Picard had the idea of iterating the integral operator U defined by (12), and
proving that, for any initial trial function x’, the successive integral transforms
(Picard approximations)
converge to a solution. This idea works under various sets of hypotheses; one
such set is the following.
THEOREM 6. Let the vector function X(x; t) be continuous and satisfy the Lip-
schitz condition (4) on the interval |t — a| = T for all x, y. Then, for any constant
vector c, the vector DE x’(t) = X(x; t) has a solution defined on the interval |t — a|
ST, which satisfies the initial condition x(a) = c.
*|
oo.
e/2
1+¢
g = ——+—
10>
l+¢t
r=
0.5
z°
0.5 1.0 t
Proof. Let M = supy—a)<7r |X(c; #)|; the numberM is finite because contin-
uous functions are bounded ona closed interval. Without loss of generality we
can assume that a 0 and ¢ = a, that is, that the interval is 0 = ¢ < T; the
=
=
proof for general a and for t < @ can be deduced from this case by the substi-
tutions i > i+ aandt—~a
— it.
By the basic inequality (3) for vector-valued functions, the function x'({) sat-
isfies the inequality
(M/L)(La)"
Ix") —x""O| = !
7 Successive Approximation 187
we infer that
M(Lt)"* 1
*(Ls)"
(14) |x") —x")| SL (#) ni L(n
+ 1)!
n+l
GED
i
(7)azo | + 1)!
The left side converges uniformly, by the preceding lemma. By the Lipschitz
condition, [X(x"(s), s) — X(x"(), s)| <= L|x"(s) — x"(s)|, and so the integrals on
the right side also converge uniformly. It follows that they have a continuous
limit X(x(#); #).[ Passing to the limit, we have
x @=e+ f xe (s),
8)ds
EXERCISES D
6. Show that the nth iterate for the solution of y’ = yx such that (0) 1 is the sum
of the first n + 1 terms of the power series expansion of e*/?
For the initial value problems in Exs. 7-10, obtain an expression for the nth function of
the sequence of Picard approximations x U"[x°] to the exact solutions
7. dx/dt ’ x(0) 9. dx/dt
=
=
tx, x(0)
x(0) 0, 1
=
(0)
= =
x(0) 0, 9(0) 1
=
=
= = =
2
For the initial value problems of Exs. 11-13, compute the functions x’, x x” of the
sequences of Picard approximations.
24 2
11 dx /dt =
=
x(0) = 0
12 dx/dt
=
=
15 Let X(x, t) = Ax, where A is a constant matrix. Show that each component of the
nth Picard approximation to any solution is a polynomial function of degree at
most n
8 LINEAR SYSTEMS
(15)
d
~ a;x(t) + bt) lsisn
In this case, we have X;,(x, ft) = Xj.) a;(t)x; + 5,(). In vector notation, the linear
system (15) is written in the form
where A(é)x stands for the matrix ||a,,(¢)|| applied to the vector x, and b stands
for the vector (6, »b,)
When b(t) = 0, the system (16) is said to be homogeneous. Otherwise, it is
called inhomogeneous. The homogeneous system obtained from a given inho-
mogeneous system (15) by setting the 6; equal to zero is called the reduced system
associated with (15)
A basic property of a linear system of DEs (16) is that the difference x — y
of any two solutions of (16) is a solution of the reduced system. It can be imme-
diately verified that any linear combination ax() + by(é) of solutions x(#) and
y(t) of a homogeneous linear system is again a solution
8 Linear Systems 189
LEMMA. Any linear system (15) with continuous coefficient functions on a closed
interval I satisfies a Lipschitz condition (4) with
Proof. Since X(x, #) ~ X(y, 4) is the vector sum of n* vectors Z,;, with ith
component a,(x;; — y,) and other components zero, repeated use of the triangle
inequality gives
sS a sup la;(O|- Ix — yl
THEOREM 7. The initial value problem defined by a linear system (15), with the
a, ;(t) and b,(t) defined and continuous for |t — a | = T, and the initial condition x(a)
= ¢, has a unique solution on |t — a | ST.
Proof. The preceding lemma shows that such a system satisfies the hypothesis
of Theorem 6. This gives the existence of the solution. The uniqueness follows
from Theorem 1, again by the preceding lemma.
For homogeneous systems, we can construct a basis of solutions, as follows.
-1 -2
du u
dt
= p(t)
di”) + pl) Face + + pri) = + pala
transforms the DE into a homogeneous linear system dx/dt = A(é)x, where the
matrix ||@,,(é)|| = A(@ is defined as follows: a,() = 0 if 1 = i= n— 1 and
J#it la .@ = liflsisn— 14,0 = p10.
We therefore obtain the following result.
In Theorem 6, it was assumed that X(x, 4) was defined for all x and satisfied
a Lipschitz condition (4) for all x. But often this is not the case. For instance,
this assumption does not hold for the DE dx/di = e* of Example 4. The ratio
THEOREM 8. Suppose that the function X(x, t) in (2) is defined and continuous
in the closed domain |x — c| = K, |t — a| = Tand satisfies a Lipschitz condition
(4) there. Let M = sup |X(x, 1)| in this domain. Then the DE (2) has a unique solution
satisfying x(a) c and defined on the interval |t — a| = min (T, K/M).
=
=
Proof. All steps in the proof of Theorem6 can be carried out, provided we
know that the functions x"(f) referred to there take their values within the
domain D,: |x — c| = K, |é — a| S min (7, K/M), in which x(é) is surely
defined. In particular, note that since D, C D, the bound M and Lipschitz con-
stant L of Theorem 6 can be used in D,. Therefore, the proof is a corollary of
the following lemma.
LEMMA. Under the hypotheses of Theorem 8, the operator U defined by (12) carries
functions x() satisfying the conditions: (i) x(0) is defined and continuous on |t — a|
<= min (T, K/M); (ii) x(@) = ¢; iii) |x@ — c| S Kon the interval |t — a| S min
(T, K/M), into functions satisfying the same conditions.
Proof. In (12), suppose that x(s) satisfies conditions (i), (ii), (iii). We must
show that y(#) satisfies the same conditions. Clearly (i) and (ii) are satisfied by y(é).
By the inequality (3) we have (taking again ¢ = a for simplicity)
ly¥® — c| =—
=
=
Therefore, (iii) is satisfied and y(¢) is defined for |¢ — a| < min (T, K/M), com-
pleting the proof.
The existence theorems for normal systems (1) proved so far have assumed
that the functions X, satisfy Lipschitz conditions. We shall now derive an exis-
192 CHAPTER 6 Existence and Uniqueness Theorems
Proof. Using an elegant method due to Tonelli, we shall consider the equiv-
alent integral equation (11) of Theorem 5,
and prove that this has a solution. Let T; = min (T, K/M). We may assume that
0 and that the interval is 0 = ¢ = Tj. In this interval we construct a
=
a =
sequence of functions x"(t) as follows. For 0 < t = T,/n, set x"() = cc. For
T\/n <t S T, define x"(é) by the formula
This formula defines the value of x"(¢) in terms of the previous values of x"(s)
for0 =sSt—T,/n.
It follows, as in the lemma of §9, that the functions x"(¢) are defined for 0 =<
t <= T;. Also, we have
Ti
Applying this result to the sequence x"(/), we see that it must contain a uni-
formly convergent subsequence x™(f), converging to a continuous function x*(é)
as n,
> ©.
It is now easy to verify that this limit function x™(¢) satisfies the integral equa-
tion (19). Indeed, (20) can be written in the form
&
Fm | = Spe “™M “= 0 21
n;
Therefore, taking limits on both sides of (21) as n; —> 00, we find that x® satisfies
the integral equation (19), q.e.d.
We shall now consider the vector DE (2) under the assumption that X (x, 4)
is an analytic function of all variables x, , x,, t. The essential principle to be
established is that all solutions of analytic DEs ave analytic functions.*
The result is true whether the variables are real or complex; we shall first
consider the complex case. To emphasize that we are dealing with complex vari-
+ Rudin, p. 164 ff. The proof given there is for real-valued functions, but the method applies to
vector-valued functions.
* This section requires a knowledge of elementary complex function theory such as is found in the
books by Hille (Vol. 1) and Ahlfors.
194 CHAPTER 6 Existence and Uniqueness Theorems
J
We assume that the Z,(z,, » Zy #) are analytic functions of the variables z,,
Zo ., 2, and ¢ in the closed cylindrical domain C: |t — a| = T, |z—c| =K,
with maximum M there. By the lemma of §2, this implies that a Lipschitz con-
dition holds in C, for some constant L.
Vector notation can be adapted to complex vectors with the following
changes. The length (or norm) of a vector z (z1, 29» , Z,) with complex
=
=
components z, is defined as
is defined as
Note that z w = (w z)*: the dot product operation is not commutative for
complex vectors. (The set of complex n-vectors with the above inner product is
called a unitary space.)
Now let y be any path in the complex ¢-plane, defined parametrically by the
equation ¢ = t(¢) = r(o) + is(o), where 7, s € @' andais a real parameter. On
the path y, (22) is equivalent to the system of real DEs
vided that this path stays within the domain C where the function Z is defined.t
By Morera’s theorem, the function w is therefore also analytic, in the sense that
each component w,(z) is. Thus, the operator W transforms analytic functions
into analytic functions. Moreover, the lemma of §9 still holds, because the inte-
grals in (23) can be taken along straight line segments in the complex {-plane.
This gives the following lemma.
LEMMA. For |t — a|< min (T, K/M), the operator Wdefined by (23) takes anal-
ytic complex-valued vectorfunctions z(t) with |z(t)— ¢|<S K into analytic vectorfunc-
tions w(t) with |w)— c|= K.
THEOREM 10. In Theorem 8, replace the real variables t, x,, X, with complex
variables t, z;, Z,. Under the same hypotheses, if the Z,(z, t) are complex analytic func-
tions, the vector DE (22) has a unique complex analytic solution 2(i) for given initial
conditions.
From this result and the uniqueness theorem, again for real DEs, we obtain
the following corollary.
+ This is true because the disk |¢ — a| << T where Zis defined is simply connected.
*Ahifors, p. 173. The result contrasts sharply with the case of functions of a real variable. By the
Weierstrass approximation theorem, every continuous function on a real interval a < x < bisa
uniform limit of polynomial (hence analytic) functions.
tt An alternative proof can be based directly on (22). If the z,() satisfy (22), they are continuously
differentiable. Hence, they are analytic (Ahlfors, pp. 24, 105; Hille, Vol. 1, pp. 72, 88).
196 CHAPTER 6 Existence and Uniqueness Theorems
Real Analytic DEs. A real function X(x, é) of real variables x,, , x, and
tis said to be analytic at (c, a) when it can be expanded into a power series with
real coefficients in the variables (x, — c,) and (¢ — a), convergent in the cylinder
|x — c| <4, |f — a] <, for sufficiently small positive y and ¢. When X(x, #)
is analytic, its power series is convergent also in the complex cylinder |z — c| <
n, |t — a| < € (¢ complex), and defines a complex-valued analytic function there.
Now, let a normal system of real DEs (1) be given, the X,(x, t) being analytic.
From Theorem 10, it follows that the resulting complex DE dx/dt = X(x, #) has
a unique complex analytic solution for given real initial values. On the other
hand, it also has a unique (local) real solution by Theorems 1 and 8. Hence the
two solutions must coincide, proving Corollary 2.
EXERCISES E
1. (a) Obtain an equivalent first-order system for d?x/dé? = t?x. Find the nth term of
the Picard sequence of iterates for the initial values x(0) = 1, x’(0) = 0.
(b) Prove that this initial-value problem has one and only one solution on (—,
oO),
(a) Obtain an equivalent first-order system for the DE d®x/dt” = x* + 2, and find
the Lipschitz constant for the resulting system in the domain |t} =< A, |x| = B,
Ix’| <C.
(b) State and prove a local existence theorem for solutions of this DE, for the initial
conditions x(0) = 6, x’(0) = ¢. Estimate the largest 7, U such that a solution is
defined on —U <1 T.
Show that, if F(y) is continuous for [y| =< K, and | F(y)| < M, every solution of y’ =
F(y) can be uniformly approximated arbitrarily closely for [x] <= K/M by a solution
of a DE y’ = P(y), where P is a polynomial.
Compute the nth Picard approximation to the solution of the complex system dw/dt
=
=
iz, dz/dt = w, which satisfies the initial conditions w(0) = 1, z(0) = i.
In the complex ¢-plane, determine a domain in which the system dw/dt = iz’, dz/at
= tw* has an analytic solution satisfying given initial conditions w(0) = wo, 2(0) = 2.
Show that the solution of the complex analytic DE
w'(z) = M | 21
—_
(1+2 yy
K
(lz1 < *)
which satisfies the initial condition w(0) = 0, is the function
i] =
w(z)= K (: + a IK
(n — 1)M
*7 Using the result of Ex. 6, show that the bound given by Theorem 8 for the domain
of existence of a solution is “best possible” for analytic functions of a complex
variable.
12 Continuation of Solutions 197
Even when the function} X(x, 2) is of class @! and is defined for all x and #,
Theorem8 establishes the existence of solutions only in the neighborhood of a
given initial value. In other words, it establishes only the local existence of solu-
tions. We shall now study how such local solutions can be joined together to give
a global solution defined up to the boundary of the domain of definition of the
function X.
THEOREM 11, Let X(x, é) be defined and of class @' in an open region # of (x,
t) -space. For any point (c, a) in the region A, the DE (2) has a unique solution x(t)
satisfying the initial condition x(a) = ¢ and defined for an intervala St<b(bxs
00) such that, if b < 00, either x(t) approaches the boundary of the region, or x(t) is
unbounded as t > b.
Proof. Consider the set S of all local solutions of the system (2) that satisfy
the given initial condition x(a) = c. These are defined on intervals of varying
lengths of the form [a, T). Given two solutions x and y in this set, defined on
intervals J and I’ respectively, the function z, defined to be equal to x or to y
wherever either is defined, and hence also where both are defined, is also a solu-
tion defined on their union J U I’.
We now construct a single solution x, called the maximal solution, defined on
the union of all the intervals in which some local solution is defined, by letting
x(é) be equal to the value of any of the solutions of S defined at the point ¢. This
maximal solution x(é) is a well-defined function of class @', by the Uniqueness
Theorem. Furthermore, the interval of definition of this solution is the union
of all the intervals of definition and, therefore, is itself an interval of the form
ast<b.
Consider the limiting behavior of x(#), as ¢ tf b. By the Bolzano—Weierstrass
Theorem,‘ any infinite bounded set of points (x(¢,), ¢,) in xf-space must contain
a limit point. Hence either 6 = +00, or lim,, |x()| = +, or at least one
finite point (d, 6) is approached by at least one sequence of points [x(,), ¢,] on
the above solution curve. In the first case, ¢ is unbounded. In the second case,
x(é) is unbounded and the maximal solution may be said to “recede to infinity.”
It remains to consider the third case. A typical example is provided by choos-
ing the region & as the left half-plane t < 0 and x’(t) = ¢”? cos ¢~', with general
solution x C—sint?.
=
=
We shall now prove that, in the third case above, every limit point (d, 5) on
t = b of the maximal solution curve must lie on the boundary of #. Indeed,
suppose that it is in the interior; there would then exist a closed neighborhood
+ In this section we consider only real vectors and functions. The results can, however, be extended
to complex-valued and analytic functions, by methods similar to those used in § 1. The continuation
so defined is then the analytic continuation in the sense of complex function theory, by Theorem 9
(cf. Ch. 9, § 1).
+ Cf. Courant, Vol. 2, pp. 95 ff., where the Bolzano—Weierstrass Theorem is proved in R".
198 CHAPTER 6 Existence and Uniqueness Theorems
A solution with a finite escape time is one for which |x(é)| becomes
unbounded or reaches the boundary of # as t — b < 00, On the other hand, a
solution with an infinite escape time is one that remains within the domain of
definition of X for all tf > a. For example, every solution of the DE dx/dt = x
2
has infinite escape time, whereas every nonzero solution of the DE dx/dt = x,
namely every function x 1/(c — 2), has finite escape time.
=
=
It is easy to derive a formula for the dependence onc of the solution x = f(¢,
c) of the initial-value problem defined by the system x’() = X(x, é) and the initial
condition x(a) = c. For simplicity, consider first the case n = 1 of a single first-
order DE. Assuming that /(, c) is analytic, that is, that fhas a convergent Taylor
series expansion, we have
a off\=3c
9g of;_ 9
(25)
ot (
Oc (at Oc
[X(t ©), o)]
= |Xoe.0.0] -| of(t, ¢) |
“4
iC
obtained by linear perturbation are accurate near stable exact solutions of initial
value problems, but they can be misleading near unstable solutions.
We now drop the assumption that X is a one-dimensional vector, as well as
the assumption that the solution has a convergent Taylor expansion, and we
derive analogous results. We show that the solutions of a normal first-order sys-
tem (1) depend differentiably on the initial values, thus proving (at long last!)
that the solution curves of any normal first-order DE or system form a normal
curve family.
THEOREM 12. Let the vector function X be of class C', and let x(t, c) be the
solution of the normal system (2), taking the initial value c att = a. Then x(i, c) is a
continuously differentiable function of each of the components c, of c.
where x(t, c) is the solution of the normal system (2) for which x(é, a) = c. We
assume that the functions H; are bounded for |t — a] <= Tand |h — #| =
K, where 6 is the vector whose components are the Kronecker deltas 6]. We also
assume that H, tends to zero as y, > 0, uniformly for |f — a] <= T and
|h — 6’| < K. We define hi = hii, c, n,) as the solution of (28) that satisfied the
initial condition h/(a) = 4/. Applying the Corollary of Theorem 3, with « = 9,
we find that h/ tends, as y, > 0, to the solution f? of the linear system
th _ > OX,(x(¢,c), #)
(29) f
dt k=1
Ox,
dgi(t, c, ny) -
dt
ny[Xx ©) + ng’ ©), t) — XGx(, ©))]
200 CHAPTER 6 Existence and Uniqueness Theorems
We now use the assumption that X, is @’. By Taylor’s theorem for functions of
several variables, we infer that the right side equals
AX(x(t, ©), t)
A(t, c) + | g,(t, ¢)|
k=l Ox;
We have, therefore, shown that the derivative 0x/dc, exists and is indeed a solu-
tion of (29), q.e.d
The linear DE (27) is called the perturbation equation or variational equation of
the normal system (2), because it describes approximately the perturbation of
the solution caused by a small perturbation of the initial conditions
In the course of the preceding argument we have also proved the following
result
COROLLARY. /f x(t, c) is a solution of the normal system (2) satisfying the initial
condition x(a) c for each c, and if each component of the function X is of class @
then for each j the partial derivative Ox(t, c)/dc, is a solution of the perturbation equa
tion (28) of the system
(30)
& = X(xa) iy
>-
=
=
Y(x,9)
di dt
THEOREM 13. Any plane autonomous system where X andYare of class @' is
equivalent under a diffeomorphism, in some neighborhood of any point that is not a
critical point, to the system du/dt = 1, do/dt = 0.
Proof. Let the point be (a, 6); without loss of generality, we may assume that
X(a, b) # 0. Let the solution of the system for the initial values x(0) = a, (0) =
cbhex = &t, c), y = nt, c), so that 0&/dt = X, On/dt = Y. Then by Theorem 12,
the transformation (t, c) F (E(t, ¢), n(t, c)) is of class @'. Moreover, since x(0) does
not vary with ¢, the Jacobian
ag, 7) 0& On of on
-_ SS xa, )- 1—0- Ya, b) = X(a,b)
=
—_—
At, c) at Oc Oc ot
is nonvanishing at (a, b). Hence, by the Implicit Function Theorem, the inverse
transformation u
—
=
t(x, y),
=_
=
c(x, y) is of class @'. In the (u, v)-coordinates,
the solutions reduce to u = 4, v = ¢ constant; hence, the DE assumes the
=
=
COROLLARY 1. Any two plane autonomous systems are locally equivalent under
a diffeomorphism, except near critical points.
The system u 1,ov 0 is, therefore, locally a canonical form for plane
= =
= =
COROLLARY 2. [If the functions X andY of the plane autonomous system (30)
satisfy local Lipschitz conditions, then tts integral curves form a regular curve family
im any domain that contains no critical points.
EXERCISES F
1. Let F(x, y) be continuous for |x — a] =< T, [y — c] < K. Show that the set of all
solutions of y’ =
=
F(x,y), satisfying the same initial condition f(a) oo
=
c, is
equicontinuous.
202 CHAPTER 6 Existence and Uniqueness Theorems
2. Show that, if X(x, #) is continuous and satisfies a Lipschitz condition for a = t <= 3b,
every solution of the DE (2) satisfying x(a) = ¢ is bounded for a = ¢ = db. Show that
the corresponding result is not true for open intervals a < t < b.
Let «X(x,y) + °¥(x,y) = 0, where X andYare of class @!. Show that the system x’
= X(x,y), y’ = Y(x,y) has infinite escape time. [Hint: Show that 2x” + y* is an integral
of the system.]
Let the function X(x, /) be defined for 0 < ¢ < © and for all x, and let
Let X(x, #, s) be of class @! for [x — ec] <= K, |t — af ST, [s — 5| SS. Let x(t, s)
be the solution of x’ = X(x, é, s) satisfying x(a) = c. Show that x is a differentiable
function of s.
*6 Under the assumptions of Ex. 5, suppose that X(x, ¢, s) is of class @*. Show that
x(t, s) has n continuous partial derivatives relative to s.
*7 Show that if there are two distinct solutions fand g of y’ = F(x, y) satisfying the same
initial condition ¢ = f(a) = g(a) (F continuous in [x — a] = T, |y ~— c| = 4A), there
are infinitely many of them.
*8 Show that there is a maximal and a minimal solution fy(x) and f,(x) of the DE in Ex.
7, such that f,,(x) <= f(x) < f(x) for any other solution f such that f(a) = f(a) =
Fal). UHint: See Ch. 1, Ex. F4.]
ADDITIONAL EXERCISES
*1. Let dx/dt = X(x, y, t) and dy/dt = Y(x, y, t), where
(a) Prove that, if there is a normal Ath-order ordinary DE satisfied by two functions
uand v and if n > &, there is a normal nth-order DE satisfied by both functions.
State your differentiability assumptions.
(b) Prove that, if the given kth-order DE is linear, then the nth-order DE can also
be chosen to be linear.
13 The Perturbation Equation 203
(c) Prove that there is no fourth-order normal DE u” = Flu, u’, u”, u’”, t) satisfied
by both u =
=
t* and v =
=
i for all real ¢.
8
(d) Prove that «
=
=
satisfies no normal linear homogeneous DE of order six or
less with continuous coefficients.
Show that, if X(@) = ||x,(¢)]| is a matrix whose columns are solutions of the homo-
geneous linear system X’ = A(#)X, then det X(f) = [det X(a)] exp ff E ay,(s) ds.
A matrix X(é) is a fundamental matrix for a <= t = a + T of a homogeneous linear
system X’
=
=
A(X if its columns are solutions of the system and det (X(t) # 0. Show
that, if the columns of X are solutions of the system and if det X(a) # 0, then X is
a fundamental matrix.
*10. Show that, if X(¢) is a fundamental matrix of the reduced linear system, the function
x(t) = X(t) JLX7'(s)b(9) ds is the solution of the inhomogeneous system such that x(a)
= 0 (X7! is the matrix inverse of X).
CHAPTER 7
APPROXIMATE
SOLUTIONS
1 INTRODUCTION
During the past 40 years, the accurate numerical solution of initial value
problems for ordinary DEs has become routine, because of the availability of
high-speed programmable computers. Even fairly large systems of DFs can be
treated similarly in many cases, although “stiff? systems involving time scales of
different orders of magnitude can be troublesome.
This development has not only made the study of classical numerical methods
(e.g., Runge-Kutta methods) more important, as practical substitutes for
involved analytical considerations, it has also increased interest in numerical
mathematics from a theoretical standpoint. In particular, the power series methods
explained in Chapter 4, together with techniques of numerical linear algebra,
have provided the basis for a new field of research.
Because of this changed emphasis, a few simple numerical methods for solv-
ing DEs were already described in Chapter 1, §8. In this chapter and the next,
we will treat the numerical solution of ordinary DEs and systems of DEs more
carefully. This chapter will concentrate on the underlying ideas, while the effec-
tive technical implementation of these ideas will be the subject of Chapter 8.
Since these ideas are applicable to systems of first-order DEs, we will adopt
throughout Chapters 7 and 8 the vector notation introduced in Chapter 5.
Thus, we will consider vector DEs of the form
at+T
=
(2) Tv €=ty<t<tp<t,<---
<4, =
2 ERROR BOUNDS
This formula is recursive; each value x, can be computed knowing x,-_, alone.
From the approximate function table just defined, one can also construct an
approximate solution by linear interpolation. This approximate solution is defined
by the formula
Example 1, When the DE (1) is of the special form x’ = f(t), the preceding
method reduces to the Riemann sum formula of Ch. 1, (5’):
+ It was Cauchy who first proved their convergence to exact solutions, though Euler had used “Cau-
chy polygons” a century earlier.
206 CHAPTER 7 Approximate Solutions
15 q T
1.25 -
h=0.2
1L0r
y mer
h=mAar=|xij=}
75
! l f J
0.55
0.2 0.4 0.6 0.8 1.0
t
Figure 7.1 Cauchy polygons for dx/dt = x, x(0) = 1/2.
5 ||
—
=
max(At,, ., At,) = max lee — te)
k=1,...,
THEOREM 1. Let X€ @' satisfy |X| S M, |OX/dt|S C, and (5) in the cylinder
atSa+t/T, |x — cl S MT. Then the Cauchy polygon approximation xz(t)
differs from the true solution x(t) by at most
3
(n — 1)h? n
(*) = (lth =1l+nh+n
2 + h 3!
+
This section will be devoted to proving Theorem 1, that Euler’s method has
O(h) accuracy. The proof will be based on a new concept: the deviation of a func-
tion from a DE. This concept is of theoretical interest in its own right.
for all except a finite number of points ¢ of the interval [a, a + T}.
THEOREM 2. Let X ¢ @’ satisfy |X| <= M, |9X/dt|<S C, and (5) in the cylinder
D: |x — c| SK,aStSat T. Then any Cauchy polygon in D is an approximate
solution of x’(t) = X(x,t) with deviation at most (C + LM)|x|.
In proving this theorem, we will use the fact that any Cauchy polygon approx-
imation is continuous in [a, a + T], and is differentiable at all points not mesh
points. At these, it still has a left and a right derivative.
THEOREM 3. Let x(t) be an exact solution and y(t) an approximate solution with
deviation ¢, of the DE x’ (t) = X(x,t). Let X satisfy the Lipschitz condition (5). Then,
fort 2 a, we have
Now, set ¢ = v’; the foregoing gives v’ < Lv + ¢ (for ¢ > 0). Applying Theorem
7 of Ch. 1, §12, we get the desired inequality (9), much as in proving the lemma
of Ch. 6, §4.
A slight variant of the analysis leading to Theorem 3 yields a closely related
bound to the cumulative error of the Cauchy polygon approximation, as follows.
Define the directional derivative 0X/0£ of the vector function X(x,?) in the
direction
& = (&,€,, » §,) in t,x-space, for any vector & of unit length, as the
sum £ 0X/dt + Lin, & IX/dx,. It follows as in the proof of the lemma of Ch.
6, §2, that
term of the inequality (9) vanishes if we let y(é) be the Cauchy polygon approx-
imation for the initial value x(a) in Theorem 3, and so (9) simplifies to
N
=
=
E + | [expLT— 1]
we obtain the following simple corollary of (6).
COROLLARY. Under the hypotheses of Theorems 1 and 2, let the interval [a, a
+ T) be divided into n equal parts of length h = T/n. Then the error of the Cauchy
polygon approximation is bounded by Nh, where N is a constant independent of h.
EXERCISES A
1. (a) What is the deviation of the approximate solution x = #°/2 — t*/24 of the initial
value problem defined by dx/dt = sin t, x(0) = 0 on the interval 0 <= ¢ =< 1?
(b) Compare the difference 1 — (cos 1) — 34 with the bound given by formula (*),
for the deviation computed in (a).
(c) For the initial value x(0) = 1, bound the difference between the solutions of
dx/dt = sin t and dx/dt = t — (/6).
210 CHAPTER 7 Approximate Solutions
In Exs. 2-5, for the initial value problem specified: (a) use the Cauchy polygon method
to compute an approximate function table for ¢, = 0.1, 0.2, ..., 1.0; (b) find the devia-
tion of the approximate solution obtained from this table by linear interpolation; (c) find
the exact solution, (d) find the error.
= x,
2.x (0) = 1 3. *¥ = 1 — 2x, x(0) = 0
6 (a) Find the deviation of the approximate solution y = 107!° of the DE x’() =
5x,
(b) What is the exact solution of this DE on [0, 0%) for the initial value (0) = 107!°
= (0)?
(c) Prove in detail] the uniqueness of this solution.
On the interval [0, 1], for any « > 0, construct an approximate solution with devia-
tion € to a suitable first-order DE, for which the exact solution with the same initial
value is unbounded.
*9 (a) Sharpen (9) and (*) in the stable case [X(x, t) — Y(x, )] - [x — y] = 0. Compare
with the limiting case L = 0 of these formulas.
(b) When X(x, #) satisfies the one-sided Lipschitz condition [X(x, 1) — X(y, )] -
[x — 9] = L |x — y|?, how can (9) and (*) be sharpened? [Hrnr. See Ex. 8.]
In practical computation, one can often reduce the error by a large factor by
accepting as a working hypothesis, the theoretical result that, in a wide variety of
situations, the truncation errort under repeated mesh-halvings is of the form
As has just been emphasized, the order of accuracy v 1 for the Euler—Cauchy
=
=
polygon method. For the modified and improved Euler’s methods to be dis-
cussed later in this chapter, »y = 2. For other methods to be discussed in Chapter
= 4.
8,»
If one knows p a priori, as one does for the methods of Euler just mentioned,
one can determine the unknown constant C in (10) with fair accuracy, by com-
paring the computed value Y) for a given partition x9, with the corresponding
value Y, for the partition 7, obtained from it by mesh-halving.
This is because formula (10) implies that
Y=¥Y,-
(Yn — Yon)”
(14)
Yan —- 2Yon + Y;,
L 1 1
A= 2 4 8 16
A= $ i & 1
Y, 2.640625 2.694856 2.711841 2.716593
Error .077657 023426 006441 .001689
(4¥,
— 3¥a4)/3 2.6875 2.712933 2.717503 2.718177
Error 030782 .005349 .000799 -000105
Roundoff Errors. The preceding discussion has set no limits to the fineness
of the mesh used in solving DEs numerically, and it has been tacitly assumed
that all arithmetric operations and function evaluations are exact. Actually,
however, the floatin -point arithmetic on many computers has an accuracy
of only around 10
-
5. On such computers, the dominant source of error
when h = 1/1024 (say) may well be due to so-called ‘roundoff errors” in
floating-point arithmetic. This is especially likely if values of the x, that are
not exact “binary decimals” are used—e.g. if h = .001 is used instead of
h = 1/1024.
Roundoff errors will be discussed again in Chapter 8, §6.
5 MIDPOINT QUADRATURE
_ G1
+ x)
(15) SF(x) dx = M,[F] = > F(m,) Ax,,
1=1
t
2
a =
computed; it takes its name from the fact that m, is the midpoint of the ith inter-
val of subdivision. We now derive an error bound for the midpoint quadrature
formula (15).
5 Midpoint Quadrature 213
| fF #9 ax - ~ F(m,) Ax, =
JF” | max|r} 2(b _a)
(16) —
t=]
24
PF"(m; + 7)
F(m, + }) — Fm) — tF’(m,) =
2
3 Xe
Fmin z < Ft, Ax?
—_ ; F(x) dx — F(m,) Ax, < =
ae
24 i 24
Summing over i and noting that 0 < Ax? < |x|, we get (16).
Theorem 4 shows that the midpoint quadrature formula (15) has order of
accuracy O(h*), one order higher than the Cauchy polygon method.
5
Fmt" FOr!
Fm, +0 =>-
r=0 (7!) 720
214 CHAPTER 7 Approximate Solutions
where £ is some number between m, and m, + t. On each ith interval (x,_,, x,),
the final term (“remainder”) is bounded in magnitude by Mk°/720, where M =
max | F(£)|, the maximum being taken on the entire interval a < & = 5. Inte-
grating over —k =i Sk, we get
x1 Be Fe (m,)
WF?(m)
F(x) dx = 2kF(m,) + +
+ O(k*) Ax,
x:-1 3 60
where the factor O(&°) is bounded in magnitude by Mk°/720. When we sum over
z, there results the estimate
- 120
> F°(m,) Ax, + O(F°)
emt
An application of (18) to the function F”(x) € @* gives similarly (one term being
dropped because of the loss in differentiability),
Substituting from (18’) and (18”) back into (18), and combining terms, we get
7k
x F(m,)Ax, = f F(x)dx — . f F(x) dx + 360 J"F"(x)dx + O(k°) ——ee
21
IF) — P@)
(19) f " F(x) dx = 3 F(m) Ax, +
2=]
24
_ TF") — F"@)]
+ O(n)
5760
6 Trapezoidal Quadrature 215
EXERCISES B
In each of Exs. 1-4, a numerical quadrature formula is specified for approximately eval-
uating f”, f(x) dx. In each case: (a) compute the truncation error for f(x) = x", n = 0,
1, 2, 3, ..., and (b) find the order of accuracy of the formula, using Taylor’s formula
with remainder, assuming f(x) to be analytic.
(a) Show that all odd-ordered derivatives F°**(0) of F(x) = 1/(1 + x?) vanish when
x = 0.
(b) Show that F(1) = —}and F’’(1) = 3
(c) Knowing that 7/4 = arctan 1 = f,3 dx/(1 + x*), derive the formula
he
+— -—
7h*
a/4 = M,[F] + O(n)
48 7680
6 TRAPEZOIDAL QUADRATURE
(20)
J” Rex) dx = T,[F] = 3 [F(x,-1) + Flx)] Ax,/2.
i=]
216 CHAPTER 7 Approximate Solutions
We will now use the concept of the Green’s function for a two-endpoint prob-
lem, as defined in Ch. 2, §11, to obtain an exact expression for the error in
trapezoidal quadrature over a single interval. Consider the linear function
defined by linear interpolation between the values F(a) and F(6), and let R(x) =
F(x) — L(x). Then R” (x) = F’(x), and R(a) = R(b) = 0.
Now consider a single interval of length h, = x, — x,-; = 2k, and translate
coordinates so that (x,_,, x,) becomes the interval (—k, k). As in Ch. 2, §11, we
have
x =&,
(23) G(x,2) = |(fx/k + & — x — k)/2,
(fx/k — E& + x — h)/2, =x.
he
Substituting for R(x) the integral expression displayed above and interchanging
the order of integration in the resulting double integral, we get
Since h,? < ||*h,, summation over i now gives our final result.
R2 4
The right side of (25) can be evaluated by repeated use of the midpoint quad-
rature formula error estimate (17). The conclusion is the truncated Euler—
Maclaurin formula.
THEOREM 8. For F € @®, let all intervals of subdivision have the same length
Ax; = 2k = h. Then
b 2 4
J a
F(x) dx — & [FO — F(@) + =
360
[F”(b) — F”(a)] + O(K*)
218 CHAPTER7 Approximate Solutions
while the third term gives (k*/24)[F” (b) — F”(a)] + O(k°). Adding these three
contributions together, simplifying, and writing k = h/2, we get (26).
Simpson’s Rule. Comparing the error estimates (17) and (26) for midpoint
and trapezoidal quadrature, we are led to an error estimate for Simpson’s rule.
For a given partition 7, this is defined as
Forming the linear combination indicated for subdivision into double steps of
constant length 2k = h, we obtain
EXERCISES C
ne-[ 1, 1 1 1
20
—
=640 >
k=1
10 +k 400 |
In Exs. 2—5, use (26) with A = 0.2 to evaluate the following numbers approximately:
4. fi VT + x dx 5. Si sin (x?) dx
In Exs. 6-9, use Simpson’s rule (28) with double step 2k = h = 0.2 to evaluate approx-
imately the numbers defined in Exs. 2—5, respectively.
10. For a subdivision into 2n intervals of length kh = (6 — a)/2n, Simpson's approxima-
tion to f? f(x) dx is 5%, (4/3)[fx2,-1) + 4/flre,-1) + f(xs,)]. Show that the truncation
error is (#°/90) E%, f"(x2,-1) + OCA).
11 Show that f*, F(x) dx =
=
*7 TRAPEZOIDAL INTEGRATION
The rest of this chapter will be devoted to the theoretical analysis of three
classical methods for integrating first-order ordinary DEs (and systems of DEs).
Like (uncorrected) midpoint and trapezoidal quadrature, these methods have
7 Trapezoidal Integration 219
only O(h) accuracy. Since Runge-Kutta and other algorithms having at least
O(h*) accuracy are readily available and easy to use, readers who are primarily
interested in applications of numerical methods may wish to proceed directly to
Chapter 8, which will take up such more efficient methods.
For various reasons, the errors arising from the use of these more efficient
methods cannot in practice be predicted purely theoretically. Therefore, the dis-
cussion to follow will have no parallel in Chapter 8.
Our theoretical analysis will first take up trapezoidal integration. For any sys-
tem dx/dt = X(x, i) of first-order DEs, this is defined implicitlyt by the recursion
formula (difference equation)
where 6, = At,/2, A, = A(t), and b, = b(i). Hence, for At, small enough, the
system (30’) can be solved for y,, given y,-,, by Gaussian elimination.
(31) + 2x 1
=
=
taking the initial value x(0) = 0. By the formula of Ch. 1, §6, the solution is
the function x
—
=
e~® fi, e* ds. Looking up values of the definite integral in a
table,tt we get the first row of entries in the following display.
+ For large At,, Eq. (80) may have more than one solution. But usually, y, can be computed by
iterating (30) two or three times.
[This does not mean that reduction to a first-order system is recommended in numerical
integration.
+{ E. Jahnke and F. Emde, Tables of Functions, Dover, 1943, p. 32; W. L. Miller and A. R. Gordon,
J. Phys. Chem. 35 (1931), p. 2878.
220 CHAPTER 7 Approximate Solutions
The second row of entries gives values y, of the approximate function table
constructed using the trapezoidal integration formula (30), with x(0) 0 and
constant mesh length h 0.1. With this value of 4, the formula in Example 4
reduces to
1 + (10
— t-1)y-1
10 + &
The entries tabulated in the second row were calculated from the preceding
formula, rounding off all numbers to six decimal digits and then rounding off
the final values to four decimals. The last row of the table gives values z, com-
puted by the “improved Euler method” of §8 below.
Above, we used a table of [6 e* ds; we next ask: How should one construct
this table? For large i, numerical quadrature formulas tend to be inefficient,
because the integrand e” varies so rapidly (by more than 10% between 5 and
5.01, for example). For this reason, rather than computing x(f) as we did, it is
more efficient to solve the DE (31), for the initial value x(0) = 0,by an accurate
numerical method (see Ch. 8), and then to compute {5 eds = é x(t) as a prod-
uct, than to compute x() as we did.
rt 6
ot [ea f (ree |a
Tr
+5t+qt-
2!
tt 1 (Qk)lt 2k-1
Moreover, by truncating the series at the kth term, an asymptotic error estimate
can be obtained
Knowing (or having guessed) the form of the asymptotic series x(t) ~ Lj-o
a [ee we also can derive from (31) purely formally, by the method of unde-
termined coefficients, that a) = 1/2 and a4, = (2k + 1)a,/2, from which
(asymptotically) we obtain
3 15
x~o~—+ 4° set abe
_——
as t—>
©
2t 81 16
7 Trapezoidal Integration 221
Though the series is (ultimately) divergent, the first few terms give an extremely
accurate approximation to x(¢) for ¢ 2 25, which is much more accurate than
could be obtained by general numerical methods
Qe = e-vT
9) = Ye + MAT + s Tet by
2 At
interpolates not only to y,_, and y,, but also to y,_, and yj at ¢,_; and &, respec-
tively.t This gives us a piecewise quadratic approximate solution of class @', which
satisfies. the given DE exactly at all points ¢,
We next bound the deviation (§1) | y’(é) — X(, y()| of this approximate solu-
tion. Since the deviation is zero at t,_; and ¢,, while y’(¢) is linear in the interval
[t,-1, t], it follows (by Theorem 1 of Ch. 8) that the deviation there is at most
t(h —1)|X} maxy Where h = t, — 1, and |X| ax signifies the maximum absolute
value of the second time derivative of X(f). On the other hand, we have
ax
_ 9X OX OX OX AX Ox
(32) x~
de
ye
€t
+ 2x
OxOt
+
at Ox
ax ot (Ox
where, in any subinterval [#,-), ¢,] of length h or less |e(7)| S or(h — 1)|X} mays
tT = ¢— t,-,;. A more careful repetition of the proof of Theorem 3 shows that
since f§ t(h — #) dt 7/6 and |x(a) — y(a)| = 0, the cumulative (truncation)
error must satisfy
h?
(33) 7,
Ix) — 91 SEF |Xlmaater® “9 1}
t It is actually the “cubic” Hermite interpolant to the y, and y) (to be discussed in Ch. 8), but this
happens to be quadratic in the present case
222 CHAPTER 7 Approximate Solutions
EXERCISES D
In Exs. 1-4, compute the trapezoidal approximations to the solutions of the initial value
problems specified, over the range 0 = ¢ = 1 withh = 0.1.
1 x
=
=
—Ix, x(0) = 1 2.2= 14+ x7, x(0) = 0
3 x =
h
—
2
|X .50 + 70.10%0.59] +00, 90= 90)
In Exs. 6-9, compute the truncation error for trapezoidal integration over one interval
of length h, in terms of the Taylor series expansions of the functions involved:
6. % = plt)x, p analytic 7.
% =
=
1+ x?
xX= y,
8. 9.
x
=
y=
0 =
ys jaxty
10. For the DE y” = F(x, y), what is the order of accuracy of the formula
if all Ax, = A?
r+]
(33) yz = yr-1 + EX(ya-1, te-1) + Xi, t)] At,/2
+ The vector equation (30) is, of course, equivalent to a system of n simultaneous equations in the
components, For linear systems, these equations can be solved by Gauss elimination, instead of
iteration.
8 The Improved Euler Method 223
Instead of solving the implicit equation (30) precisely by using many itera-
tions, one usually gets greater accuracy for the same amount of work by using
a finer mesh and stopping after one or two iterations. If one stops after a single
interation, one has the improved Euler method, which is adequate for many non-
linear engineering problems requiring moderate accuracy (say, two significant
decimal digits).
(0)
(34) Yk = ¥i-1 + X(y,-1, ty—-1) At,
(r+ 1)
(34’) Yi =y,-1 + EXY,-1, &-) + X(y??, t,)] At,/2
to be solved iteratively if necessary. In most cases, the full order of accuracy of
the implicit equation [O(h?) in the present instance] is achieved after one
iteration!
The improved Euler method consists in computing the sequence of z, by per-
forming these two substitutions in alternation. In the case X(x, i) = F()) of quad-
rature, it is equivalent to the trapezoidal method.
Applied to the initial value problem of Example 4, the improved Euler
method gives the approximate solution tabulated in the last row of the table of
§7; in this example, the work was carried to three decimal places to reduce the
cumulative error to 0.01.
To apply the improved Euler method to first-order systems, simply substitute
vectors for scalars in formulas (34), (34’). We illustrate the procedure by an
example.
O = x2 )
(35) + 9%, —=
There is little hope that formal methods of integration will help in the compu-
tation, but a straightforward application of the improved Euler method enables
one to calculate an approximate solution.
224 CHAPTER 7 Approximate Solutions
The calculations to be performed give, for this example, the double sequence
of numbers x,, y, defined by x9 = yo = 0 and by the formulas
Each step requires squaring four numbers and performing 14 additions and
subtractions and four multiplications.
For a subdivision into intervals of constant length h, the relative error com-
mitted in using the improved Euler method is Oh’), provided that the function
X(x, #) is of class @?. For, expanding the exact solution x(#) of dx/dt = X(x, t) by
Taylor’s Theorem with remainder, we have
h2
h?
oY
(38) De = Me + hY,-1 +
—_—
2
Eay
Ox
+ —
Ot
ln + Oh’)
where Y,_, denotes X(y,_, ¢,—1), and so on, The relative error committed in sub-
stituting (38) for (37) is thus O(4°).
A more explicit error bound is deduced in §10.
The improved Euler method has the advantage over the trapezoidal method
of being explicit. Various other explicit methods about as accurate as the
improved Euler and trapezoidal methods can also be constructed. For instance,
one can use the following adaptation of the midpoint quadrature formula:
2
’ ty-1 + A2 ) h= At,
This midpoint or modified Euler method is about twice as accurate as the trape-
zoidal and improved Euler methods in the special case dx/di = F(i) of quadra-
9 The Modified Euler Method 225
ture, as a comparison of formulas (17) and (26) shows. (In this case, the
improved Euler method is the trapezoidal method.)
But for first-order DEs generally, no such simple error comparison holds. To
see this, let X = Lb,tix be expanded into a double power series, and let x(t)
satisfy x = X(x, é). Then, just as in Ch. 4, §2, we have ¥ = X, + XX,, and
d°x
at
= Xy + 2XXq + X°X,, + XX, + XX
Expanding out to infinitesimals of the fourth order, we find that the exact solu-
tion of the DE dx/di X(x, #) for the initial condition x(0) = 0 has the
=
=
expansion
WC n°(B + B*)
(40) x(h) = hbp + — + + O(n’)
2 6
WC W(2B + B*)
(41) yh) = hbyy + — + + O(n)
2 4
giving a truncation error h3(B/3 + B*/12) + O(h*). With the improved Euler
approximation (34)-(34/), we get
WC WB
(42) z(h) = hbo + — +— + O(n)
2 2
with error h3(2B — B*)/6 + O(h*). With the midpoint approximation (39), we
finally obtain
WC WB
(43) w(h) = hbo + — + va + O(n’)
2
where X, denotes X(x,, t,) and X denotes 8X/dt + X 8X/dx. Dropping the last
cerm, we get a corrected trapezoidal integration formula, which may be expected
to have a cumulative error of only O(|7|°).
For instance, when applied to the inhomogeneous linear DE ¥ = 1 — 2éx of
Example 4, in which X = —2x — 2t + 4x”, this formula gives the approximate
recursion formula
At?
= x, 1[] — tj, Ate + (—1 + 2-17) Ate/6] + At, +=e
EXERCISES E
In Exs. 1-4, compute approximate function tables on [0, 1], with At, = 0.1, by the
improved Euler method for the following initial value problems:
1. x = —tx, x(0) = 1 2,%= (1+), x0) =0
3. x = y, 7 = 0, x(0) = 0, 9(0) = 1 4.%= 9,9 = x+y, x(0) = 0, y(0) = 1
In Exs. 5-8, compute approximate function tables for the data of Exs. 1-4, using the
midpoint (or modified Euler) method, instead of the improved Euler method.
11 Show that through the terms computed in Ex. 10, the improved Euler method gives
the same result as trapezoidal integration for co, ¢), cg.
All the methods for constructing approximate function tables that have been
described in this chapter have had one feature in common. Namely, the kth
entry in the table has been constructed from the immediately preceding entry
alone, the (k — 1)st entry, without reference to the earlier entries. Such methods
are called one-step (or “two level’’) methods.
Given a one-step method for numerically integrating the DE dx/dt = X(x, 6),
that is, for constructing an approximate function table with entries y, = y(,),
10 Cumulative Error Bound 227
Here¢is the function expressing y, in terms of y,_, and the data of the problem.
Bounds for the errors associated with one-step methods can be obtained by
using the following general theorem, which applies equally to the Cauchy poly-
gon method, the trapezoidal method, and the improved Euler and midpoint
methods.
let the relative error at each step be at most «. Then, over an interval of length T, the
cumulative error is at most (e/L)(e“7 — 1).
Proof. For any partition a, let ¢, denote the error introduced at the kth step.
That is, if £,(é) is that exact solution of the given DE satisfying the initial condi-
tion %,(¢,) = », where y, is the value of the computed approximate solution at
ts let
By the definition of “relative error,” we have «, < ¢ At,. The magnitude of the
cumulative error is, by definition,
But | %,(t,,) — %,-1@n)| is the magnitude of the difference, at t = ¢,,, of two solu-
tions of the given DE that differ by ¢, at t = ¢,. By Theorem 2 of Ch. 6, this is
at most
since ¢, << ¢ At,. Here ¢ is an upper bound to the relative error. Summing over
k, we get the following upper bound to the cumulative error:
But the final sum is the Riemann lower sum approximation to the definite inte-
gral fim exp (—Lt) dt = [exp (—Lto) — exp (—Lé,)]/L. Hence
eee? — 1]
[Yr(tm) — %(tm)|
L
(11?
1X max) Ate
12[1 — (L At,/2)]
XImale“ — 1)
[171 f jx] =-
(46)
12L[1 — L|x|/2] L
Similar error bounds can be found for the midpoint and improved Euler
approximate integration methods.
OX/dt + X 0X/x, we see that the order of accuracy of this approximate solution
is also O(|7|?).
EXERCISES F
1. Let x(@) and y(#) be approximate solutions of the system (1) with deviations ¢, and
€, defined for a <= t = b. Show that (7) implies that
€ + &€
|x — xO S [x@ — p@le
4 + et“! — 1)
L
2. Let F(x, y) and G(x, y) be everywhere continuous; let F satisfy a Lipschitz condition
with Lipschitz constant L, and let | F(x, y) — G(x, y)[ <= K. Show that if f(x) and
g(x) are approximate solutions of the DEs y’ = F(x, y) and z’ = G(x, z) with devia-
tions ¢ and 7, then
K+e+n
If) — g&)) = FO — g@ie" + (er4| — 1)]
10 Cumulative Error Bound 229
*3 Assume that, for equally spaced subdivisions of mesh length h, the truncation error
of a given approximate method J,[f] is Mh" + O(h"*'), where M is independent of
h. Prove that the extrapolated estimate
*5 For the DE dx/dt + tx = 0 and the mesh length A, = 1/(10]é,| + 1), show that
the truncation error of the trapezoidal method tends to zero as t — 00, regardless
of the initial value x(0). What is the limiting truncation error as t > — 00?
*6 Let f() be an analytic function, and f(¢ + 1) = f(@. Show that trapezoidal integra-
tion of fj f@ dt for h = 1/n has an infinite order of accuracy as n — 00. [HINT:
Expand /() in Fourier series.]
xh
h
J ~
; F(x) dx = 200 {114F(x) + 84[F@e +h) + Fe — A]
z)
(*)
20
P(e) de = = [24K + AUF, + F) — (Fa + FO
(b) Infer that, if F(z) is a complex analytic function, (*) holds with an error that is
O(n’).
*Q Let F(x, y) be bounded and continuous on the strip 0 = x <1, —0 < y < +00;
let {x,} be any sequence of partitions of [0, 1] with |x,[ — 0; and let f,(x) be the
Cauchy polygon approximate solution defined by z,, for the initial value (0) = 0.
(a) Show that, if the f,(x) converge to a limit function f(x), then f(x) is a solution of
y = Fu, y).
(b) Show that, in any case, a uniformly convergent subsequence { f,)(x)} can be
found, n(i) < n(i + 1). [HinT: See Ch. 6, §13.]
*10 In Ex. 9, show that, if the DE y’ = F(x, y) admits only one solution for the initial
value (0) = 0, any sequence of Cauchy polygon approximations defined for (0)
= 0 by partitions whose norms tend to zero must converge to the exact solution.
CHAPTER 8
EFFICIENT NUMERICAL
INTEGRATION
1 DIFFERENCE OPERATORS
(1) a=
d
X(x, 0)
of ordinary DEs. The simplicity of the methods considered, all due to Euler,
facilitated a rigorous theoretical analysis of their errors. In general, they had
O(h?) accuracy for X € @?.
In this chapter, we will describe some more efficient methods having higher
order accuracy, usually O(h*) for X € @*. We will explain the guiding ideas that
motivated the construction of the algorithms used but will not push the analysis
to the point of getting rigorous error bounds. This is partly to avoid lengthy
discussions of complicated formulas, but mostly because errors are usually esti-
mated in practice by studying the numerical output.
Such higher order methods are almost always used in practice when more
than two or three significant digits are wanted. If their errors are accurately and
reliably known, the errors should be subtracted to obtain improved results, as
in Richardson extrapolation (Ch. 7, §5).
Like the schemes already analyzed in Ch. 7, many of the schemes for numer-
ical integration to be studied later will refer to an assumed partition 7 of the
interval [a, 6] of integration by a finite number of points (the mesh),
‘<t,=b=a+T
(2) Ta=itgp< ti <ig<:
Typically, the partition is made into steps At, = t, — t—, of constant length h,
so that ¢, = a + rh; we then speak of a uniform mesh.
On the mesh (2), the DE (1) is approximated by a suitable difference equation.
This difference equation is then solved step by step in hand computations, using
ordinary arithmetic supplemented by readings from available function tables. In
machine computations, however, function tables are usually replaced by simple
subroutines that give accurate approximations by rational functions.
230
1 Difference Operators 231
fats
=
=
In the preceding formulas, the symbols A, V, 6 stand for linear operators that
transform functions into functions. Unlike the linear differential operators of
Ch. 2, §5, they apply to all functions
These operations are useful in obtaining approximate solutions because they
yield approximations to the derivative f’(x). If f € @', the derivative J’(x) is the
limit of the difference quotients
VS =lim of
i’) =lim +" = lim
0 mo h ano h
For obvious reasons, these are called the forward, backward, and central divided
difference approximations to f’(x)
The difference operators (4a)—(4c) can be applied to anyfunction table defined
on a uniform mesh with step h, consisting of the equally spaced points
(5) X, = Xo + th r=0 +2
=
> >
h>0
232 CHAPTER 8 Efficient Numerical Integration
(6b) VF
(x) = V(VF(x)) = VOf(x) — fx — h)) = flx) — fle — h) + flx — 2
(6c) Sf(x) = fx + h) — 2f(x) + fle — hy
We easily verify the identities 5°f(x) = A(Vf(x)) = V(Af(x)).
In formulas (6a)—(6c), the exponent 2 describes the effect of applying the
operators of formulas (4a)—(4c) twice, or ‘‘squaring” them. More generally, we
can form polynomials (with constant coefficients) of difference operators, like
A® — 3A + 2. Such linear difference operators with constant coefficients com-
mute (are permutable); those with variable coefficients do not commute (cf. Ex.
A4 below).
2. POLYNOMIAL INTERPOLATION
+ It is also inconsistent with the usual notation At, = ¢, — ¢—, employed in writing Riemann sums,
used in Ch, 7,
2 Polynomial Interpolation 233
1f second differences are tabulated, as they are in many tables, to use this for-
mula requires only two multiplications and three additions.+
In a similar way, we can derive the quartic (fourth order) interpolation formula
(9) f(xg
+k)=ye x( yn+ Hy
+ A(R? *) | (8°y312n3
— d°y1)
|
Here 5+y = 67(8°y9) = yg — 4y3 + Bye
— 49, + 90
The preceding formulas are based on central differences. For polynomial
interpolation between n + 1 successive values on a uniform mesh, one often
uses the Gregory—-Newiton interpolation formula.
o)
(xo)
(10) p(x)
= f(xo) + > A*f(%
ie
hX(k!)
| Th - =~3]
where n = (x, Xo)/h, and where A* is the iterated forward difference operator
This formula gives an approximation to f(x) in terms of the differences of n
equally spaced values of f/ This formula is a difference analog of Taylor’s for-
mula, without a remainder term
showf that there exists a unique polynomial p(x) of degree n or less which sat-
isfies p(x,) = y, for k = 0, 1 , n, that is, which assumes the n + 1 given
values at the points specified. Let
_
Q) =[[«-
plx)
=
(x — X) S#k
Pxlx)
Dy
(11) p(x)
= >4ie
pals) >" = Qe)>:Faw
Wene2
+ We do not count the division required to calculate k/h, since this requires only a decimal-point
shift in most tabulations.
Dy, _ Pi)
pilX)
=a
Pils
x)
k=O ~ pyle) 7h ~ J
Formula (11) is the Lagrange interpolation formula. Since the polynomial p(x)
is (11) is unique, (11) is equivalent to (10) ifx, = a + jh,j = 0,1 n. Hence
as interpolants to a function tabulated at equal intervals, (10) and (11) have the
same error
EXERCISES A
1. (a) Show that, if the function f(x) = a,x" is a polynomial of degree m, then A" =
vp = "fF = h"f™, where f denotes the mth derivative of f.
(b) Show that, if y € @’, then 6 = O(h’) and A = O(h’)
2. Define the divided differences [up, u;] and {tp, w, Ug] as (2, — tUt9)/(x; — Xo) and ([u,
Us] — {uo, u1])/(%2 — Xo), respectively. Show that lim,,,,, xote [to U1, Ug] = u(x)
3 Interpolation Errors 235
Show that
rl
ay a Br — Rune
Show that, if f(x) = x?, then x(Af) = Qhx? + hx, yet A(xf) = 3x7h + Sxh? + HP.
Solve the AE u,,; = 2u, — U,-1, for the initial conditions wy = 1, u, = —1.
Fv =F, + Fy-1
Estimate the largest 4 such that parabolic interpolation in a six-place table of log, x
on 2 =x = 3, with mesh length 4, will yield five-place accuracy.
Show that, with parabolic interpolation between y(—A), (0), and y(h), the maximum
error is normally near x = +h/ V3 and is about h? if’) | /9V3 there.
Show that the cubic Hermite interpolants to given values of f(x) and f’(x) at the end-
points of the intervals (a@—h,a) and (a,a+h) define a “cubic spline” function f ¢ @?
[a—h,a+h] if and only if
*10 Let h = (6 — a)/nand x, = at+ih. Prove that, given y,[i = 0 . ,n] and 94,¥;,, there
is one and only one cubic spline function s(x) ¢ @7(a,6) which satisfies: (i) the inter-
polation conditions s(x,)} = y, for i = 0, ... ,n; (ii) s(a@) = 6, s(6) = yy; and (iii) is a
cubic polynomial in [x,_),x,] for i = 1,...,n
(HinT: Use Ex. 9.]
*11 (a) Show that if s(x) is the (piecewise) spline interpolant to given y, and 9, 4, specified
in Ex. 10, and fix) = s(x) + v(x) eC 2(a,b) is any other interpolant with the same
properties, then
(Hint: Show that f s” (x)u”(x)dx = 0 for all continuous piecewise linear func-
a
tions v”(x).)
#3 INTERPOLATION ERRORS
In this section, we shall do much better, by giving explicit expressions for the
magnitude of the error, by formulas involving an appropriate derivative of the
function. More precisely, let a function f(x) be tabulated at n + 1 points x) <
x< * + <x,, and let p(x) be the unique interpolation polynomial, of degree
n or less, which satisfies p(x,) = f(x,),k = 0,1... , n. How big is the error of
p(x), considered as an approximation to f(x)? An answer to this question, when
J (x) is sufficiently smooth, is provided by the following result.
(x — Xo) °° (x — ,) fOMw}
(12) S(*) — px) =
(n+ VD)!
Proof. Let e(x) = f(x) — p(x) denote the error function. Since p(x) is a poly-
nomial of degree n, p"+ (x) = 0. Therefore
for all x, and e(xo) = e(x;) = - + - = e(x,) = 0. Consider now the function
Q(x)
f(x) — pe) = ex) =
_—_
(n + 1)!
f° * %, q.e.d.
< N,
If) — plx)| nm
~ (n+ 1)!
When the mesh points are equally spaced, we can compute N,, explicitly.
Thus, we have N, = h?/4; if x) < x < x9, N3 = 9h*/16, and so on. Moreover,
similar arguments show that for k <= n, the error in the kth derivative of the
Lagrangian interpolant is O(h""**?). It follows (though we shall not prove it) that
one can develop multipoint AEs that approximate DEs to an arbitrarily high
order of accuracy.
4 STABILITY
The accurate numerical solution of initial value problems for ordinary DEs
(and systems) involves much more than interpolation error bounds and esti-
mates. For one thing, the relative error involves stability considerations. These
are most easily explained in the special case of linear DEs with constant coeffi-
cients, previously discussed in Ch. 3.
+ See J. F. Steffensen, Interpolation, Williams & Wilkins, Baltimore, 1927, pp. 35-38 and the refer-
ences given there.
238 CHAPTER 8 Efficient Numerical Integration
For each root p, of this characteristic equation, the sequence y, = p,’ is the
solution of the linear AE (14), which satisfies the initial conditions yp 1, M
= =
= =
m—1
Ps +> » Ym-1 = Pr
For example, consider again the DE y’ = y for the initial condition y(0) = 1,
whose exact solution is y = e*. In this case (16) reduces to
For h = 0.1 and yp = y(0) = 1, the exponential series truncated after five terms
gives y, = 1.1052, rounded off to four decimal places. Substituting into (16’)
we can compute the approximate function table for y, = exp (7/10):
and so on. After 10 steps, this gives the approximate value e 2.714, whose
=
=
For A = 0.1, this gives p, = 1.10499, pg = —0.90499 when rounded off to five
decimal places. The first root differs from the growth factor e°' = 1.105171 of
the exact solution by about 0.00018. This is about 0.27% of 0.105, which
explains why the O(4”) approximation (16) gives only two-digit accuracy for
h= 0.1.
4 Stability 239
Stability. By analogy with Ch. 3, §4, we will say that the homogeneous linear
nth order AE with constant coefficients (14) is stable when all solutions y, are
bounded sequences and are strictly stable when all solutions are sequences tending
to zero, as r — ©,
Since we can obtain a basis of solutions of (14) of the form r’p,’, the AE (14)
is strictly stable if and only if all roots of the characteristic equation (15) are less
than one is absolute value. This condition is obviously necessary; it is sufficient
because lim,_.., 7p" = 0 whenever |p| < 1.
The concept of stability brings out a significant aspect of the effectiveness of
the central difference approximation (16) for integrating numerically the DE
y’ = y. The general solution of (16) is Ap,” + Bps’, where p, = h + (1 + hh?!”
as above, and A, B are arbitrary constants. The positive root p;, which is approx-
imately equal to e°| is dominant in the sense that |p,| > |p9|. Therefore, the
term Bp; can be neglected in comparison with Ap} for large 7, provided that
A # 0.
Example 1. For the DE y’ = —y, the situation is reversed: the (central) dif-
ference approximation y,4; — 9,—1 = —2hy, to the stable DE y’ = —y is unstable.
Thus for h = 0.1, we have p,; = 0.904988, p. = —1.10499. Although p,
approximates the exact growth factor p = 0.90584 reasonably well, it is domi-
nated by the “extraneous root” py which is introduced in approximating a first-
order DE by a second-order AE. As a result, the “approximate solution”’ will
x
ultimately grow like e*, whereas the true solution decays like e~
whose roots are £1 and 4 + V15. One of these is near 8, so that errors tend
to grow by a factor 8 per time step. The AE is thus very unstable!
EXERCISES B
?
In Exs. 1-4, verify the formulas indicated for f € @*, a uniform mesh with step A,
0<6<1,and6 = (1 — 4).
4. f(x + 6h) =
=
3(Yo + 9) +
6 — Yip — 6/4)6 + 8)
[00(20 — 1)/12] 6,2 + O(h*) (Bessel)
5 Find the truncation errors of the formulas of Exs. 1-3, for quartic polynomials
q(x) = a + bx + cx* + dx” + ex
(a) Find the cubic polynomial ¢(x) that satisfies c(0) = yo, (0) = yo, c(h) = 1,
"(h) = 9%
(b) Derive your formula asa limiting case of the four-point Lagrange interpolation
formula
Test the following AEs for stability or instability, by calculating the roots of their
characteristic equations
(a) Un+l Qu, Unt (b) Unt) =
Uy + Uy
Up-]
*8 Show in detail that |@ + ab] + 6° < 1 is a necessary and sufficient condition for the
strict stability of the AE 9,49 = @yn+; + by,
Derive necessary and sufficient conditions for the strict stability of a general linear
third-order AE with constant coefficients
Using Taylor series with remainder, one can in principle obtain approxima-
tions to the derivatives of tabulated functions having arbitrarily high orders of
accuracy, from suitably designed difference quotient and divided difference
formulas
For example, whereas the usual forward difference quotient formula Af/h has
only O(h) accuracy, and even the central difference quotient formula 6f/h has
only O(h?) accuracy, formula (3) of §1 has O(h*) accuracy. Likewise, we have the
truncation error estimates
Lew fe (x)h?
(19a)
a + Lor
— f'®) =
6 24
(at r
@r*
fO
(19b) —f'®)=
+"T920
where £ is in the interval over which the difference is being taken. This illustrates
the general principle that central difference quotients give more accurate
approximations to derivatives than forward or backward difference quotients of
the same order. We can obtain truncation error bounds similarly:
Note that the interval in (19c) is twice as long as in (19b); hence, the truncation
error is multiplied by about four.
Similar approximations can be made to f”(x), using second difference quo-
tients. For f € @?, we have
(20) SF
— WF" (x) = A*f?(®)/12, x—-hs&ESxth
he?
Sf’) —_—
ef
~ T9 ”(6)
—_=
he
o"f nh?
oa SO ®),
(21) fO@) gel
oo e_
h”
n/2 n/a
of(x) = ph @ F 9 de Sf(x) = f “e dt pl ttt w du,
h/2 h/2
a"f(x) = f . dt, f =-
A/2
anf pl &t th +++: +4,)dt,
where m and M are the least and greatest values of f*®(g) for & in the given
interval. Since
4/2 3
h/2
i,” dt, = —
12
and since f“+®(g), being continuous, assumes all values between its extreme
values m and M, formula (21) follows.
nMh?
(22) [f%e) ~ Of/h"| = 24
where M is the maximum of |f*(&)| as & ranges over the given interval.
5 Numerical Differentiation; Roundoff 243
(23)
and tabulated values are correct to $ in the last decimal place, the roundoff error
is at most one in the last decimal place.
In numerical quadrature formulas, which also have the form Dw,y, per step
with Dw, equal to the mesh length, the maximum roundoff error is similarly
bounded by the length of the interval multiplied by the maximum tabulation
error (ordinarily 3 in the last decimal place).
However, the effect of roundoff can be dramatic in other cases, as Example
2 of §4 demonstrates. The truncation error is zero in this example; hence if
=1l
h =
8 or some other binary fraction, the computer printout will also be exact.
But if h = 0.1 (which is not binary), the small initial roundoff error is amplified
by a factor 8 at each step, and dwarfs the true solution after 10 or 20 steps.
Empirically, roundoff errors are nearly independent, and randomly distrib-
uted in the first untabulated decimal place with a mean nearly zero. Hence,t the
cumulative roundoff error has a roughly normal distribution on a Gaussian
curve, and the probable cumulative roundoff error with n equal subdivisions is
only O(1/ Vn) times the maximum cumulative roundoff error.
Similar results hold for the numerical integration formulas to be considered.
The roundoff errors may be thought of as “noise,” superimposed on the system-
atic truncation error. Both are amplified in the course of the calculation by a
L(b—a)
factor of at most ¢ , where L = sup OF/dx is the one-sided Lipschitz con-
stant, and (6 — a) is the interval of integration.{ The reason for this is that, for
Las defined above,
It is interesting to compare the truncation error bound (19c) with the cor-
responding roundoff error bound, which is 10~"/2hA if an m-place table is used.
When h < (3/10"|f” | max)'”8, therefore, the roundoff error exceeds the trun-
cation error. To minimize the sum (|f” | max#?/6) + (107"/2h), which is the max-
imum total error if both terms have the same sign, set 2h° = 3/10" |f” | max: This
shows that the maximum total error with parabolic interpolation into an m-place
table cannot be reduced below
1 VIB" lmn/10™
and that one loses accuracy if |ff:,| < 2, by choosing h smaller than 0.04, ifa
four-place table is used to approximate f’(x) by a central difference quotient.
For example, for the function f(x) = sin x, where f(x) = sin x, since f(x)
ranges between —1 and 1, the maximum truncation error is approximately h?/
12, and the maximum roundoff error is 2 X 1075/h®, using five-place tables. To
minimize the greater of the truncation error (which tends to zero with h) and
the maximum roundoff error, we must make h?/24 = 1075/h. Hence, we min-
imize the maximum total error of f” =~ 5*f/h? near h* = 2.4 X 1074 or h &
0.13 radian = 8°, roughly, a surprisingly large interval!
Roundoff errors are not considered further in this chapter. This is partly
because, with high-speed computing machines, truncation errors are usually big-
ger unless h is very small (most modern machines carry at least ten decimal dig-
its), and partly because the analysis of roundoff errors involves difficult statistical
considerations.
EXERCISES C
1 Show that the effect of roundoff errors on tenth differences is bounded by about 500
* 107” in n-place tables.
Show that hf’(x9 + 4/2) = 8y12 — (1/24)8%y,2 + O(F').
Show that hf’(x9 + h/2) = dy12 — 8°y12/24 + 5°912/1920 + OH’).
Given a six-place table of sin x (x in radians), show that the approximate formula
{3 & &f)/h has a combined truncation and roundoff error bounded by 2/10°h? +
#?/12, and that this expression has a minimum of about 0.0008, assumed for h
about 0.07.
(a) Show that h2f” = Of — d4f/12 + O(n’).
*(b) Show that hf” =
ical quadrature. In this section, we will derive a rigorous error bound for Simp-
son’s Rule (Ch. 1, §8), by a method which is applicable to a wide variety of
numerical quadrature formulas. In the next section, we will discuss two other
remarkable quadrature formulas, which seem to have no analogs for other DEs.
Let 0 = % <7, < 7, S 1be given, and let x; = a@ + 7;h, h > 0, so that
(25) aSx%<x<-+-++<x,Sath=)d
For any function f € @"*®, let p(x) be the Lagrange polynomial interpolant of
degree n to the y, = f{x;) at the x; Then the approximation of f(x) = p(x) is
associated with a formula for numerical quadrature, namely
wyf(x)
i=0 ix=0
(26) t
=
_ f pix)dx, =p(x) =I]
p jri
(x — x,)
Formula (25’) is exact for all polynomials for degree = n, since then f(x) = p(x).
For other functions, the error in (25’) is — J° e(x) dx, where —e(x) = p(x) —
f(x). Hence, by Theorem 1, the error in (26) for any given n is of the order of
h”*? at most. We have proved the following theorem.
THEOREM 3. For any choice of real numbers 1, with 0 < ry <1) < ‘<n
< 1 and weights w, defined by formula (26), we have, for all f € @"*? [a, b):
I (x — x9) dx =
x) — Xp
Se
w= =
xO XX) — Xo
ath
(27)
a
fe) dx ~ - [le + fle + 1 =F (0 +90)
246 CHAPTER 8 Efficient Numerical Integration
is obtained as a special case of (25) and (26). In this special case, the error e(x)
satisfies by Theorem 1, applied to the linear interpolant L(x) to f(x)
x1
A(yo + 1)
J xo
L(x) dx =
2
x}
sf f th-ojdts
a
min
e(x) dx = —
2
w”
max
[aoa
Xo
1 3
f x(t) dt — (xy + x)
-( 410
is not equal to —x”(r)/12 = —(6r, 127°)/12 —(r, 27°)/2 for any r in [0, 1].
6 Higher Order Quadrature 247
satisfying P(x) = Yo, P(%1) = Mis P(x) = M1 = fH), Pg) = Jo, with xo = x) —
h, x9 = x, + h. These conditions amount to a) = 9, @, = yj, dg = 5y, /2h?,
and ag = —4a,/h? + (yo — yo)/2h°. Hence they can be satisfied for any yo, 91, Yo.
To estimate the error e(x) = p(x) — f(x), translate coordinates so that x9 =
=
x?(h? _ xf »(¢)
(32) e(x) = > —h=ét=h
24
Integrating (32) with respect to x, since x(k? — x*) = 0, it follows that the
truncation error in using Simpson’s rule for quadrature over —h < x Sh lies
between m = min f"(é) and M = max /"(€) times the definite integral
* xn? — x*)dx WP
J =
h 24 ~ 90
248 CHAPTER 8 Efficient Numerical Integration
Since f € @*, f'(€) assumes every value between m and M, which gives the fol-
lowing theorem.
THEOREM 4. Iff(x) € @”, the truncation error for Simpson’s rule on the interval
—h=x S his equal to hef(é) /90 for some & in the interval [—h, h).
The relative truncation error is therefore h*f'’(£)/180. For example, to
achieve five decimal places of accuracy in computing In 2 = {7 dx/x, about 10*
points must be taken if Riemann sums are used, about 100 with trapezoidal
quadrature, whereas 10 are sufficient using Simpson’s rule!
*7 GAUSSIAN QUADRATURE
In formulas (25) and (26) for numerical quadrature, with the x, as free param-
eters, we have 2n + 2 adjustable constants in all. This suggests the hope that by
properly locating the x,, we can get a formula which is exact for all polynomials
of degree 2n + 1 or less, since these form a (2n + 1)-parameter family.
Such a formula was obtained by Gauss; in deriving it, it is convenient to
renormalize to the interval (—1, 1) and to label the points é,, . » Em, SO that
m =n + 1. The formula uses two properties of the Legendre polynomials P,,(x)
defined in Ch. 4, §2, which will be proved in Ch. 11, §6. These properties
are:
(i) P,,(x) has m distinct zeros x = &; << &<- ++ < €, in the interval (—1, 1),
whence p,y(x) = Cy(x — &)(x — &) + + + (x — &,) for some constant ¢,,.
Proof. By centering the origin and change of scale, we can assume the
interval of integration to be [—1, 1] without loss of generality. Let f(x) be any
polynomial of degree 2m — 1 or less; let p(x) be the Lagrange interpolation
polynomial of degree at most m —1 satisfying p(é,) = f(&,), 7 = 1, 2,...,m
Then e(x) = f(x) — p(x) vanishes at &,, » &m- Hence,t we have e(x) = (x — &)
++ + (x — &,)b(x), where (x) is a polynomial of degree at most m — 1.
f e(x)dx f 5(x)P,,(x)dx = 0
=
=
so that (25’) and (26) are exact for e(£) by the choice of the £,. Moreover, Gaus-
sian quadrature is exact for p(x) by the choice of the &; hence, it is also exact
for the given f(x), completing the proof.
Let f(x) € @?"[a, a + h], and let q(x) be the polynomial of degree 2m — 1 (or
less) satisfying 9(x;) = f(x;), for x, =
=
a + [jh/(2m + 1)],j = 1,..., 2m. Then,
by Theorem 1, we have e(x) = q(x) — f(x) = O(h?") and so
i- ae) in - oe) dx
=
=
f ™ e(x)dx = O(h?"*)
But, by Theorem 5, we have
m
gal gol
ath m
(33)
a
f(x) dx — ° wfla + 1h) = OW?"*)) if r, _(d-¥§
J
2
joi
and let T® = [4"7T&*) — TE) iJ/(4" — 1). Then the T? converge extremely
—
+ See F. L. Bauer, H. Rutishauser, and E. Stiefel, Proc. XV Symposium on Applied Math., Am. Math,
Soc. 1963, 199-218.
250 CHAPTER 8 Efficient Numerical Integration
EXERCISES D
1. Using a five-place table of sin x, x in radians [but not tables of Si(x)], evaluate
*5 Show that
11
J's as =5 E +f ~ eh + Bf) + 720 (“fo + a) + O(h")
*6 Show that, for n = 3, the Gauss quadrature formula on (— A, h) is
fora =4 0+ -Eos—90
is h°f(£)/720, where 0 < E <A, ify = f(x) € @*.
(Simpson’s Five-Eight Rule). Show that, if f€ @, then
8 FOURTH-ORDER RUNGE-KUTTA
n(n)
h
(34) (x, + h) = 9 + hy +
ye4 —_—_
+ O(h"*)
2 n
ignoring the remainder O(h"*}). The relative error in the recursion relation so
obtained
2a
h Ye 4. h‘yf
(35) Yer1 = MH hy +
—
2 (n!)
dx
(36) — = X(x, 2), ast=sb
dt
+Zeits, Math. Phys. 46 (1901), 435-453; C. Runge and H. Kénig, Numerische Rechnung, 1924, Ch.
10.
252 CHAPTER 8 Efficient Numerical Integration
me
(37) k, = Xi; 4), ky = X(x + hk,2 —,iitr-
h
2
ls = X(y+ hk,
2
+e2 ). ky= X(y;+hks,t, +A)
where the mesh length h = A; may vary with 2.
We now show that the preceding Runge-Kutta method has an error of only
O(h*) per step.t For simplicity, we restrict attention to the first-order DE
and to the initial condition x(0) = 0. Formula (37) reduces in this case to
a X(y 4), hg
=
=
x(4 2°
hky
+t
2
(39’)
hks
x(n 2”
hs
2 J. X(y; + hks, t, + A)
where x) = x(0), %1/2 = x(6), and subscripts x stand for partial derivatives. Using
primes to indicate total derivatives with respect to t, so that X’ = OX/dt +
X 0X/dx, we get
+ O(h'*)
2 2 6
so that
we obtain
Similarly, we have
and
so that
hy = X(x, hb) — 4X(x1, AX" (10, O° + X,(e1, MX _(2e12, OX" 1/2, 0? + O(h4)
254 CHAPTER 8 Efficient Numerical Integration
we see that
EXERCISES E
1. (a) Derive a power series expansion for f(a + h) through terms in h® for the solu-
tion of y’ = 1 + 9? satisfying fla) = .
(b) Truncating the preceding series after terms in h?, evaluate approximately in
three steps the solution of y’ = 1 + y? satisfying (0) = 0, setting x, = 0.5,
Xo = 0.8, x3 =
=
1. What is the truncation error? [HINT: Consider tan x.]
3. (a) Apply the Picard process to the DE y’ = x” + 9? for the initial value yp = 0 and
initial trial function y (O) = 0. Calculate the first four iterates.
(b) Using the power series method of the text, calculate the Taylor series of the
solution through terms in x!7, and check against the answer to (a).
(c) Evaluate (1) numerically at x = 1, using the preceding truncated power series,
and compare with the answer of Ex. 2.
8 Fourth-Order Runge-Kutta 255
4. (a) Apply the Runge-Kutta method to the DE y’ = 1 + ¥? for the initial value
(0) = 0, setting x9 = 0, x; = 0.5, xy = 0.8, x3 = 1.
(b) Same question for the DE y’ = x? + »°, with (0) = 0, and the same mesh.
5. For the first-order linear system dx/di = A(/)x, show that the Runge-Kutta method
is equivalent to
6. (a) Show that the system u’ = v, v’ = —wis neutrally stable, and indeed that |(u(2),
v())| = const. for any solution.
(b) Show that the Runge-Kutta method is strictly stable, and satisfies
6 8
9. Show that, with midpoint integration, y(h) is given by the approximate formula
3B
yulh) = hboo = © + bobo) + — + orn
2
B* ) + or
yh) = hbo + 9 (bio + Boobor) + 3
(
=+
2 4
12. Check the formulas of Ex. E8 against those of Exs. 9-10 above in the special case
t-; = 0,4, = h, and AW = pl) = fo + pil + pol? + of a first-order linear
DE.
*13, For the linear DE dx/dt = p(x, p) = Lae pt — a)*, evaluate x(a + h) through
terms in A® by the Runge-Kutta method. Compare this with the Taylor series for
the exact solution.
*9 MILNE’S METHOD
A very different method for solving initial value problems with fourth-order
accuracy is due to W. E. Milne. Whereas Runge-Kutta methods are based
directly on power series expansions, and the Euler methods of Ch. 7 (and Ch.
1, §8) basically approximate derivatives by difference quotients, Milne’s method
replaces x’(f) = X(x; ¢) by the equivalent (vector) integral equation:
(40) x(t)
=
=
x(a) + f X(x,s) ds
[15 + xp40)
[3 + (15 — x) — 40419041]
=
Dr+2
=
Evaluating (0.1)
=
=
+ W. H. Milne, Numerical Solution of Differential Equations, John Wiley & Sons, 1953.
9 Milne’s Method 257
The truncation error is about 107°, as can be verified by use of the power series
solution
2
=
x—- tx +— xP +--
3 15
= > aaa Gon+)
Aap]
=
2k + 1
Starting Process. Given the initial value yo c = f(a), one must compute
=
=
4 by a one-step method before one can begin to apply a two-step method. For
analytic F, it is usually best to calculate y, = f(a + h) by expanding f(x) in a
Taylor series as in Ch. 3, §7. WhenF is not analytic, but fairly smooth (say, if
F € @*), good approximations to f(a + h) are often obtained by repeated mesh-
halving of the interval [a, a + h], using a one-step method with a lower order
of accuracy. For instance, we might first compute fla + h/8) by midpoint inte-
gration, and then use Milne’s formula to get f(a + h/4) = 91,4 from yp and 9g.
Next we compute 9; 2 from yp and 4, 4 by a second application of Milne’s formula
with mesh length h/4, finally getting y, from yp and 9j/2 by a third application of
the same process.
Iterative Solution. Although (41) can be solved algebraically for linear DEs
and systems, for nonlinear DEs one must resort to iterative methods to compute
X,+9 from x, and x,,;. One can do this by a method analogous to Picard’s
method of successive approximation (Ch. 6, $7), as follows. First rewrite Milne’s
equation (40) in the following form
Xi+2 =
|x,(+1)
+2
(”)
~ Kare
=_
9)
For h < 3/L, 6 < 1 and so the sequence of x h+2 is a Cauchy sequence; let x,,5
be its limit. Moreover U is a contraction which shrinks all distances by a factor 6
or less, and so is continuous. Hence, passing to the limit on both sides of the
(+l)
equation x,k+2 =
U(x’?..), we get (42).
Milne’s method (41) is evidently a two-step method in the sense that each new
value of an approximate solution is computed using the two preceding values.
In this section, we will study Milne’s method more critically, and describe other
multistep ““Adams-type’”’ methods.
Multistep methods are usually best executed as predictor-corrector methods,
in the sense of Ch. 7, §8. An explicit “predictor” formula based on extrapola-
tion is made to yield higher order accuracy by one or two iterations of an implicit
“corrector’’ formula.
Milne’s Predictor. Thus with Milne’s method, we can use the predictor
which has O(h*) absolute, and O(h?) relative accuracy. To get O(h*) relative
accuracy from this, we must iterate twice with the corrector (41). Alternatively,
for k = 2, one can use the four-step (five level) predictor
4h
(43 Zerg = Z—g+ ( —
3
{2X,_) — X, + 2X41}
which has O(h*) accuracy, and apply the corrector (41) once.
Ahp 3—h
—__ =9
(44) pe t+
3+h 3+h
+
O(h°). This corresponds to relative O(h*) accuracy. Unfortunately, the other root
po = —1 — h/3 + O(h?); hence, the magnitude of the error will grow exponen-
10 Multistep Methods 259
tially like e? with alternating sign. Therefore, computed values will ultimately
oscillate with increasing amplitude, whereas the exact values tend smoothly to
zero.
L
(47) M = |6"
|max(@”— 1/901( 1-—
Finally, we can express ¢(x) in terms of F and its derivatives, just as in Ch. 4,
§8:
¢’ =
=
and so on. Combining these results, we can compute M a priori in terms of the
values of F and its derivatives, thus getting an explicit error bound.
260 CHAPTER 8 Efficient Numerical Integration
A priori error bounds, like the preceding, are seldom useful for methods hav-
ing O(h*) accuracy. One reason is that they are so complicated. In practice, reli-
ance is usually placed on a posteriori error estimates, which utilize computed val-
ues. Another reason is that they neglect roundoff errors.
EXERCISES F
1 (a) Show that, for the DE y’ = y and hk = 0.1, Milne’s method amounts to using the
AE y49 = Bly, + 4yn41)/29.
(b) Integrate the DE y’ = y from x 0 to x =
0 and y, =
=
0.00033.
4 (a) Show that, if y’ = F(x, y), where F € @?, then y = f(x) satisfies the two-level mid-
point formula (16) with discrepancy O(h’).
(b) Show that, if F € @* and y,_), y, are exact, the truncation error is Kk? + O(h’).
(a) Integrate y’ = y approximately by the two-level midpoint method with h = 0.1,
taking yy = 1, y, = 1.1052 as starting values and integrating to y;9 = 2.7145.
(b) Estimate the discrepancy and the cumulative truncation error in (a), comparing
them with the roundoff error.
(a) Do the same as in Ex. 5a but for the system y’ = z — 2y, z’ = y — 2z and the
starting values yy = 1, % = 0,9, = 0.8228, z, = 0.0820, computed by the Taylor
series method.
(b) Show that the system in question is stable but that the approximating AE is not.
Explain why the computed table is approximately correct, although the method is
unstable.
he
y=1, natn 2 6 24
—
Show that the truncation error of (A) is O(h‘) per step, while that of the implicit
method (Aj) is O(h’).
CHAPTER 9
REGULAR SINGULAR
POINTS
1 INTRODUCTION
over the complex plane. In the rest of that chapter, however, we assumed the
independent variable to be real.
In the present chapter, we shall consistently be considering complex-valued
functions w = f(z) of a complex independent variable z, and the behavior of
such functions as z varies in the complex domain. Specifically, we shall usually
be studying functions that satisfy some second-order, linear homogeneous ordi-
nary DE of the form
+ Ahlfors, p. 24; Hille, p. 72. Some knowledge of complex function theory is assumed in this chapter.
261
262 CHAPTER 9 Regular Singular Points
solution of the DE, in a domain D that includes every point of the complex z-plane
except the isolated singular point at z = 0; hence, it includes both real solutions.
The general solution of the same DE is the complex-valued analytic function
w = 1/(c — z). This function has an isolated singularity at z = c.
The DE dw/dz
=
=
(1)
_ yw
dw ’ y=at #, a, 6 real
dz z
(1’/)
=
When 6 = 0 and y = ais real, the analytic continuation of the real solution
u
—
=
when z describes a simple closed counterclockwise loop around the origin, mak-
ing 8 increase by 27 and In z by 277. This example shows that solutions of a linear
DE can have branch points where the DE has a singular point,} even though the
DE has single-valued coefficient-functions.
For positive z x > 0, a basis of real solutions is provided by the real and
=
=
imaginary parts of the functions z’ = e”'" *, where ¥ is either of the roots of the
indicial equation
(2’) P+ (p-lvytq=0
are vy = +i. A basis of complex solutions, real on the positive x-axis, is therefore
provided, as in Ch. 3, §3 by the real and imaginary parts of the functions
The analytic continuation of the solution cos (In x), real on the positive x-axis
6 = 0, through the upper half-plane to the negative x-axis 6 = a is not the
solution cos (In |x|) = cos (In 7) given in Ch. 3, §3 but is the complex-valued
function cosh z cos (In r) — 7 sinh sin (In 7).
z/w. The general solution of this DE, obtained by separating variables, is the
two-valued function w = (z? — c*)'/””, which has branch points at z =
=
te. Since
c is arbitrary, the general solution has a movable branch point.
There is no significant class of nonlinear first-order normal DEs whose solu-
tions have fixed singular points. However, the solutions of the generalized Ric-
cati DE w’ = fo(z) + pi(z)w + po(z)w® have fixed branch points.t This can be
shown by representing w = v’/pov as a quotient of solutions v of the linear DE
v” + pi’ + popov = 0, as in Chap. 2, §5.
A second-order nonlinear DE having a fixed singular point at z = 0 is
w” = (w?/w) — (w'/2),
The study of singular points of linear DEs in the complex domain begins with
first-order DEs of the form
(3) w’ + p(zjw = 0
+ The Riccati DE is the only first-order nonlinear DE with fixed branch points.
3 First-Order Linear Equations 265
exp[—Jp(z) dz]
oo
A}
3”) C exp[—a_,Inz — >
k=] ( )e+
k3 (He)
As a corollary, we can represent the solution w in the form
r once
a(z) = 1 — age +
(aj — a2" +
In this case p(z) is said to have a branch pole of order ¥ at (Q, 0), and the DE (3)
to have a regular singular point at z = 0. The number + can be real or complex,
rational or irrational.
Finally, if p(z) has a pole of order exeeding 1 or an essential singularity at z
= 0, then the DE (3) is said to have an irregular singular point at 0. In this case,
one can show that g(z) in (4) has an essential singularity at z = 0. For example,
the DE w’ + z7?w =
=
Q has the solution w =
=
exp[1/{] = Upio z “/Al, easily
found by separating variables, with an essential singularity at z = 0.
Toprove this, suppose the contrary: that [z’g(z)]’ + p(z)z’g(z) = 0, with
aw
Not all branch points are simple; for example, the functions In(z — zo) and
(2 — %)* + (% — 2%)? have branch points at z zo but do not have simple
=
=
branch points there, unless a — 6 is an integer. Any branch pole (4) is a simple
branch point, but a simple branch point need not be a branch pole; thus,
consider z!/2e71/” at z = 0.
3 First-Order Linear Equations 267
We shall now give another proof of the first result of Theorem 1, namely that
every solution w(z) of (3) that is not a holomorphic function has a simple branch
point at z 0
=
=
the DE does not have a branch point there; the DE w’ = w/2z is a case in point.
However, the general solution of the DE (3) has the form cw(z); from this it
follows that
252
wz) = wie z) = cw(z), c#0
erm
Now, write c , where @ is a suitable complex number, and consider the
=
=
g(z) = 27 7w(2)
This shows that g is a single-valued function in the punctured disc A. Thus, the
function w is the product of z* and a function without a branch point.
The idea behind this second (and deeper) proof of the first result of Theorem
1 can be applied to linear DEs of any order, as we show in the following sections.
EXERCISES A
1. Show that no solution of w’ = 1/z that is real on the positive x-axis can be real on
the negative x-axis.
2 (a) Setting z
=
=
re®, discuss the analytic continuation to the negative x-axis of the
solutions z and z Inz of z’w” — zw’ + w = 0 on the positive x-axis.
(b) Show that no nontrivial solution of z2w” + 3w/8 = 0 that is real on the positive
x-axis can be real on the negative x-axis.
Let p(z) be holomorphic and single-valued in |z| < p except at points a and b. Show
that any solution w, of (3) can be written in the form
where fis single-valued and holomorphic in |z{ < p except at a and bd.
Generalize the result of the preceding exercise to the case that p(z) is single-valued
for |z| < p and holomorphic at all points except a, + Dye
can have branch points only where the p,(z) have singular points.
d"w
*) = Fw, w',w”,... ,w-”, z)
dz"
With this theorem in hand, let w,(z) and w(z) be a basis of solutions of the
second-order DE
where the functions p and q are single-valued and analytic in the punctured disc
A: 0 < |z| <p. Analytic continuation of each of these solutions counterclock-
wise around acircle |z| =
=
y < p with center at the origin yields two functions
29%
(in general different): @(z) wy,(ze ) and wo(z) w,(ze""’). These are, by the
Continuation Principle, also solutions of the DE (5). Since every solution of (5)
is a linear combination of w, and w,: the continued functions w,d can be
expressed as linear combinations of the solutions w, and w», thus
QR
@(z) wy(e Z) = 41, (zZ) + a)qW9(z)
252
Woz) Wole Z) = ag W,(z) + aggwo(x)
—1 0
( 0 —1
wo O
(0 w
aly
where again all the coefficient-functions ,(z) are holomorphic in the punctured
disc A
LEMMA. Givena basis of solutions wz) of (6), analytic continuation of the (2)
around any circle |z| =
=
of (6)
Proof. By the Continuation Principle, the w,(z) satisfy (6); since the w,(z) form
a basis of solutions, the result follows.
The matrix A = |{a,|| so defined is the circuit matrix of the DE (6) relative
to the basis w,(z), , w,(z). It represents a linear transformation of the vector
space of all solutions of the nth-order linear analytic DE (6), in the following
manner. If
is the general solution of the DE, analytic continuation of w around the same
circuit y carries w(z) into the solution
(6/) wW(z)
=
=
wer™z) = y C,,w,(Z)
pk=l
(7) w(z)
—
=
2*flz)
5 CANONICAL BASES
0
if and only if it is carried into a constant (scalar) multiple of itself by continua-
tion around the circuit y. From the discussion in the preceding section, we see
that a solution w(z) = Xj.) ¢w;(z) of (6) has a simple branch point at z = 0 if
and only if the vector ¢ = (¢, 9, ,» €,) and the circuit matrix A satisfy the
+ Birkhoff and MacLane, p. 293. As stated there, A is a root of the characteristic equation |A =
MI = 0.
5 Canonical Bases 271
relation cA = dc, where the constant A will then be necessarily different from
zero. In other words, a linear combination Lj., cjw, of solutions of (6) has a
simple branch point at z = 0 if and only if the vector c is an eigenvector of the
circuit matrix A associated with the basis (w,, wo, , w,) of solutions. Thus,
there are as many linearly independent solutions of (6) with simple branch
points as there are linearly independent eigenvectors of the matrix A.
We shall now look for a basis of solutions with simple branch points for any
second-order linear DE (5), with coefficients holomorphic in A.
Given two linearly independent solutions w, and wy of (5), we can construct
the circuit matrix A = ||a,|| as in §4. The linear combination w =
=
CW,(2) +
€9W(z) then will have a simple branch point if and only if Dc,ay, = de, By the
theory of linear equations, this system of equations has a nontrivial solution if
and only if the following determinant (the characteristic equation of the circuit
matrix A) equals zero:
Ordinarily, this characteristic equation has two distinct roots A,, A». These
roots give two linearly independent solutions F(z) = ¢,w,(z) + cgwo(z) and
G(x) = d,w,(z) + dgwo(z), having simple branch points: Fez) = d F(z) and
G(e""z) = A»yG(z). Relative to the canonical basis of solutions F, G, the circuit
matrix is thus a diagonal matrix
A 0
( 0 dg
As in §4, F(z) =
=
z*flz) and G(z) = z8g(z), where A; = e?™*, dy = 6?™8, and fand
g are holomorphic in A. Such a basis is called a canonical basis.
When the characteristic equation has a single solution \, the solutions may
stillt sometimes have a basis of the form (7). Every solution of the DE is then
2
2a
multiplied by the same nonzero constant A = ¢ when continued around a
counterclockwise circuit ‘.
Otherwise, we choose a basis as follows. Let w,(z) be the solution of the form
z*flz), where f is one-valued in the punctured disc whose existence was estab-
lished in Theorem 2, and let w9(z) be any other linearly independent solution.
Continuation of w.(z) around the circuit y gives, as in §3, wo(ze?™) = aw,(z) +
bw,(z). The circuit matrix for this basis of solutions is therefore the matrix
A 0
(
=
=
a b
+ This occurs when the matrix A is a multiple of the identity matrix. Otherwise, any 2 * 2 matrix A
with only one eigenvalue is similar to a matrix of the form . This fact is not assumed in the
1 A
present discussion.
272 CHAPTER 9 Regular Singular Points
Since the eigenvalues of such a “triangular” matrix are \ and , and since the
only eigenvalue of A was assumed to be A, we must have 6 = A and
dA 0
). a#0
A
(
=
=
a,x
The continuation around the circuit y of the function h(z) = w(z)/w;(z) is easily
computed to be
We(z)
hie") =< + =
= + hz)
r WwW,(z) r
Ai® = A(z) — 4 In z
Qrin
is single-valued in 0 < |z| < p and, therefore, that the function w,(z) can be
written in the form
The functions f(z), g(z), and f,(z) ave holomorphic and single-valued in the punctured
disc
0 < [z| <p.
more generally, consists of zeros except on the main diagonal and (in the case
of a repeated eigenvalue) just above it. If the eigenvalues of the circuit matrix
are Aj, do, ., A, the continuation of the solution w, around a small circuit y
is given by one of the two formulas
By Theorem 3, there always is at least one solution that goes into a multiple of
itself. If all the eigenvalues of the circuit matrix are distinct, then the circuit
matrix can be reduced to diagonal form by suitable choice of a ‘‘canonical basis”
of solutions, and the exceptional case of formula (10b) does not arise.
dw
ea tw Co d* wy
(11) L{w] = dz"
+—=w=0
z dz”!
The trial function z’, with unknown exponent », satisfies the DE if and only if
the exponent » is a root of the indicial equation of Ch. 4, §2,
(12) CC)
=
=
vpy~-1)--+-@—nt))
+evpe—-—D-- YV-—ntAaterr+
te wt c 0
=
=
When the roots of the indicial equation are distinct, the z” are a canonical basis
of solutions of the Euler DE (11). The circuit matrix is the diagonal matrix with
diagonal entries }, e exp 2xiv,, where », is a root of the indicial equation.
—
=
When»is a k-tuple root of the indicial equation, then the functions z’ log z,
z’(log z)", and so on, formabasis of solutions. This basis is not “canonical” when
n > 2 (see Ex. B2), though the circuit matrix for (11) is always triangular relative
to it.
EXERCISES B
1 0
(0 1
for which the functions f and g in Theorem 4 have (a) essential, and (b) removable
singularities at 0.
2. (a) Compute the indicial polynomial for the homogeneous Euler DE zw” + 32z2w”/
2 + zw’/4 = w/8.
(b) Compute the circuit matrix of the preceding DE relative to the solution basis
z/?, z'(log z), and z'(log z)”.
274 CHAPTER 9 Regular Singular Points
FindaDE(5)withcircuitmatrix(0 1/A
for which the functions f and g in Theo-
rem 4 have (a) poles, and (b) essential singularities. Can f and g have removable
singularities?
Show that the requirements 0 S Re{a}, Re{@} = 1 uniquely determine the exponents
a and @ in formula (9a).
Show that, in the exceptional case of Theorem 4, the eigenvalues of the circuit matrix
are equal.
Show that eV*, e~V” forma solution basis for the DE w” + }w/ — kw = 0. Compute
the circuit matrix for the given DE relative to this basis, and also find a canonical
solution basis.
circuit matrix has the form , such that fin Theorem 4 has an essential singu-
1 1
larity at z 0
=
=
Many of the ordinary DEs of greatest interest for mathematical physics have
singular points which are ‘“‘regular’’ in the sense of the following definition.
DEFINITION. A second-order DE
analytic for 0 < [z — z| < p, has a regular singular point at z) when p(z) has at
worst a simple pole at z Zo, and q(z) at worst a double pole there.+
—
=
In the next several sections, we shall show how to adapt the power series
methods introduced in Ch. 3 to solve ordinary DEs in the neighborhood of any
regular singular point. In particular, we will show that a singular point of the
second-order linear DE (13) is “regular” if and only if the functions f(z) and g(z)
of the canonical basis (9a), of solutions constructed in Theorem 4, have at worst
branch poles there.{ Equivalently, the condition is that a basis of solutions have
the form (25) below.
+ That is, # may either be holomorphic (have a removable singularity) or have a simple pole, and q
may be holomorphic or have a pole of first or second order.
{Or, in the exceptional case of (9b), poles times logarithmic branch points (order of growth log
r/r).
6 Regular Singular Points 275
We now show that, near the regular singular point z = 0, there always exists
a formal solution of the DE (13), namely, a formal power series of the form
which, when substituted into (13), satisfies the DE. To calculate the coefficients
c, and the exponent p of (14) it is convenient to rewrite (13) in the form
where P(z) = LfioP,z* and Q(z) = XfoQ,z* are convergent for |z| <p. Sub-
stituting (14) into (15) and equating to zero the coefficient of z’, we obtain the
indicial equation
(16) Iv) = vv — 1) + Pw + Q = 0
for the exponent v. The roots of this equation are called the characteristic expo-
nents of the singular point, and J(v) is called its indicial polynomial.
Equating to zero the coefficients of the higher powers of z, namely
ann y+n
z 3% ,+.., we obtain the relation
[~ + ly + Poe+ 1) + Qole; + Pv + Q, = 0
and, recursively,
n—-1
Since the left side of the preceding equation is [(v + n)c,, the equation can be
written in the form
The above equation for the coefficient c, can be solved recursively for ¢,, Cy, ¢3,
..., except in one case: when, for some positive integer n, both vy and (» + n)
are roots of the indicial equation. By taking a characteristic exponent having the
largest real part, we can make sure that I(v +- n) does not vanish for any positive
integer n, even in this case. We therefore obtain the following theorem.
af a(a + 166 + 1) 2
Fa,
8, y;2) = 1 +maet
(20) y¥y¥+)) 2!
ac + 1)(a + 2)B(8B + 16 + 2) 2? +
vty + 1) F+ 2) 3!
From the Ratio Test, it follows that the radius of convergence of the series is at
least one, and is exactly one unless a, 8, or 7 is a negative integer. This also may
be expected from the existence theorems of Ch. 6, since the radius of conver-
gence of a solution extends to the nearest other singular point of the coefficients
of the hypergeometric DE, which is at z = 1. The function F(a, 6, ; z) defined
by the power series (20) is the hypergeometricfunction, to be studied in §10 below.
7 BESSEL EQUATION
The Besse] DE has a regular singular point at the origin, with indicial equation
Iv) = v? ~ n®? = 0. In physical applications, n is usually an integer or half-
integer. But, for theoretical purposes, it is interesting to let n® be an arbitrary
complex number.
By Theorem 5, we can compute a formal power series solution beginning with
z", of the form z"(1 + cz + coz” + - - -). From the recursion formulas (17) or,
more simply, by direct substitution into the DE, we obtain the recursion rela-
tions for the coefficientsc¢,:
( )-
1 1
of z z
n+1 (_
2 (2!)(n + 1)(n + 2) 2 |
The series in square brackets is an entire function (convergent for all finite z).
Multiplying by the normalizing factor 1/2I'(n + 1), we obtain the Bessel function
of order n, already discussed in Ch. 4, §4:
1 z 1 Zz
(2}[- n+1 (_
()-
1 z
(22) :
(2(n + 1)\(n + 2) 2 |
The Bessel function J, is an entire function (Ch. 4, §5) if n is a nonnegative
integer. Using the functional equation for the gamma function, Iz + 1) =
zI(z), this formula can be recast in the form
co
(- 1)*(z/2)"**
we (— 1)*(z/2)-"**
J-.©
=>
k=O |
Tk + DE(—n + k + 1) |
This solution has a branch pole at the origin. Unless n is an integer, J, and J_,
form a canonical basis of solutions of the Bessel DE.
Jnl2)b"2) (2)6@) = —
We)
_ A
gi) =
Tee) nz)
(23) Z,(2)
=Jul) | +A f Fi,|
= K,z-""4{1 + Die b,x") (cf. Ch. 4,§8],
for any indefinite integral of 1/z]2(z)
where K”is a suitable constant. Hence, we have
dz
(23’) =KA [xvin fy + Yo dz
Jr(2)
0, we easily
compute 1/zJo"(z) = z7' + 2/2 + 523/64 + , and so the general solution
of (21) when n = 0 is
2
524
(23”) Zo(z) = to B + A[ Inz+>+ 5+:
256
)
This shows that any solution of the Bessel DE of order zero that is not a constant
times J,(z) is logarithmically infinite near z = 0
More generally, the preceding formulas show that, when the parameter n in
the Bessel DE (21) is not an integer,} we obtain the general case (9a) of Theorem
4, with a = n and @ = —n. When nis an integer, we have the exceptional case
(9b) of Theorem 4. We shall now discuss this exceptional case
+ Note that, though the Bessel DE of half-integral order x + } has characteristic exponents which
differ by an integer, it has a basis of solutions of the form (9a). This is because the recurrence rela-
tions for the coefficients in its expansion express ¢ 49 as a multiple of ¢, without involving c,,,
7 Bessel Equation 279
will give a possible second member of the canonical basis (9b). Thus, when n =
0, we can choose A = 1, B = 0 in (23”). Relative to this choice, the circuit
1 0
matrix is
2 1
(24) j= /3[o(o-8)+00)]
oe VE ( }+o(%)]
.
wv
xo
sin
Zari = —2"[2z-"Z,(z))’
Since the function Y(z) is of the form Y)(z) = Jo(z)Lfol) + Ko log z], where fp
is holomorphic and single-valued in a punctured disk with center z _
=
0, we
verify by straightforward differentiation that
Thus, all Neumann functions Y,,(z) have a branch point at z 0. From this it
=
=
follows that Y,, and J, are linearly independent, and indeed are a canonical basis
of solutions for the singular point z 0 of Bessel’s DE.
=
=
280 CHAPTER 9 Regular Singular Points
When pv
=
=
n is an integer, let Y,(z) = lim,_,,Y,(z). The limit can be evaluated by
PHOpital’s Rule as
van
1t can be shown that this definition of the Neumann function for integral n
agrees with that above
cw k
od'l dl
J dy? ty — (+ n)1=0
EXERCISES C
1. (a) Show that the DE zw” + (1 — z)w’ + Aw = 0 hasa regular singular point at the
origin, and that y = 0 is a double root of the indicial equation.
(b) Find the power series expansion of the solution of this DE that is regular at
z = 0. (Hint: Derive a recursion formula for ¢,41/¢
q
(*)
ool q )}e+[r-(t)ealens
has a regular singular point at z = 0
(b) Show that, if w,, wy are a basis of solutions of (*), then w{, w are a basis of
solutions of (**)
8 The Fundamental Theorem 281
*4 Show that, if the roots of the indicial equation at a regular singular point of (5) differ
by an integer, the eigenvalues of any circuit matrix are equal.
Let (13) have a regular singular point at z 0, and let a be a root of its indicial
=
=
equation having largest real part. Show that, if w is a solution, then v = z~*w satisfies
a DE of the form (15) with Q) = 0 and Re{P)} = 1.
6. (a) Given three pairs of nonzero complex numbers (A, Ag), (41, Ha), (21, 29), Construct
a holomorphic second-order linear DE (13) having regular points at z, and 2,
whose circuit matrices at these points have eigenvalues (A;, Ag), (41, Hs).
*(b) Generalize the above to n points 2), » Xp
7. Show that, for x a nonnegative integer, J,(z) and its complex multiples are the only
solutions of the Bessel DE that are holomorphic at the origin.
a JnlZ)
~(-1)"21-26)
| a on
10 Show that, if u), uw. and v;, Us are bases of solutions of the Besse] and modified Bessel
DEs, respectively, %,, tg, U;, Vg form a basis of solutions of the DE
(2n? + 1) 4 — n’)
w® + (z2 ) a _— emt 2] w+ (n
J»-
Zz w+ —1
11 Show that the self-adjoint form (Ch. 2, §8) of the hypergeometric DE (18) is:
We now establish the fact that the formal power series solutions obtained in
§6 are convergent. We begin by proving the converse of this result.
have simple branch poles at z 0 of different orders a # 8. Then the normal second-
=
=
for the unknown coefficient functions p, q, as in Ch. 2, Ex. B7. The result is
p=
—(wywz — ww) _ (wjuh = whut
’
(8 ~ a)ztt8) (: + > a)
and does not vanish near z = 0, since a # 8. The numerators are the powers
yotB-2 a+B-3
and z multiplied by holomorphic functions of z. Dividing out, we
get
where P(z) and Q(z) are holomorphic in some neighborhood of z = 0. This com-
pletes the proof of the theorem.
THEOREM 7. Let the second-order linear DE (5) have a regular singular point
at the origin, and let a be the larger root of its indicial equation Iv) = 0. Then the
formal power series (1 + Xa,z") of Theorem 5 converges to a solution of (5) in a
domain
0 < |z| <o, o> 0.
It will be recalled that, in Theorem 5, » was any root of the indicial equation
Iv) = 0 such that Jv + n) = 0 for no positive integer n.
Hence, we have
M|yp| + N
Pal lol + 1Ql ’ kz=0
From formula (16) we obtain the bound |I@ + ”| = n?/K for some constant
K = 1: since Iv + n) # 9 for all nonnegative integers n by hypothesis, the
sequence n“/|1(» + n)| is bounded (it tends to 1 as n — 00). Therefore, if
the recursion formulas (17) give the following bound for the coefficient c, in
the formal power series solution (14)
M|yp| +N cA
len| =
<=
£5 (k=0
n
") n—k
£Falal
Jeal = ARSlal
leak
(25/ lc,
|=
n m0 nM 4x9?”
KA AK 1
lc,| = —— >
nN 4=0 ( p
n—k lcol
= OOo
[1 +(AK)
+ (AK)’ + + (AK)"""]
and this shows that the formal power series solution (14) has a radius of
convergence at least equal to p/AK. The proof is therefore complete, with
o = p/AK.
For this, one only has to choose a canonical basis of the form (25), which
exists by Theorem 7
Exceptional Case The exceptional case, namely when the roots of the
indicial equation differ by an integer, can be treated by the following method
Select for a a root of the indicial equation having maximum real part. Then
8
=
=
a —~— n for some integer » 2 0 and so, since by the indicial equation
a+ 8 = 1 — Po, it follows that 2a — Po 1, where Po is the leading
coefficient of p(z) Po/fz +P) + and n is a nonnegative integer
Moreover, we know that I(a@ + ) # 0 for all integers n > 0, and so by Theo-
rem 7 the given second-order linear DE (5) has a solution
w, = zflz) = (1 + Yan!)
which has a branch pole at z 0 and is nonvanishing in the punctured disk
—
=
0 < |z| < o for some o > 0. Hence, if we set w = wih = z"f(z)h(z), the DE (5)
is equivalent to
Po =
Cz In
z + z-"d{z) ifn #0
where ¢(z) is holomorphic. This shows that the exceptional case of Theorem 4
always occurs when the indicial equation has a double root and also when the
roots differ by an integer n, unless ¢c, = 0
Collecting results, we have proved (for C = c,) the following theorem
9 Alternative Proof of the Fundamental Theorem 285
Relative to the canonical basis (26), the circuit matrix has the form
dX 2xiC
(0 r
Theorem 7 also can be given a more intrinsic proof by relying on the follow-
ing characterization of poles of analytic functions.
Theorem 7 can be deduced quite easily from this and the following basic
result.
Proof. For any solution w(z) of (5), consider the real-valued function
Setting z = re”, we shall majorize the derivative of this function relative to r for
fixed 6; its differentiability follows by the Chain Rule.
“Vv aul =
<{ |V(t)| dt
Applying this inequality to U(re) we obtain
ze 72
1] au
| = wu +|2 |
+ [z*w'w"
where z
=
=
Using the fact that w”
=
=
The functions P(rei) and Q(re6) are holomorphic in some closed disk 0 = |z|
<= R, R > 0. Let M be a common upper bound for their absolute values, for
0 = 6 = 2m and for fixed r. This gives the inequality
1|%] « ww'| + (M
+ 1)[2w?
|
<= (M + 1)|ow
| fou
(aU = (4M+4) U(re®) _ KU
)] | <
rT
K>0
r
R¥U(Re®) — U(re*) = 0
and hence,a fortiori, that | w(re") |? < (NR*)r-*. By the Order of Growth Theo-
rem, with C = NR* and a = K, the conclusion of the lemma follows
Consequently, if the DE (5) has a regular singular point at the origin, then it
has a solution given by a locally convergent power series of the form described
in Theorems 5 and 7 The construction of a second solution can then be
achieved as in Theorem 8
The preceding result can be generalized to nth order linear DEs
10 Hypergeometric Functions 287
EXERCISES D
_G-= ay
C +1
3
(j+1P
with
A = n— (k — 1)/2,B = —4,C = (1 — h9)/4.
Show that, if 6(0) # 0, the substitution w = ¢(z)v carries second-order linear DEs
(5) having a regular singular point at the origin into DEs having the same property.
Do the functions log z and (log z)? satisfy a second-order linear DE (3) with a regular
singular point at z = 0? Do they satisfy a third-order linear DE with regular singular
point at z = 0? Justify your answers.
*10 (a) The DE w” + C3 p,(z)w®-” = 0, p,(z) holomorphic for 0 <= |z| < 7, has a
regular singular point at z =
=
0 if p, has, at worst, a pole of order k. Derive an
analog of the indicial equation (16) and generalize Theorem 7 to this DE.
(b) Generalize Theorem 8 for this DE when two exponents coincide.
So far in this chapter, the behavior of solutions of DEs has been studied only
neara single isolated singular point. A fascinating topic of analysis is the relation
between the behavior at different singular points of analytic functions defined
by DEs. This topic is beautifully illustrated by the hypergeometric functions,
288 CHAPTER 9 Regular Singular Points
21 ~ zw” + fy +1 —-@+t1+B+1+4+
Dqw" —-@+)E+ lw =0
= =
the solution w)(z) = F(@;, 81, ¥:; z), we obtain at once a power series solution
of (18) corresponding to the exponent 1 — y in the form
The two solutions are a canonical basis of solutions of (18) at the regular singular
point z = 0.
The change of dependent variable w = (1 — z)’~*~8u gives another hyper-
geometric DE of the form (18) in the variable u with a, = y — a, 8; = y — B,
11 The Jacobi Polynomials 289
and
It carries the regular singular points 0, 1 of (18) into 1 and —1, respectively.
Note that the Jacobi DE (31) goes into itself under the transformation z +> —z,
a= b.
Multiplying by (1 — z)*(1 + 2)’, we get the self-adjoint form (Ch. 2, §5)
—
d
d:
la — 21 + 2)! a +nn+a+b+ 1)(1 —(1 + 2)’u = 0
When a = 8, this reduces to the ultraspherical DE
+ Assuming, of course, that the parameters a, 8, y are not chosen in such a way that the solutions
coincide: thus y # a + 8.
290 CHAPTER 9 Regular Singular Points
Ta
+ n+ 1) (a—1/2,a—1/2)
PI) = Tati/2+n+1)""
(2)
An expression for the Jacobi polynomials often more convenient than (34) is the
Rodrigues formula
pe )( )= 1"
(-1)"
(36)
ni2”
qd —- yd + yd — ond + 7]
We shall derive this formula from the identities for the hypergeometric function
established in the preceding section. First, since (1 — #)* F(—a, b, 6, t), the
11 The Jacobi Polynomials 291
bin
binomial series is a special case of the hypergeometric series: (1 — #
=
Fa+ 1, —n — b,a + 1, t). Using also the Jacobi identity, we justify the first
two steps of
nm
71 — or ern _ ‘?t")
=
=
In the last step, identity (30) for the hypergeometric function is used.
The Rodrigues formula (36) follows by making the change of variable
t= (1 — 2/2.
EXERCISES E
| z—|z—+y7-1
az
2
(z— +a
\ z—+8
}
(b) Show that the eigenvalues of the circuit matrix of the hypergeometric DE at z =
0 are equal if y is an integer.
(c) Show that the eigenvalues of the circuit matrix for z = 1 are equal if y — a —
6 is an integer.
(a) Show that, if @ is zero or a negative integer, the hypergeometric DE (18) has a
polynomial solution unless -y < a is a negative integer.
(b) Using (34), express this solution as a Jacobi polynomial.
[1 — 2*)w’! + Aw = 0
*(b) Describe corresponding circuit matrices, taking as basic solutions an even and
an odd solution.
Derive from (31) the self-adjoint form of the Jacobi DE displayed in the text.
( n I 2
(Hint: Show that the right-hand side satisfies (31), using suitable identities for F.]
Find the roots of the indicial equation of the Jacobi DE (31) at z = 1 andz = —1,
10 Show that (34) defines a solution of (32) even when n is not a positive integer. What
happens when7 is a negative integer?
0, form #n
d’v 2 1 1 1
a
(37)
dt” | i (} _—
t di
f?
(po
where v(t) = w(1/t). The point at infinity is said to be a regular singular point
of the DE (5) when the origin is a regular singular point for the DE (37). This
happens when the function
2 1 1
| (2 a }
—_
t
—
has, at worst, a pole of the first order at ¢ = 0, that is, when the first coefficient
in the power series expansion of p(1/t) vanishes. Also, the function ¢~*9(1/1)
must have, at most, a pole of the second order at ¢ = 0; this happens when the
12 Singular Points at Infinity 293
first two coefficients in the power series for q(1/é) vanish. This gives the follow-
ing theorem.
THEOREM 9. The point at infinity is a regular singular point for the second-order
linear DE (5) if and only if the coefficients p and q have power series expansions
convergent for sufficiently large |z|, of the form
ple) = PL 4 Bey
(38) ’ q) = f++-
That is, it is necessary and sufficient that the function p have a zero of at least
the first order and the function g have a zero of at least the second order at
infinity. In particular, the solutions of the DE are holomorphic at z = ©, or
0, if and only if the coefficients
2 1 1 1
J} ana (
| ( HK
=
t
2
_-
t
_—
tt
K _
COROLLARY. [If the coefficients p(z) and q(z) of the DE (5) are holomorphic for
sufficiently large z, then all solutions of (5) have removable singularities at z
=
oo if
and only if p,
= 2 and q2 = 93
= 0 in (38).
the indicial equation of (37) at ¢ = 0 has roots e and 6 not differing by an inte-
ger, then the DE (5) has a basis of solutions of the form
Its roots are called the characteristic exponents at z = 00. If they differ by an
integer, then there is still a solution of the form (1/z’)(1 + (a)/z) + ), but
every second linearly independent solution may contain a logarithmic term
convergent when |z| > 1. From the symmetry between @ and 8, we obtain a
second solution
wv+{> A,
ju =o
k=] * Zk
Second-order Fuchsian DEs offer much more variety; they are classified
according to the number of their singular points. When the number of these is
small, their study is greatly simplified by making linear fractional transformations}
of the independent variable, of the form
—(zt4) ad # bc
(a ta)’
form ¢ = (z — z,)/(@ — 29), we can send these singular points to zero and infin-
ity. 1t follows from the definition of a regular singular point and from Theorem
9 that the Laurent series of p(z) and g(z) reduce to p,/z and q/z”, respectively.
Hence, the most general Fuchsian DE of the second order with two regular sin-
gular points is equivalent to the Euler DE of Example 2,
w +ew+ By =0
0
and z = 1. It can therefore be written in the form
pl)
etna tne
—
=
where the function /;(z) is regular throughout the plane. However, by Theorem
9, the function zp(z) has a finite limit as |z| tends to infinity. Since the function
z[A,/z) + (B,/(@ — 1))] is bounded as |z| tends to infinity, it follows that zp, (z)
is uniformly bounded. By Liouville’s Theorem} it must, therefore, vanish
identically.
Similarly, the coefficient q(z) has at worst poles of the second order at z =
=
0
and z = 1, and can therefore be written in the form
A, As __ By_
+ +
3
qz) = at + qi(z)
Zz Zz «-1% z<-1
where the function q¢,(z) is holomorphic in the finite complex plane. By Theorem
9, the function z°q(z) remains bounded as |z| tends to infinity; hence, so does
the function
| As
Zz
+
_—
5 + ata} = 2 zz —- 1)
Therefore, Az; = —Bs and, again by Liouville’s Theorem, the function q,(z) van-
ishes identically. This completes the proof of the following theorem.
THEOREM 10. Any second-order Fuchsian DE with three regular singular points
can be transformed by a linear fractional transformation into the form
(40) w++
A,
z
=
z-]
__Bo As
@-1? 2z—-1
leno
where the A, and B, are constants.
The differential equation (40) is called the Riemann DE; it evidently depends
on five parameters.
With the Riemann DE are associated three pairs of characteristic exponents
Ai, Ag), (Hi, Ha), (1, ¥9), belonging to the singular points 0, 1, 0, respectively.
These exponents are the roots of the indicial equations [cf. (16) and (39)]
He — 1) + Bye + By 0
=
=
vy? + (1 ~ Ay — Bw + Ag + By — Az = 0
By 1—
m4 — be By
=_
My He
A, + B, =
yp ttl Ay + By — Ag =
VV
From the identities in the first column, we obtain the Riemann identity
(41) Ay + Ag + ey + Hg $y + vg = 1
w" +
1—dr, —~dg 1 ~ 4 — be /
(42) z z-l1
+
AiAg +
be VYg — Ajrg — Hills
0
=
=
2
z (2 — 1)? zz — 1)
The preceding discussion shows that the Riemann DE (40) is completely deter-
mined by the values of the exponents and the location of the singular points.
THEOREM 11. A Fuchsian DE of the second order with three regular singular
points in the extended complex plane is uniquely determined by prescribing the two expo-
nents at each singular point. The exponents satisfy Riemann’s identity (41).
13 Fuchsian Equations 297
0
and to y, — A at infinity. A similar result holds for the more general change of
dependent variable
v(z) = Zz ~— 1)w(z)
Proof. The general solution w(z) of the Riemann DE can be written in the
form
Y= 1— do + ALq.e.d.
EXERCISES F
1. Show that the only second-order linear DE that has just two regular singular points,
at 0 and 0, is the Euler DE w” + (fp/z)w’
+ (qo/z*)w = 0.
2 Show that no analytic linear DE (5) can have only removable singularities, if the
point z = 09 is included.
Prove in detail that any linear fractional transformation carries regular singular
points into regular singular points.
lf p and gq are constant in (5), is the singular point at oO regular? Justify your
statement.
ag
+ 5b.
298 CHAPTER 9 Regular Singular Points
Find necessary and sufficient conditions on p(z) for w’ + p(z)w 0 to have (a) a
=
=
v+| > Uk
Jeno
k=l @ — 4)
*9 Find the most general second-order linear DE (5) having regular singular points at
a, , 4, and 00,
*10. Find the most general linear DE having regular singular points at 0, 00 and no other
singular points. Show that any such DE can be integrated in terms of elementary
functions.
ADDITIONAL EXERCISES
Find an entire function f(z) and constant ¢ for which the functions
dz
w, = f(2)'?exp + { f
ea V20l — 2] |
are a basis of solutions of z(1 — z)w” + (1 — 2z)w’/2 + (az + bw = 0.
Show that this has a regular singular point at £ = 0, calculate the exponents, and
find a recurrence relation on the coefficients of the power series solutions.
*6 (a) If P and Q are given polynomials without common factors and if ¢,41/¢, = P(n)/
Q(n) and © ¢,z” is convergent, show that the function © ¢,z” satisfies the DE
2P(zd/dz) w — Q(zd/dzjw = 0.
(b) Find all quadratic polynomials P and Q for which the preceding DE has regular
singular points only, and express the solutions in terms of hypergeometric
functions.
Show that the function In (In z) satisfies no linear DE of finite order with holo-
morphic coefficients.
13 Fuchsian Equations 299
9. Show that, if (0) = 0 but (0) ¥ 0, the substitution z = f({) carries a regular
singular point of (5) at z = 0 into one at ¢ = 0 having the same indicial equation.
10. Show that, for any nontrivial solution of the Euler DE Zw" + ww +w =
=
0 and
any integer n, there exists a spiral path @ = A(r) approaching the origin, along which
lim...) [z"w| = 00.
Show that this singular point is regular if and only if, for some n > 0, every solution
satisfies lim, z-*w(re"*) = 0for0 <6 = 2a.
CHAPTER 10
STURM-LIOUVILLE
SYSTEMS
1 STURM-LIOUVILLE SYSTEMS
Here a, a’, 8, 8’ are given real numbers. We exclude the two trivial conditions
a=o'
= 0andé = @’ = 0.
2
d
du Ju =o, asr=xb
(3)
| I+(
dr “dr
Rr — —
are sometimes imposed and give another type of S-L system, a periodic S-L
system.
will be periodic with period 27. Moreover, since cos 2x = cos (—2x) is an even
function, any solution of (5) is the sum u(x) = $[@(x) + ¥(x)] of an even solution
@(x) = u(x) + u(—x) and an odd solution ¥(x). The Mathieu functions are suitably
normalized even and odd solutions of (5), of periods « or 27 [i.e., satisfying
(5’) or (5”)).
2 STURM-LIOUVILLE SERIES
Examples 1 and 3 define two S-L systems from the same S-L equation,
u” + du 0, but with different endpoint conditions. The eigenfunctions of
=
Example 3 are the functions used in the theory of Fourier series, studied in the
advanced calculus. It is shown there that these functions are orthogonal on
the interval —x < x < x. This means that the following relations hold.+
wT x
J sin mx sin nx dx
J cos mx cos nx dx = 0, ifm #n
_
=
Ww =
t
J =
WT
sin mx cos nx dx =
We will now show that analogous orthogonality relations hold for the eigen-
functions of regular S-L systems generally, and for the eigenfunctions of S-L
systems with periodic endpoint conditions.
(6) f p(x)f(x)g(x) dx = 0
I
The interval J may be finite and open or closed at either end; or it may be semi-
infinite or infinite.
(7) f p(x)u(x)v(x)dx = 0
Proof. We use the operator notation L[u] = [p(x)u’)’ —q(x)u. Then u and v
are eigenfunctions of (1) with eigenvalues ) and yp if and only if
x=b
To prove (8), we apply the hypothesis to the Lagrange identity of Ch. 2, §8,
namely the identity
E2a [u’(a)v’(a)—v’(a)u’(a)] = 0
Hence p(a)[u(a)v'(a) — v{a)u’(a)] = 0 unless @ = a’ = 0. Similar formulas cover
the boundary term at x = b. Since the possibilities a =
=
a’ = 0and 8 = p’
=
=
are excluded, this shows that the right side of (8) vanishes. Formula (7) now
follows from identity (8), after dividing through by the nonzero factor (A — y).
304 CHAPTER 10 Sturm-Liouville Systems
COROLLARY. The result of Theorem 1 holds also for S-L systems with periodic
endpoint conditions.
For, in this case, the contributions to the boundary term on the right side of
(8) from x aand x = bare equal in magnitude and opposite in sign; hence
=
=
they cancel.
It is shown in the calculus that any reasonably smooth periodic function f(x)
can be expanded into a Fourier series
EXERCISES A
1. (a) Show that every solution of the Airy DE v” + xv 0 vanishes infinitely often
=
=
the eigenvalues are (k + 12,4 +1 /2)%, (zk + 1/2)%, and Rk, respectively. What are
the eigenfunctions?
4. (a) Show that u = U(kr) satisfies (3) if and only if U(x), (x = kr), satisfies the Bessel
DE U" + (1/x)U' + [1 — (n2/x2)]U = 0.
3 Physical Interpretations 305
(b) Show that if U(x) and V(x) satisfy the Bessel DE and if
then f U(kr)V(k,)r dr = 0.
(a) Show that any two Mathieu functions having distinct eigenfunctions are
orthogonal, in the sense that
f “ u(x)v(x) dx =
=
(b) Show that the even Mathieu functions are the eigenfunctions of the regular S-L
system defined by (5) and
u(0) = w(x) = 0
Show that, if f(x) and f2(x) are eigenfunctions of Ex. 6 having distinct eigenvalues,
A, # do, then [5 fi(falx) dx = 0.
Show that the substitution £ = cos’ x transforms the Mathieu equation into
and one nontrivial side condition B[u] au(a) + a’u'(a) + Bulb) + B’u'(b) = 0.
=
=
*3 PHYSICAL INTERPRETATIONS
2 =
T
Mt = CM
ex where c =
Here y is the lateral displacement from equilibrium; T is the tension and p the
density of the string, both assumed constant. Simply harmonic standing waves
are defined by the separation of variables
For (x, #) to satisfy the vibrating string equation y, = c’y,,, it is necessary and
sufficient that u” + Au = 0, = k®/c?, where k depends on the endpoint
condition.
For the vibrating string, it is natural physically to have fixed endpoints, so that
y(a, t) = 9(b, t) = 0. This makes u(a) = u(b) = 0 and leads to the S-L problem
of Example 1. The eigenvalue belonging to each eigenfunction is proportional
to the squared frequency k®/4x*. This relation, combined with the analogy
between mechanical and electromagnetic waves, has led mathematicians to call
the set of eigenvalues the spectrum of an S-L system.
Another physical interpretation of S-L systems is furnished by the longitudi-
nal vibrations of an elastic bar of local stiffness p(x) and density p(x). The mean
longitudinal displacement v(x, f) of the section of sucha bar from its equilibrium
position x satisfies the wave equation
2.
0
p(x
—_=_—=«—_>
ot?
—
0
| pe) |
The simple harmonic vibrations (the normal modes of vibration) given by the sep-
aration of variables
d
dx
|p(x)“| + h'p(x)u =
=
0
+ Widder, pp. 413-421. In this section, subscript letters denote differentiation with respect to the
variable indicated.
3 Physical Interpretations 307
EXERCISES B
1. Show that, if U,(x) satisfies the Bessel equation of order n (Ex. A4), then ¢ = U,(kr)
sin nf? and U,(kr) cos n# satisfy the Helmholtz equation V7@ + k®@ = 0 for polar
coordinates in the plane.
The partial DE of a vibrating membrane is V?U + #2U = 0. Using Ex. 1, show that
this equation has solutions satisfying U(x, y) = 0 on x® + y? = 1, for all numbers k,,,
such that J,(Rmn) = 0.
3. A string of density she + ahs cos 2x grams/cm is stretched taut between pegs at x =
—a/2 and x = 7/2, under a tension of 2 kg. Determine its natural frequencies, in
cycles per second.
4. For the Bessel DE (xu’)’ + Axu = 0, with the endpoint conditions that u(1) = 0 and
508 CHAPTER 10 Sturm-Liouville Systems
u is bounded on 0 < x = 1, show that the first five eigenvalues are approximately
A, = 5.78, Ay = 30.5, Ay = 74.9, Ay = 139, and A, = 223. [HinT: Consult a table
of zeros of Jo(x).]
A vibrating reed, with one clamped end and one free end, executes simply harmonic
vibrations with transverse displacement y(x, t) = u(x) cos kt if and only if u® =
=
ku,
u(0) = w(0) = 0, and w’() = u” (2) = 0. Find the characteristic frequencies 27/k.
Show that the function J,,,(e* Vb/c) satisfies the DE u” + bce** — d®)u = 0.
Show that the function x/,,,(e/* Vb/c) satisfies the DE
wt x74 (be**/*
a) =0
is the function v(x) = VxU(bx’), whereUis the general solution of the Bessel DE of
order n.
4 SINGULAR SYSTEMS
5
An S-L equation (1) can be given ona finite, semi-infinite, or infinite interval
I. In the finite case, J may include neither, one, or both end points. The exclu-
sion of an endpoint a may be necessary when lim,,., p(x) = 0, lim,_., p(x) = 0,
or when any one of the functions , q, p is singular at a.
Only when/ is a closed, finite interval a < x = 6 can an S-L equation be
associated with a regular S-L system. If J is semi-infinite or infinite, or if J is finite
and p or p vanishes at one or both endpoints, or if q is discontinuous, we cannot
obtain from (1) a regular S-L system. In any such case, the given S-L equation
(1) is called singular.
We can obtain singular S-L systems from singular S-L equations by imposing
suitable homogeneous linear endpoint conditions. These conditions cannot
always be described by formulas like (2). For example, the condition that u be
bounded near a singular endpoint is a common boundary condition defining a
singular S-L system.
(9) [1 ~ xu)’ + Au =
=
0, -Il<x<]l
P,(x) are real eigenfunctions of this S-L system belonging to the eigenvalues
An n(n + 1).
=
=
2
d du
Ju =o, Q@<rxa
_—
dr | |+(
rr
d
k?y - —
The eigenfunctions of the preceding singular S-L systems are the Bessel func-
tions J,(k,r), where £,a is the jth zero of the Bessel function J,(x) of order n. It
has been shown in Ch. 2, §6, that J,(x) has infinitely many zeros; it follows that
the singular S-L system just defined has infinitely many eigenvalues.
The eigenfunctions of singular S-L systems are also orthogonal, provided that
they are square-integrable relative to the weight function p, in the following sense.
When the weight function p is identically equal to 1, we say simply that the
function fis square-integrable on the interval I.
B
(12) lim p(x[u(x)o'(x) — v(x)u/(x)] 0
x=
The conditions p(a) = p(b) = 0 and u’(x) bounded on the interval [a, 6] imply
this property, for example.
When (12) holds, we obtain from (8) the identity
(A — #) f p(x)u(x)v(x)dx = 0
for any two square-integrable eigenfunctions u and v with eigenvalues ) and y.
The integral here may be an improper integral. If \ # y, this implies, as in the
proof of the lemma of §2, that u and v are orthogonal. This proves the next
theorem.
after verifying that the boundary term vanishes. Applying it to the Bessel equa-
tion, we obtain the orthogonality relations for Bessel functions
(14)
J 0
xJn(k,x)Jn(k,x) dx = 0, kh, # k,
if J,(k,a) = Jy(k,a) = 0.
(Qk — dja,
(16) k=0,1,2,...
mee + DEF 2D)’
among whose solutions for \ = 2n are the functions ee ?H,,(x); these functions
are square-integrable and tend to zero as x > +00.
That is, the functions ¢,,(x) = et ?H,(x) are eigenfunctions for the singular
S-L system defined by (17) and by the endpoint condition that a solution y(x)
must tend to zero as x > +00, We shall now derive the orthogonality relations
for the Hermite polynomials
(18) f . H,,(x)H,(xe7™"dx =
=
0, men
2(m — n) H,,(x)H,(x)e~* dx =
=
EXERCISES C
Show that the Legendre polynomials (and their constant multiples) are the only solu-
tions of the Legendre DE that are bounded on (—1, 1).
Show that the S-L system [(x — a)(6 — x)u’!’ + Au = 0, a < b, with u(x) bounded on
a<x <5, has the eigenvalues \ = 4n(n + 1)(6 — a)*. Describe the eigenfunctions.
(a) (Laguerre polynomials). Consider the singular S-L system
(xe™*u’y’ + Ae *u = 0 on O<x< +0
with endpoint conditions that u(0*) is bounded and that e*u(x) -> 0 as
x — +00, Show that the values A = n give polynomial eigenfunctions.
(b) Show that the preceding system has no other polynomial eigenfunctions. [HINT:
Obtain the power series expansion of the general solution of the DE.]
312 CHAPTER 10 Sturm-Liouville Systems
d gat “|
d. ja x?) = 0,
+ A(1 — x*)*u a>-—l
and the condition of being bounded on (—1, 1), are A, = n(n + 2a)
5 PRUFER SUBSTITUTION
We now develop a powerful method for the study of the solutions ofa self-
adjoint second-order linear DE
(20)
d
d.
|Pe a + O(x)u : a<x<b
where P(x) > 0 is of class @’ and Q is continuous. One may want to find out
how often the solutions of (20) oscillate on the interval under consideration, that
is, the number of zeros they have for a < x < 6. This can be done by using the
Poincaré phase plane, already introduced in Ch. 2, §7. Modifying slightly the
formulas used there, we make in (20) the Priifer substitution
ris called the amplitude and @ the phase variable. When r # 0, the correspon-
dences (Pu’, u) = (r, 0) defined by (21) are analytic with nonvanishing Jacobian
For nontrivial solutions, r is always positive because, if u(x) ”(x) 0 for
a given x, by the Uniqueness Theorem of Ch. 2, §4, u would be the trivial solu-
tion u=0
We now derive an equivalent system of DEs for r(x) and 6(x). Differentiating
the relationt cot @ = Pu’/u, we get
Pu?
d@ (Puy
c? 6 = — =F = —Qlx) — 5 cot”
dx u ue
d
(22) — = Q(x) sin 294 costa = F(x,6)
P(x)
+ When @ = 0 (mod x), the relation is not defined. But the final equations (22)—(23) can still be
derived by differentiating the relation tan? = u/Pu’.
6 The Sturm Comparison Theorem 313
(23)
=
dr
~_
dx |a
1
P(x)
Q(x)| rsin
60 cos @ = —
2 |
ee
1
P(x)
Q(x)| r sin20
The system (22)-(23) is equivalent to the DE (20) in the sense that every non-
trivial solution of the system defines a unique solution of the DE by the Priifer
substitution (21), and conversely. This system is called the Priifer system associ-
ated with the self-adjoint DE (20).
The DE (22) of the Priifer system is a first-order DE in 8, x alone, not con-
taining the other dependent variable 7, and it satisfies a Lipschitz condition with
Lipschitz constant
L= sup —
The constant L is finite in any closed interval in which Q and P are continuous.
Hence, the existence and uniqueness theorems of Ch. 6 are applicable, and
show that the DE (22) has a unique solution (x) for any initial value #(a) = y,
provided P and Q are continuous at a.
With 0(x) known, 1(x) is given by (23) after a quadrature:
where K = 7(a). Each solution of the Priifer system (22)—-(23) depends on two
constants: the initial amplitude K = r(a) and the initial phase y = 6(a). Changing
the constant K just multiplies a solution u(x) by a constant factor; thus, the zeros
of any solution u of (20) can be located by studying only the DE (22).
The zeros of any solution u(x) of the DE (20) occur where the phase function
6(x) in the Priifer substitution (21) assumes the values, 0, £2, +27, . , that
is, at all points x where sin 0(x) = 0. At each of these points cos? 6 = 1 and
d6/dx is positive, by (22) [recall that P(x) > 0]. Geometrically, this means that
the curve (P(x)u’(x), u(x)) in the (Pu’, u)-plane, corresponding to a solution u of
the DE, can cross the Pu’-axis @ = nx only counterclockwise.
Now compare the DE (22) with a DE of the same form, dé/dx Fy(x, 9),
—
=
1
—
=
=
Q,(x) sin? 6 + cos? 6 = F,(x, 6)
d. P(x)
314 CHAPTER 10 Sturm-Liouville Systems
If Q,(x) = Q(x) and P,(x) = P(x) in an interval J, then F,(x, 0) = F(x, 6) there.
By the Comparison Theorem of Ch. 1, §11, we conclude that, if @,(x) is a solu-
tion of the second DE whose initial value satisfies @,(a) = 0(a), and @(x) is a solu-
tion of (22), then 6,(x) = @(x) for a <= x = b. Furthermore, we have 6,(6) = 6(0)
only if 6(x) = 0,(x), which implies that u(x) = cu;(x), whence F(x, O(x)) =
F\(x, 8,(px)). This implies that Q(x) = Q,(x) since dO/dx = 1/P,(x) > 0, where
sin @ = 0; therefore, sin @ can vanish only at isolated points. It also implies that
P(x) = P\(x), except in intervals where cos 6 = 0, and so Q(x) = Q(x) = 0
(cf. Ch. 1, §12, Corollary 1). Therefore, if sin 6(a) = 0, the number of zeros of
sin 6,(x) for a < x < b is at least the number of zeros of sin 6(x), except when
P= P, and Q = Q,, when it is equal, and when Q = Q, = 0 in an interval,
when it may be equal. This completes the proof of the following theorem.
d.
—
d. (
P(x) me + Qy(x)uy = 0
Then, between any two zeros of a nontrivial solution u(x) of the first DE, there lies at
least one zero of every real solution of the second DE, except when u(x) = cu,(x). This
implies P = P, and Q= Q,, except possibly in intervals where Q = Q, = 0.
In the case of S-L equations, since p(x) > 0, Q(x) = Q,(x) evidently implies
that A = A).
Sturm’s Separation Theorem of Ch. 2, §6 follows as a corollary, by comparing
two linearly independent solutions of the same DE.
A short and easily remembered, if somewhat imprecise, summary is this: as Q
increases and P decreases, the number of zeros of every solution increases.
Maxima and Minima. For the self-adjoint DE (20), the inequality Q(x) > 0
implies that
For, in (22), cos 6 = 0 and |sin 6| = 1, if9 = (n + 3)m. Since cos 6 = 0if and
only if u’ = 0, it follows that, if Q(x) is positive, any nontrivial solution of (20)
has exactly one maximum or minimum between successive zeros.
We now consider the variation with \ in the number of zeros of the eigen-
functions of a regular S-L system (1) to (2). Setting P(x) = p(x) and Q(x) =
Ap(x) — g(x) in (1), we obtain (20). Since u 0 if and only if sin 6 = 0 in (21),
=
=
7 Sturm Oscillation Theorem 315
the zeros of any solution of (1) are the points where 6 = 0, ta, +2z, ’
(25) _-
d.
=
Dro(x) — q(x)] sin 29 4 costa, azx=zdb
(x)
/
u(a) a
(25’) tany = O=y<a7
playu'(a) playa.’
The constants a and oe’ come from the initial condition au(a) + a’u’(a) = 0. For
fixed y, the function (x, \) is defined on the domain a S x = b, -oOO<)\A<
oo; we shall consider its behavior there.
Applying the comparison theorem of Ch. 1, $11 (and especially Corollary 1
there) to (25), we obtain the following lemmas.
LEMMA 1. For fixed x > a, 0(x, d) is a strictly increasing function of the variable
r
Proof. If x, is any point where 6(x, A) = na, then by the DE (25), we have
d6(x,, \)/dx, = 1/p(x,) > 0. Thus, the function 0 = 6(x,, d), considered as a
function of x,,, is increasing where it crosses the line 6 = nz, as shown in Figure
10.1. Hence, @(x, A) stays above this line for x > x,, q.e.d.
Lemma 2, combined with the condition 0 <= y = @(a, \) < a, makes the first
zero of u(x) in the open interval a < x < b occur where 6 = =, and the nth
zero, where 0 = nz.
Our next aim is to show that, for fixed x > a, 0(x, 4) > C© asi = OO.
In view of Lemma 2, we will have shown that lim,_... 6(x, 4) = 00 for each x,
if we can show that for every integer n > 0, we can find a number x,(A) be the
smallest x such that 6(x, A) = nz. Then, all we need to show is that x; <x such
that 6(x,; d) =
=
LEMMA 3. For a given fixed positive integer n and sufficiently large \, the func-
tion x,(A) ts defined and continuous. It 1s a decreasing function of X, and lim--o
x,(A) = a.
6 5
=
Qu 65
| 5
32/2, OOOO
OO
2S OOOO
nn
acs
a
qm
(26) pure” + APn — Inu = 0, A> _—
Pm
is the function u,(x) = sin k(x — a), where k? = (Ap, — qu)/Pm. The successive
zeros of this function are spaced at a distance 7V py/(ApPm — qu) apart. By the
Sturm Comparison Theorem (Theorem 3 above), any nontrivial solution u(x) of
the Sturm-Liouville equation (1) must have at least one zero between any two
zeros of the function w,(x). Since u,(x) has zeros on (a, 6) whendis sufficiently
large, it follows that u(x) has at least n zeros and, therefore, that 6(x, ) takes the
value na for sufficiently large , as we wanted to show.
The number x,,(A) falls between the (n — 1)st and the nth zero of u(x), and
both these zeros tend to a as \ — ©. Therefore, we have x,,(A) — @ as A > ©,
q.e.d.
forax<x=sb.
The first sentence was proved in Lemmas 1~3. The first formula of (27) was
proved in Lemma 1.
We shall now prove the second formula of (27). Choose numbers y < y, <
aw and ¢ > 0. The slope of the segment in the x@-plane joining the points (a, 7;)
and (x), €) where a < x, = b, equals (€ — y,)/(x, — a). For a point (x, 6) on this
segment, the slope of 0(x, A), as given by (25), will be less than the slope of
the segment for large negative \. Therefore, the function @(x, d) will lie below
the segment for a = x = x, for all sufficiently large negative 4. We conclude
that @(x,, A) < e for sufficiently large negative \. Since, by the argument used to
prove Lemma 2, @(x,, A) > 0, it follows that |@(x), A)| < «And since ¢ and
x, are arbitrary, the proof is complete.
We now derive an estimate for the positions of the zeros of a solution of a
regular S-L equation (1), by comparing it with equation (26) and with
where p,, and q,, are the minima of p(x) and q(x), and py the maximum of p(x)
foraxx=b.
Consider solutions of (26) and (28) for which u(a)/p(a)u’(a) = tan y. The
zeros of these solutions can be determined by inspection. They are a + (nx —
Y)/V An — 9u)/em and a + (nx — ¥)/VQpm — Inem respectively. Applying
the Sturm Comparison Theorem, we obtain the following Corollary.
COROLLARY. Let x, be the nth zero of a nontrivial solution of the S-L equation
(1). Then
Pu
(29)
dpm — Qn
The preceding results have been proved under the assumption that a # 0 in
(2). If a = 0, we can use the same argument when § # 0, by changing the
independent variable tot = a + b — x. Ifa = B = 0, we can still prove the
foregoing results with y = 7/2.
EXERCISES D
Show that the number of negative eigenvalues of a regular S-L system is always finite
and is at most 1 if g(x) > 0.
Show that any finite sequence of eigenvalues of a regular S-L system is unbounded.
Show that, at all points x where a solution u(x) of (Pu’)’ + Qu = 0 has a minimum
or a maximum, d0/dx = Q(x).
Extend the Sturm Oscillation Theorem to the case where a = 6 = 0 in (2).
10 (a) Show that /,(x) is increasing for 0 < x < |n|. [HinT: Use the identity x(x/fy =
(my — x*)J,]
(b) Prove that, if xo is the first positive zero of J, and yo is that of J), then
in| = yo < x9
*11 (a) Let u(x) be a solution of (Pu’)’ + Qu = 0, where P > 0, P’ > 0, Q> 0, and
(P’/Q)’ > 1. Show that the zeros of u, wu’, u” follow one another cyclically.
(b) Infer that the zeros of J, Jn+is Jn+2 follow one another cyclically.
13 For the modified Bessel function Jo(y) = Jo(#y), without considering its Taylor series,
show that Jj(y) > 0 and 1 < Jo{y) < cosh
y for all y > 0.
14 Show that, in the Sturm Oscillation Theorem, @(x, 4) > 00 as \ —> 00, uniformly in
any subinterval a’ = x < 6b, a’ > a.
with the DE (1). If a # 0, then the function 6(x, \) must satisfy the initial con-
dition 6(a, \) = y, where y is the smallest positive number 0 = y < a such that
pla) tan y = —a’/a. When a 0, we chose a/2. Similarly, we choose
= =
= =
Clearly, any value of for which conditions (31) are satisfied is an eigenvalue of
the given regular S-L system, and conversely. Let @(x, 4) be the solution of (25)
for the initial condition @(a, \) = . Figure 10.2 shows graphs of the function @
= 6(x, \) for various values of the parameter A. The waviness of the lines
expresses the fact that 1/P(x) in (23) is independent of A, whereas Q(x) = Ap —
q tends to infinity with A. As a result, the slope of the graph is 1/p(x) for all A
when 6 = 0 (mod =), although it tends to infinity with A for all other 6.
Since 9(6, A) is an increasing function of 4, and A(b, A) > 0 by Lemma 2 of §
7, as \ increases from —©6, there is a first value \,) for which the second of the
conditions (31) is satisfied. For this eigenvalue, we have 6(6, Ay) = 6. As A
increases, there is an infinite sequence of X,, for which the second boundary
condition is satisfied, namely those for which 6(4, A,) = 6 + na, for some non-
negative integer n. Each of these values gives an eigenfunction
of the S-L system. Furthermore, the eigenfunction belonging to A,, has exactly
n zeros in the interval a < x < b, by Theorem 4. This proves all but the last
statement of the following theorem.
A=4
5n/2F 5x /2
2xfF 2x
3x /2 5 rA=1 3x /2 A=]
re
x
7=0 Yro
THEOREM 5. Any regular S-L system has an infinite sequence of real eigenvalues
Wy <M << with lim, .. Ay = ©. The eigenfunction u,(x) belonging to
the eigenvalue X,, has exactly n zeros in the interval a < x < b and is uniquely deter-
mined up to a constant factor.
Only the last assertion wants verification. Any two solutions of (1) that satisfy
the same initial condition au(a) + o’u’(a) = 0 are linearly dependent, by the
Uniqueness Theorem of Ch. 2, §4.
EXERCISES E
1 Show that for a regular S-L system, if q(x) is increased to q,(x) > q(x), each nth eigen-
value of the new system is larger than that of the old.
Show that for a regular S-L system, if p(x) is increased to p,(x) > p(x), all positive
eigenvalues decrease and any negative eigenvalue increases.
Discuss the asymptotic behavior, as n — 00 of the nth eigenvalue of the S-L systems
defined by u” + Au = 0, and the endpoint conditions:
(a) u(0) = 0, u(r) + w(x) = 0.
(b) u(0) = 0, u(r) = u’(x).
That is, find constants ap, a; such that VA, = 2 + ay + a,/n + O(1/n9.
For regular S-L systems with two sets of endpoint conditions, (30) and
show that, if aj/a, < o’/a, the eigenvalues of the second system are smaller than the
corresponding eigenvalues of the first.
(a) Given (Pu’Y + Qu = 0, and (P,v’)’ + Qyv = 0, P,(x) > 0, Q,(x) continuous,
establish Picone’s identity.
v(x)
| a= 0
where u(a) = u(b) = 0 and v(x) ¥ 0 in [a, 8].
(b) Infer the Sturm Comparison Theorem from Picone’s identity.
we can simplify the S-L equations (1) considerably. If the functions y and h are
positive and continuous in the given interval, the first substitution leaves the
location of zeros unchanged, while the second one distorts the range of the
independent variable, preserving the order, and leaves the number of zeros of
a solution in corresponding intervals unchanged. The equivalent DE in w and ¢
is obtained from the identity d/dx h(x) d/dt, which is obtained from the sec-
=
=
ond of equations (33). When substituted into the S-L equation (1), this identity
gives
0 = hlhpQw)], + Ae — Qyw
Dividing through by the coefficient pyh? of wy, we obtain the equivalent DE (for
h, y € @),
THEOREM 6. Liouville’s substitution (34) transforms the S-L equation (1) with
coefficient functions p, p € @? and q € © into the Liouville normal form
2
ou + A — q@]lw =0
(35)
dt”
where
(36) q=
1+ oF top"
az
Evaluating the second derivative in (36) and using the identity d/dt —
=
Pp , p p / ,
3 1
--4,2
+ 1
()
p p p
(36’)
p
~
4p ( }+(
p p 4 (p 2 (I
p p 4 p
322 CHAPTER 10 Sturm-Liouville Systems
Let u(x) and u(x) be transformed into the functions f(#) and g(#) by Liouville’s
reduction (34). From the identity
2 1
dw
Jw = 0, w= xy
—s
kt
4
<=
dx? + 2
EXERCISES F
1. (a) Show that the self-adjoint form of the Hermite DE (15) is the S-L equation
[eu + reu = 0
(b) Show that the Liouville normal form of this is the S-L equation (17) for the Her-
mite functions.
10 Modified Priifer Substitution 323
Show that the Liouville normal form of the self-adjoint form of the Jacobi DE is, for
=
x =
cos t
w+| 4 @sin?
— a’) G — 6’)
+(n4 @@+6+1)
(t/2)
TT
[9 — xt IW — fab — xP]y =
=
Compute the Liouville normal form for the Legendre DE, setting x = —cos t, —a
<t<0.
the S-L equation [xe~*w’]’ + dAe~*u = 0. What is its Liouville normal form?
*7 Show that the Legendre polynomial P,(x) has-exactly zeros. [HinT: Reduce the
Legendre DE to Liouville normal form and apply Ex. E6.]
*8 If x, = cost), »*X, = cost, are the zeros of P,(x), x, < x,41, show j that 2x(—1)/
(Qn + 1) <4, < Qaj/(2n + J), for 2 = 7 <n. [Hint: Use the Liouville normal form
and Ex. E6.]
(39) u” + [A — gq(x)]u
—
=
u” + Q(x)u = 0, Q(x) = A — q(x)
The constants a, @’, B, 6’ are usually changed, but we still have a” + a’? # 0
and 6? + 6” # 0. By Theorem 6, Corollary 1, the eigenvalues of this system are
the same as those of the original system, and the eigenfunctions are obtained
from those for the Liouville normal form through the Liouville substitution. To
study the distribution of eigenvalues and magnitude of the eigenfunctions, it,
324 CHAPTER 10 Sturm-Liouville Systems
therefore, suffices to treat the system (39)—(40). In §§10—-11, we shall use mainly
(40)
We shall assume from now on that Q(x) > 0 for a = x = 8, that is, that \ >
q(x) and Q € @'. We introduce the functions R(x, \) and (x, A), the modified
amplitude and modified phase, which are defined in the terms of a given solution
u(x, d) of (39) by the equations
These equations constitute the modified Priifer system for the DE (39)
We shall now derive a pair of DEs for R and ¢ that are equivalent to (39). We
havet
VQu +
(42) cot¢ =
Ye
—_—— —
Yo
u’
Qu? + u” 1 Qu
(csc* $)¢’ = Q'? 2 9 OF? u
Q’ JQ
2RR’ = 2Q7/?(Quu!+ u’u”) + & 2 Q-V2y%
2Q
Q’ —Q’
(44)
F_Yg (sin* @ — cos* ¢) cos 2¢
R 4Q 4Q
+ When u # 0, these equations are valid. When u = 0, set tang = V Qu/w’ and proceed similarly.
10 Modified Prifer Substitution 325
,
q
(45a) = VA sin 26
~ 1 4a= 9)
q’
(45b) Re cos 2¢
R 40-9
Clearly, to every nontrivial solution of (39) there corresponds a solution of
the modified Priifer system, and conversely. Furthermore, we know that R > 0,
unless R vanishes identically.
Equations (45a) and (45b) determine the asymptotic behavior of the solutions
of (39) as \ — 00, The fundamental result is the following.
THEOREM 7. Let $(x, ) and R(x, X) be solutions of the system (45a) and (45b),
where q(x) € @' is bounded. Then, as \ > 00,
O(1)
(46) d(x, A) = o(a, ) + VA(x — a) + —_——
Vi
and
ov)
(47) R(x, A+) = Ra, ») + 7
Intuitively, Theorem 7 states that for large \ the modified phase ¢ is approx-
imately a linear function of VA, and the modified amplitude function R is
approximately constant.
The Symbol O(1). The symbol O(1) used here and later signifies a function
J(x, ») of x and X, defined for all sufficiently large A, which is uniformly bounded
for a = x = bash — ©. Hence, O(1)/M signifies a function f(x, \) such that
A‘f(x, A) is uniformly bounded. The symbol O(1)/X’ is also often written O(\™*),
as has been done in analogous contexts in Chs. 7 and 8.
The formula fx, 4) = O(1), where fis a given function, is not an ordinary
equation. Thus, to write O(1) = f(x, A) would be meaningless, since O(1) is not
a function. The formula means simply that f remains uniformly bounded for all
x as AX — 00, and that no other property of the function f is needed for the
purpose at hand. Using this definition, the following important properties of
the symbol O(1) can be easily verified:
for any finite a, b. Again, if a and @ are real numbers with a = 8, then O(1)/A*
+ o(/r® = O(1)/A*. Finally, if q(x) is any bounded function of x, then by
326 CHAPTER 10 Sturm-Liouville Systems
[A — g(x)]* = A*
1—q(x)
r
| = d* —ag(x)A*! + O(I)AT?
The preceding formulas will be used freely in subsequent computations
Proof. For all X for which |q(x)| <A on [a, 4], we have as before
q
—.
q
A
(1+ O(1)
r
-£, 00
n2
r
Aq
"
V i-~q=vx(1 -4) =
=
Vx
O(1)
x3?
We now compare the solutions of the DEs (45a) and (45b) with the solutions
x(x, ) = o(a,) + VA(x — a) and R(x, ) = R,(a) of
Vx
d’ —
=
and (log RY = 0
using Theorem 3 of Ch. 6. In making this comparison, we set € = O(1)/ Vy, and
replace x and y with the functions $(x, A) and (x, A), respectively. If $,(a, )
= (a, d), the inequality (7) of Ch. 6 gives |o(x, A) — $)(x, A)| = O(D/ Vx, and
since #,(x, A) = (a, A) + VA(x — a), equation (46) follows.
Similary, to derive (47), compare R(x, A) with R,(x, d), using the identity
eOOWR me y+ O(1)/\ obtained from Taylor’s formula.
We shall now use the modified Priifer substitution to study the asymptotic
behavior of solutions of the Bessel DE (3) as x — 00. The substitution u = w/
Vx reduces (3) to the Liouville normal form
2 1
(48) w” + [1 — (M/x*)]w = 0, <
0O<xom, M =n*—F
whose solutions are w(x) = VxZ,(x), where Z,(x) is a solution of the Bessel DE
[see (38)]. The modified Priifer system for (48) is then obtained by setting Q(x)
= 1 — M/x* in (43) and (44). This gives
x?
M sin 26
2(x® — Mx)
R(x) _ ~Mcos 26
(49b)
R(x) 2(x? — Mx)
11 The Asymptotic Behavior of Bessel Functions 327
=]--
1M , O(1) R(x) o(1)
’(x) 2 x2 ~~ 3 ’ 3
x R(x) x
Here O(1) denotes a function of x that remains bounded as x ~> 00. Integrating
the first of these equations between any x > VM and y > x, we obtain
M M. O(1)
(x) o(y)
—_—
a
2
2x 2y x
where R, Ry)
It follows that every solution of the Bessel DE (48) has the asymptotic form
na) = x? EB ()
|sn(6 2
+
O(1)
x
2
Since sin (A + O(1)/x*?) = sin A + O(1)/x?, the preceding display can be rewrit-
ten as
1psin( M
nl) «+ bo — =
2
)+ O(1) B/2
The solution Z,, is uniquely determined by the constants R,, and ¢.. above. For
if two solutions had the same asymptotic amplitude R.. and phase ¢.., their dif-
ference would be a solution having modified amplitude R(x) O(1)/x°. Since
R(x)
= Raoexp oo)
this would imply R = u = 0. Setting x. = 7/2 + d., this proves the following
theorem.
(n? — 1/4
(50) Z,(x) = —= cos
x
(++ x- 2
O(1)
xo?
For the Bessel function J,(x), it can be shown that x. = n1/2 + 7/4 and that
Ro = V2/". The Neumann function Y,(x) is defined likewise by the conditions
Xo = nw/2 + 37/4 and R, = 2/x. Thus, the Neumann function Y,,(x) is
defined by the condition that it has the same asymptotic amplitude as /,(x), with
an asymptotic phase lag of 2/2 radians.
That is, the asymptotic relation between J,(x) and Y,,(x) is, for large positive
x, the same as that between cos x and sin x. The Hankel function H,(x) =
Jnlx) + 7Y,(x) is, therefore, analogous to the complex exponential function
e* = cos x + isin x.
12 DISTRIBUTION OF EIGENVALUES
Weshall next show that the asymptotic distribution of the eigenvalues of all
regular S-L systems is the same: the trigonometric DE u” + Au = 0 is typical.
We shall treat in detail the case of separated endpoint conditions (2), also assum-
ing a’6’ # 0 for uniformity. We can assume the given S-L system reduced to
Liouville normal form (39)—(40), because this does not change the eigenvalues
or the condition a’f’ # 0.
For the trigonometric DE and the boundary conditions u(a) = u(b) = 0, the
nth eigenfunction is sin [ma(x— a)/(b — a)] and the nth eigenvalue is »,, =
2a?/(b — a)*, n = 1, 2, 3,.... For u(a) = w/(b) = 0, u,(x) sin VA,(x — a),
where A, = (n_+ 3)*x?/(b — a)®. For u’(a) = u’(b) = 0, the (n + 1)st eigenfunc-
tion is cos A(x — a), where A, = n?x?/(b — a)? andn = 0,1,2,....
We will treat in detail, here and in §13, regular S-L systems satisfying sepa-
rated endpoint conditions (2) with a’f’ # 0. We will show that VA, = [nz/(b —
a)] + O(1)/n in this case, n = 0, 1, 2,.... That is, unless o& = 0 or #’ = 0 in
(40), the asymptotic behavior of the eigenvalues and eigenfunctions is similar to
that of u” + Au = 0, with the endpoint conditions a = 6 = 0.
THEOREM 9. For the regular S-L system (39)—(40), let a’B’ # 0. Then the eigen-
values \, are given, as n — 00, by the asymptotic formula
nT o(1)
(51) Vin b-a
+
Here O(1) denotes a function of n that is uniformly bounded for all integers
n = 0.
o()
(54) a) =F +t n3/2
+
O(1)
(54’) (6, r,) = 9 +
Vin
Subtracting (54) from (54), and comparing with (46) of Theorem 7, we
obtain the equation
O() _ O(1)
(55) (6, d,) = O(a, A,) = ne + Vib— a) +
Vin Vin
Letting \,, > 00 we obtain lim,_... nA, 2 = (6 — a), or VA, = K,n, where the
K,, tend to 1/(b — a). Substituting into (55), we obtain
J u?(x)p(x) dx = 1.
a
In the ease of the eigenfunctions of (39), p(x) = 1. Our aim is to show that
the normalized eigenfunctions of (39) and (40) behave approximately like cosine
functions, provided that a’6’ # 0. [The cases a = 0 and 6 = 0 are similar, after
phase-shifts of /2 in (54) and (54’.]
330 CHAPTER 10 Sturm-Liouville Systems
n(x — a) O(1)
(56)
b—-a n
The proof of this theorem will be carried out in three steps. For an eigen-
function u,(x), with eigenvalue A,, we have by (41)
R(x, d,)
(57) u,(x) = = sin (x, d,), a=x=b
An — q(x)
Proof. Using (x, d) as the variable of integration in (55), and recalling from
(46) that dx/dp = (dp/dx)"' = —d~? + O(1)A~*”, we have
(bd) dx
f sin?o(x, A) dx = f ¢ — dd
sin?
(@,)) dp
(6,2) $(6,r)
@ _ sin 2¢ _ Ne -
J ¢ (a,A)
sin? ¢ dd =
2 4 | @r) 2
+ O(1)
Substituting into the previous displayed formula and simplifying, we obtain (58)
2
(1+ O(1)
uN?
)+ O(1) no
13 Normalized Eigenfunctions 331
vr )+ O(1)
R(a, d) O(1)
vA ve nef?
R(a, »)
1-
O(1)
n° 4 A
VF *) O(1)
rn
Solving for R, and taking the asymptotic form of the quotient, we get (60), q.e.d
LEMMA 3. Let X,, be the nth eigenvalue (Ag < dy < dg of the S-L system
(39)-~(40). Then, as n — 00, unless a’B = 0
nx(x 4
(61) sin $(x, A,) = cos + O(1)Az¥?
b-
O(1)
(x, 4.) = oa, A) + VAx
— a) + —
Ve
332 CHAPTER 10 Sturm-Liouville Systems
Moreover by (54), (a, A,) = 7/2 + O(1)/ Vs Substituting back into the pre-
ceding formula, we get
(62)
sin $(x, A,)
=
=
We now apply Theorem 9 to this formula. By formula (51) and the mean value
theorem, we have
ug()= 2s n= 2) + Or,”
Since \,/? = O(1)n“', this gives Theorem 10.
EXERCISES G
A
P,(cos 6) —_
= mip
(sin 6)!
COs
( nts 2
9-2
4
+On"*) for O0<0<a9
Show that the relative maxima of x!/*|J,(x)| form an increasing sequence if 0 <n
< $and a decreasing sequence if n > }.
*4 (Sonin-Polya Theorem). Show that if in (Pu’)’ + Qu = 0, P, Q€ @'[a, bI, Q(x) # 0,
and P(x)Q(x) are nondecreasing, the successive maxima of |w(x)| form a nonincreas-
ing sequence, and that equality occurs if and only if Q(x) = 1/P(x). [Hint: Show that
the derivative of o(x) = u(x)? + P(x)u’(x)?/Q(x) is nonpositive.]
*5 Show that the values of | PX) Q(x) [77 | u(x) | at those points where u(x) = 0 are a
monotonic increasing or decreasing sequence, according as the values of P(x)Q(x)
are decreasing or increasing. [HinT: Consider v(x) = P(x)Q(x)@(x), ¢ as in Ex. 4.]
14 Inhomogeneous Equations 333
14 INHOMOGENEOUS EQUATIONS
Let U(x) be the solution of L[u] = 0 satisfying the initial conditions U(a) =
a’, U'(a) = —a; let V(x) be the solution of L[u] = 0 satisfying Vid) = 6’, V(b)
=
=
—; let F(x) be the solution of L[u] = f(x) satisfying F(a) = F(a) = 0. The
existence and uniqueness of these functions follow from Theorem 7, Corollary
2, of Ch. 6, §8. For any constants, c, d, the function
THEOREM 11. Either DE (63) has a solution w satisfying the boundary conditions
A [w] =a, and B [w] =8), for any given constants a, and B,, or else the homogeneous
DE L[u] = 0 has an eigenfunction with eigenvalue 0, satisfying the homogeneous
conditions A[u] = 0 and Biu] = 0.
334 CHAPTER 10 Sturm-Liouville Systems
15 GREEN’S FUNCTIONS
We now show that, in the first case of the preceding theorem, there exists a
Green’s function G(x, &) defined for a = x, § = b, such that the solution of (63)
subject to the boundary conditions (30) is given by
where the factor e(£) above is chosen to give 0G/0x a jump of 1/p9(€) across
x = & Thus
U(x) VE)
sxx
pol&) WEE)’
(67) G(x, &)
UG) Vix)
Exx=b
po(é) W(E)’
THEOREM 12. Unless W = 0, equations (66) and (67) yield for any continuous
function f on [a, b] a solution u(x) of the DE L{u] =f(x) that satisfies the boundary
conditions A{u] =B[u] = 0.
15 Green’s Functions 335
That is, unless the homogeneous linear boundary-valué problem L[u] A[u)
= Blu] = 0 admits an eigenfunction, the function defined by (67) is a Green’s
function for the system L[w] f, A{u] = Blu] = 0
The proof is like that given in Ch. 2, §11. Rewriting (66) as
oy= | Ge of ae + | Oe, of at
The endpoint contributions give G(x, x" )f(x") — G(x, x*)f(x*) = 0; they cancel
since G(x, £) and f are continuous for x = & Differentiating again, we have, by
Leibniz’ rule
(x) f Gyle, Of
EO) d& + G(x, x) fx
+ | Gale, B® a8 — Gx, x*Yfle")
The two terms corresponding to the contributions from the endpoints come
from the sides x > £ and x < é of the diagonal; since f is continuous, their
difference is [G,(x*, x) x(x~, x)]f(x) S(*)/po(x). Simplifying, we obtain
Sf)
u(x)= f Cale,of
e at + 1%.
polx)
EXERCISES H
x for xe
|
a”
1 u
=
=
g for x>€&
u”
2 u(—1) = u(1) = 0; G(x, &) Cle — €| + x€ — 1172
=
f
=
u = =
0;
for xsé
G(x, = | log &
log x for x>&
u” — u = f, u(x) bounded as |x| —> 00; G(x, &) = —exp (|x — &|)]/2.
— e%,
G(x,) = |e*(1
(ef — 1),
xe
x ee
Find Green’s function for u” — u = f with u(—a) = u(@) = 0. Show that, as a >
00, it approaches that of Ex. 4.
*8 Show that the Green’s function G of a regular S-L system is a symmetric function of
x and &, in the sense that G(x, £) = G(&, x).
2m
(68)
v + —
2 ) [E— Vix)ly = 0
Physically, the function V(x) has the significance of potential energy; the con-
stant m stands for the mass of the particle; the constant E is an energy param-
eter; h = h/2z is a universal constant, whose numerical value depends on the
units used. The “wave function” (x) may be real or complex; yyy* dx = |y|?
dx is the probability that the particle under consideration will be “‘observed” in
the interval (x, x + dx). The eigenvalues of (68) for varying E are the energy levels
of the associated physical system.
The DE (68) is precisely the Liouville normal form
(69) u” + [A — q(x)]u 0
=
=
this interval, the ‘“‘endpoint’”’ condition that a solution remain bounded as x >
+00 defines a singular S-L system (cf. §4). In problems involving the Schroedin-
ger equation, it is customary among physicists to define the spectrum of this S-L
system as the set of all eigenvalues for which eigenfunctions exist. The set of
isolated points (if any) in this spectrum is called the discrete spectrum; the part (if
any) that consists of entire intervals is called the continuous spectrum. We shall
adopt this suggestive terminology here; unfortunately, its logical extension to
boundary value problems generally is very technical, even for ordinary DEs.t+
For regular S-L systems, we have proved that the spectrum is always discrete,
and the eigenfunctions are (trivially) square-integrable. We now describe a sim-
ple singular S-L system whose spectrum is continuous and whose eigenfunctions
are not square-integrable.
For every positive number \ > 0, this DE has two linearly independent bounded
solutions sin (VAx) and cos (VAx). For \ = 0, it has the bounded solution u =
1, and no other linearly independent eigenfunction. For \ < 0, it has the lin-
early independent unbounded solutions sinh (Vx x) and cosh (Vx x), and no
nontrivial bounded solution. Hence, the spectrum of the free particle is continu-
ous: it consists of the half-line \ = 0.
Comparing with Example 7 of §4, we see that this has the eigenfunctions
e” 7H,(x), for\ = 2n + 1 (n =>
=
that for the Taylor series era y 6'x”’/ (r!), namely ¢y,49 = Beo,/(y + 1). Setting
2r = k, we see that, for all sufficiently large k,
r+ >
Cn+9
A
>0 if 6<1
a Cr
Hence |Ay(x)| > Be? — pr(x), where B > 0, pr(x) is a polynomial and (say)
6 = 4. It follows that |e~*Ay(x)| > Be*/* — O(1) is unbounded, unless d is an
odd positive integer. Finally, if u(x) is any bounded nontrivial solution of (71),
the same is true of u(—x) and of [u(x) + u(—x)]/2, (u(x) — u(—x)]/2. This
shows that, if (71) has an eigenfunction, it must have an odd eigenfunction or
an even eigenfunction. Since either of these would be defined up to a constant
factor by the relation
(2k —dX¥+ Da
Me hE DR + QD)
on its coefficients, we see that the Hermitefunctions e~* 7H,(x) are the only eigen-
functions of the harmonic oscillator.
(72) wu
=| 0—C*u on
on
|x| >a
|x| <a
+ This is done with the usual understanding (Ch. 2, Ex. Al2) that a “solution” of (72) is a function
u € @' which satisfies (72), and so is of class @? where q(x) is continuous.
18 Mixed Spectrum 339
bounded for x > a are the functions A exp (— V —Ax) which satisfy u’(a)/u(a)
=
—). Writing p d\ + C?, we see that the even solutions A cos ux of
(72) satisfy the same boundary condition w’(a)/u(a) V—A if and only if
“tan pa = —X. The odd solutions B sin px satisfy it if and only if » cot pa
= —
Proof. The first statement follows from the Sturm Comparison Theorem
comparing with the DE u” + Au/2 = 0.
To prove the second statement, first change the independent variable
tot = Vax, giving the DE u, + Q()u = 0, with Q@) = 1 — git/Vd)/A. Applied
to the new DE, the Priifer substitution (21) gives, by (22),
ae, —B
(73) — = Q(f)sin?6 + cos?
@= 1 +=sin294— A
Vx
ft
f ro as= f a
+ o( )| sin20dé
The first term on the right side above can be integrated by parts
The boundedness of the first two terms on the right side of this equation is
evident; the last term is bounded because d@/ds
=
=
1 + O(1)/s. Hence, u? is
bounded because u? < 7° < K? exp {J F(s) ds}.
Combining Lemma 1 with the analogous result for negative x, we obtain the
following result.
THEOREM 13. If q € @ satisfies q(x) =A/x + O(1/x°) as x > +00 and q(x)
= B/x + O(1/x?) as x > —00, then the spectrum (69) includes the half line > 0.
As regards the discrete portion of the spectrum, the key result is the following
lemma, which characterizes the asymptotic behavior for large x of a wide class
of DEs that have nonoscillatory solutions, such as the modified Bessel equation
of Ch. 9, §7.
LEMMA 2. In the Schroedinger equation (69), let q be continuous, let lim,.4.. q(x)
= 0, and let) = —k? < 0. For any ¢, 0 <€ <k, there exist two solutions u(x) and
U(x) of (69) such that, for all sufficiently large x,
it is clear that F(x, 7) <= G(x, rT) = H(x, r) on the domain 7 = k — ¢ > 0. More-
over, the solutions p(x) = k — ¢ and o(x) = k + « of the displayed DEs satisfy
p(a) < r(a) < o(a). Hence, by the Comparison Theorem of Ch. 1, §11, we have
k —€ = p(x) S 7(x) S o(x) = k + «. Integrating, we get the first inequality of
(74).
We now derive the second inequality. As in Ch. 2, §5, a linearly independent
solution of the DE (69) is given by
ds
Ug(x) = Zkuy(x)f
x uj(s)
The first inequality of (74), applied to the integral on the right, gives the
inequalities
OD
1 —2(k—€)x >
ds 1 —QWhk+ex
=>
2 —2ex
é
—(k+Ox
(75) eOO > u(x) = é
1 — &/k 1+ ¢/k
But, for any 9 such that 0 < 3e < 9 < k we have, for sufficiently large x,
=-
- —
Applying these inequalities to (75), we obtain the second formula of (74) with 7
in place of ¢. Since, for any 7 with 0 < » < k, we can find ¢ 7/6 with 0 < 3¢
=
=
COROLLARY I. On (0, 90), let q(x) be continuous and satisfy lim,... q(x) =o
Then every solution of the Schroedinger equation with X <q that is bounded on the
interval (0, 00) is square-integrable.
COROLLARY 2. Lei q(x) € @ on the line (—00, 90), and let q(x) tend to limits qo
and q, respectively, as x — +00, Then every eigenfunction with eigenvalue \ < min
(40.9) is square-integrable.
THEOREM 14. Let q(x) be as in Theorem 13. Then, for \ > 0, the spectrum is
continuous. For \ < 0, the eigenfunctions are square-integrable.
It can also be shown that, for \ > 0, the eigenfunctions are not square-inte-
grable, and that for \ < 0, the spectrum is discrete.
EXERCISES I
2 Show that, if uw” — q(x)u = 0,0 = x < ©, with g(x) bounded, the DE cannot have
two square-integrable linearly independent solutions. [Hint: Use the Wronskian.]
Inu” + [A — g(x)]u = 0, 0 < x < 00, if g(x) > +00 as x — 00, show that, for any
A, the DE has exactly one square-integrable solution up to a constant factor.
#4, Under the assumptions of Ex. 3, show that the S-L system corresponding to the
boundary condition u(0) = 0, u(x) square-integrable in [0, 00), has an infinite
sequence of eigenvalues.
Show that, if the DE uw” + q(x)u = 0,0 = x < 0, g€ @ has a solution u(x) with
lim,-.o U;(x) = 1, it also has a solution w(x) such that lim... w9(x)/x = 1.
has a solution with lim,..,. u(x) = 1. {Hint: Show, by successive approximations, that
the integral equation u(x) = 1 — f2(¢ — x)q(i)u() dt has a solution.]
Suppose that all solutions of the DE u” + q(x)u = 0 are bounded as x -> 00 and that
Jop(x) dx < 00, p(x) > 0. Show that, for all A, all solutions of the DE u” + (g(x)
+ Ap(x))u = 0 are also bounded as x — 00, (Hint: Consider the inhomogeneous DE
u” + qu = —Apu, and show that the integral equation obtained by variation of para-
meters has a bounded solution.]
Show that, if k® > 0 and J3'|q(x) — k?| dx < 00, all solutions of the DE u” + q(x)u
= 0 are bounded as x > ©.
wt2v+| r 1 a
-—
(
=+
4
—
2
Js
are &
=
=
67/28
V2TM(x), where L®(x) = d*[L,(x)]/dx", for a = (k® — 1)/4 and
A =n — (k — 1)/2, n, k any nonnegative integers.
ADDITIONAL EXERCISES
Using Rodrigues’ formula, show that between any two zeros of P there is exactly
(a,b)
one zero of n#l> >
if a,b —1.
lim [ — x?)u’(x)] = 0
xt]
Show that there exists a bounded differentiable function g on a < x < 3, satisfying
the inequality g’ + g?/P(x) + Q(x) < 0, if and only if no solution of (Pu’’ + Qu
= 0 has more than one zero ona = x = 5.
(1/x) — 4, O<xx<}
g(x) -f Q(t)dt + 1/(x — 1) g$sx<1
18 Mixed Spectrum 343
then, between any two zeros of a solution of u” + 2p,u’ + q,u = 0, there is at least
one zero of u” + 2pu’ + qu = 0. [HinT: See Ch. 2, Ex. B4.]
9. For a regular S-L system with aa’ < 0, and 86’ < 0, and A less than the smallest
eigenvalue, show that the Green’s function is negative.
CHAPTER 11
EXPANSIONS IN
EIGENFUNCTIONS
1 FOURIER SERIES
Note that the nonzero terms a, cos x, 5, sin x, .. in (2) are actually them-
selves eigenfunctions of the periodic Sturm-Liouville system in question (Exam-
ple 3 of Ch. 10); hence f(x) is represented as a sum of eigenfunctions in (2).
However, we shall adopt the usual convention of referring to the normalized
cos kx and sin kx in (2) as the eigenfunctions of the system.
Though there exist continuous functions whose Fourier series are not con-
vergent, the following sharpened form of Fourier’s Convergence Theorem
applies to all continuous periodic functions.
+ Fourier’s theorem is proved in Courant and John, p. 594 ff; Fejér’s Theorem is proved in Widder,
p. 423.
344
1 Fourier Series 345
N-1
6)(x) = =
N | » |Oy
2
2 (a,coskx + b,sinis
n=0 k
= 4%
2
+ y (al’coskx + Bysinkx)
The preceding results, which we will assume as known, yield as corollaries the
following statements about cosine series and about sine series. Let f(x) be con-
tinuous on 0 = x < 7; define a function g(x) for —a7 <= x < =x by the equation
g(x) = f(|x|). Since g(—x) = g(x), g(x) can be extended to an even periodic
function of period 27, which is defined and continuous for all real x. By sym-
metry, all coefficients 6, are zero in the Fourier series of g(x). Applying Fejér’s
and Fourier’s Convergence Theorems, we have the following corollary.
2 ORTHOGONAL EXPANSIONS
oo
(4) f(*)
=
=
d= cabal)
h=1
Multiplying both sides of (4) by ¢,(x)p(x), and integrating term-by-term over the
interval—as is possible for uniformly convergent series—we get from the
orthogonality relations (3) the equation
iT x
“ane, f cos”kx dx = f
_
vr =
we
sin? kx dx = 4
The preceding conclusion holds provided that one can integrate the series
Xcyb,(x)@,(x)
p(x) term-by-term on the interval /. This holds much more generally
than for uniform convergence, e.g., for mean-square convergence as defined in
§3.
3 Mean-Square Approximation 347
The preceding conclusion was justified by using the fact that uniformly con-
vergent series can be integrated term-by-term on any finite interval J. Many
other series of orthogonal functions also can be integrated term-by-term, and
formula (5), therefore, also holds for them, as we shall prove in later sections.
3 MEAN-SQUARE APPROXIMATION
0
=
=
(8/) E= f E~ Tina | dx
= J fea: + Dat + on — 09 f de ax
The right side shows that the minimum is attained if and only if +, =
=
c;. This
proves the following result, for any interval J.
The coefficients c, are called the Fourier coefficients of f relative to the orthog-
onal sequence ¢,.
The partial sum ¢,¢9(x) + + - - + ¢,,(x) in Theorem 1 is thus, for each n,
the best mean-square approximation to f(x) among all possible sums y,¢,(x) + - - -
+ ¥,6,(x); it is often called the least square approximation to f(x) because it min-
imizes the mean square difference (7). The remarkable feature of least-square
approximation by orthogonal functions is that the kth coefficient ‘y, in the list
(y) > » ¥) which gives the best mean-square approximation to f is the same for
all n = k. This “finality property’? does not hold, for example, in the case of
least-squares approximation by nonorthogonal functions, or of the approxima-
tions in Fejér’s Theorem, or of best uniform approximation minimizing the func-
tional sup, <,<4| f(x) — Dh=1 €,6,(%)
|.
1
- coskx, Va sinkx
Von’
3 Mean-Square Approximation 349
(9) a f F?@)e(x) dx
For the right member of (9) to be finite, it is necessary that f° be integrable
that is, that f be square-integrable with respect to the weight function p. When
this is the case, the integrals (8) are also well-defined by the Schwarz inequality
Under these circumstances, since the right side of (9) is independent of n, if we
let n tend to infinity, we will still have
That is, the Fourier coefficients of any square-integrable function f form a square-sum-
mable sequence of numbers if the $, are orthonormal
EXERCISES A
1
Show that a, cos kx + 0, sin kx = (1/m) [® 2 f(O cos [A(t — x)] dt.
2 Show that4 + Lf_, cos kx = sin [(2n + 1)x/2]/(2 sin (x/2)]
3 Using Ex. 2, infer that
(a) Prove in detail Corollaries 1 and 2 of Fejér’s Theorem, discussing with care the
differentiability at 0 and x of the periodic functions constructed
(b) Find necessary and sufficient conditions for a continuous function on [0, 7] to
be uniformly approximable by a linear combination of functions sin kx.
» ear |
sin (nt/2)
le)= a [fle+
n (t/2)
*6. Prove Fejér’s theorem, assuming that Ex. 5 holds
For the regular S-L systems in Exs. 7 and 8, (a) find the eigenvalues and eigenfunctions
(b) obtain an expansion formula for a function f€ @! into aseries of eigenfunctions
7. u” +ru = 0,u (0) = 0,w(r)
= 0,0SeS0
350 CHAPTER 11 Expansions in Eigenfunctions
—W asx
=z Ta
4 COMPLETENESS
lim
n~e
1| 0 _ > va] 069dx= 0
it is necessary and sufficient that
lim
n7c
{| freae ~ dat foto ax| + > - aa?| op ax| =
=
Since the term in square brackets is nonnegative by the Bessel inequality (10),
and since J ¢,2p dx > 0 for any nontrivial ¢,, the limit is zero if and only if y, =
¢, for all k, and equality holds in the Bessel inequality (10). This proves the fol-
lowing results.
f f?(x)p(x) dx —
=
+ Here and below, the equation f = L7? c,¢, is to be interpreted in the sense of mean-square con
vergence, namely, that the partial sums 2j-, ¢ converge in the mean square to the function f with
respect to p.
t If every continuous function can be expanded into a series LY° ¢¢,, then many discontinuous func-
tions also have such an expansion, convergent in the mean square. The class of all such functions is
that of all Lebesgue square-integrable functions (see §11). We are here considering only continuous
functions in order to avoid assuming a knowledge of the Lebesgue integral.
4 Completeness 351
ce
For example, take the case of Fourier series. In the notation of (1), the con-
dition for the completeness of the functions 1, cos kx, sin kx on —7 Sx Sr
is that, for all continuous functions /,
N-1
(14) lim +
N-~co |
a9
—
2
+o (1 - Ay (aj + | = fre dx
k=1
(14’)
“| a% + > (a, + by)|= [Ura <o
2 k=1
Since [1 — (k/N))? <= 1, it follows that, if we replace the sum in square brackets
on the left side of (14) byay?/2 + ON, (a? + 6,2), we will get an increasing
sequence whose limit is at least equal to ff? dx. But, by (14’), this limit is at most
equal to f f? dx. Hence, the limit is exactly { f? dx, and (12) is proved. Since
any continuous function on —7 & x = 7 can be given an arbitrarily close mean-
square approximation by a continuous function satisfying f(—a) = f(x), this
proves
352 CHAPTER 11 Expansions in Eigenfunctions
S- Yah) pa <e
1f we replace each of the y, by the Fourier coefficients ¢, of f relative to ¢,—as
given by formula (5)—then by Theorem 2 the square integral on the left
decreases:
k=1
5 ORTHOGONAL POLYNOMIALS
J Pylx)P,(x)p(x) dx 0, men
=
(15)
=
a
5 Orthogonal Polynomials 353
Proof. Let p(x) be any polynomial of degree n. We can find ¢, such that p(x)
— ¢,P,,(x) is a polynomial of degree n — 1 or less. Hence, by induction on n, we
can express f(x) as a finite linear combination of Po(x), . . . , P,(x). By the Weier-
strass Approximation Theorem, we can approximate uniformly any continuous
function arbitrarily closely by a suitable polynomial p(x). By the preceding
lemma, every continuous function can, therefore, be approximated arbitrarily
closely in the mean square by a linear combination of the P,. The result now
follows from Theorem 4.
EXERCISES B
2 Show that, if f, > f in the mean square, and ¢,, cf are the Fourier coefficients of f,
f, relative to a given orthonormal sequence ¢,, then cf” — ¢, uniformly in &.
Using expansions into Legendre polynomials, obtain a formula for the best mean-
square approximation in |x| << 1 of a square-integrable function by polynomials of
degree = n.
*4 Using the Liouville substitution, obtain from Fourier’s theorem an expansion theo-
rem for functions f€ @°[—1, 1] into series of Chebyshev polynomials.
Show that, given square-integrable functions f;, » > fu a Sequence ¢,, > %, of
orthonormal square-integrable functions can be found for which f, is a linear com-
bination of ¢), »o lsik=sm.
Trying v = Df. a,x", the Method of Undetermined Coefficients (Ch. 4, §2) gives
the recurrence relation 4,,, (k — a)a,/(k + 1)?. Hence, we have
=
=
ak + 1)a
(19) a, = (—l)ala — l(a — 2)--:
(ety?
Thus, the functions L,(x)e~*/? are eigenfunctions of a singular S-L system. These
functions are certainly square-integrable, together with their derivatives, hence
Theorem 2 of Ch. 10 applies, giving the orthogonality relations
f e"L,,(x)L,(x) dx = 0, men
Proof. Suppose that P,(x) has fewer than n zeros in (a, 6). Let x, » Xm (m
<n) be those zeros at which P,(x) changes sign. Then the polynomial (x — x,)(x
— X9)+ + + (x — %,)P,(x) would be of constant sign. Hence
J (x ~ x) + + + & — X_)P,(x)p(x) dx # 0,
a
m<on
+ The normalizing condition a, = 1 is also often used, and makes some formulas simpler.
356 CHAPTER 11 Expansions in Eigenfunctions
0
a
Proof. First choose A,, such thatP,,, ,(x) — xA,P,(x) is a polynomial of degree
n or less, so that
The numerical values of the constants A,, B,, C,, in Theorem 7 depend on the
normalizing factors used to define the orthogonal polynomials considered. For
convenience, we have listed in Table 1 the recursion coefficients for some com-
mon polynomials.
EXERCISES C
Establish the following formulas for Hermite polynomials (see Ch. 4, §2):
Polynomial A, B, C,
2n+1 —n
Legendre 0
n+1 n+1
Chebyshev 2 —1
2n +X l—n-— 2A
Gegenbauer
n+1 n+1
Hermite 2 —2n
—l1 2n+1 —n
Laguerre
n+1 n+1 n+1
Infer from Ex. 5 that ¢,(x) = (n!) e-*L,(x), where L, is the nth Laguerre
polynomial.
Show that, if L,(x) is the Laguerre polynomial of degree n, then d'[L,,(x)]/dx* satis-
fies the DE
xv” + (k+1—
x)’ + (n—- kv = 0
coincides with f(x) in at least n + 1 points of the interval. [HinT: Use a method
similar to the proof of Theorem 6.]
*10 Show that, in Theorem 6, between any two zeros of P,, there is exactly one zero of
Pit
*1] Show that the Legendre polynomials are the only Gegenbauer polynomials for
which the maximum of | P3(x)|, namely P£(1), is independent of n.
*12 (a) Expand the function (1 — 2xh + h9)~'/%(|x} < 1) into a series of Legendre
polynomials, and show that the nth Fourier coefficient is 2h"/(2n + 1).
358 CHAPTER 11 Expansions in Eigenfunctions
*13. Show that the generating function for the Laguerre polynomials is,
eo
_ exp | 1-t
To i—xt
= >" #L,(x)/n!
n=O
In Exs. 14-17, D = d/dx and p(x) is positive. The method of proof is to find-by induction
a S-L equation satisfied by the expressions given.
*16. Show that the only system of orthogonal polynomials of the form
K,(0(x))~'D"[(x)(ax? + bx + o)]
*17. Show that the only sequences of orthogonal polynomials that satisfy a Rodrigues
formula p,(x) = K,[0(x)17'D"[o(x)p(~)], where p is a given polynomial, are the
Jacobi, Laguerre, and Hermite polynomials.
*7 CHEBYSHEV POLYNOMIALS
The U,,(x) defined by (21b) are also polynomials of degree m, called Chebyshev
polynomials of the second kind. Their theory is parallel to that of the T,,(x) and is
developed in Exs. D3—D7.
From the forms of the preceding explicit solutions, we see that the functions
T,(x) are eigenfunctions of the singular S-L system defined from (21) by the
boundary conditions that u’(—1) and w’(1) be finite. All solutions of (21) are
bounded at the singular points x
=
=
+1, as is apparent from inspection of
explicit solutions (21a) to (21b) and also from a calculation of the roots » =
=
0,
3 of the indicial equation 2v? — » =
=
0 of the normal form of (21). But only
multiples of the T,,(x) have bounded derivatives at the endpoints.
THEOREM 8. Among all monic polynomials P(x) = x" + URz) a,x" of degree n,
2'-"T.(x) minimizes max_)=,<,|P(x)| (Minimax property).
In order to establish the statement, it therefore suffices to show that for any
monic polynomial of degree n we have
Suppose this were not so. Then, we could find a monic polynomial p(x) of degree
n such that max_j<,<; |p(x)| < 2'~”. Now, the polynomial 2!-"7 (x) — p(x) is
of degree n — 1. We shall reach a contradiction by showing that this polynomial
has n distinct zeros.
To see this, notice that the polynomial 2!-"7 (x) takes alternately the values
+2!" at n + 1 points xo -l<x<
=
=
<x, = 1, which immediately
360 CHAPTER 11 Expansions in Eigenfunctions
EXERCISES D
1. Show that the DE [(1 — x2))u’}’ + AU. — x*)!7u = 0 can be reduced to a DE with
constant coefficients by setting v(8) = (sin 6)u(cos 8).
Show that the endpoint conditions lim,_.+; V1 — x u(x) = 0 give an S-L system with
eigenvalues X,,
=
=
n(n + 2), from the DE of Ex. 1.
Hence, with respect to the inner product ( f, g), this set of functions is a Euclid-
ean vector spacet (or “inner product space’’).
For real functions, the integral (6) in the definition of mean-square conver-
gence, is the inner product (f, — fi f, ~ f); hence it is the square of the distance
lh —fl = i -ff — f)'” between f, and fin the Euclidean vector space E.
Therefore, f, = f in the mean square, relative to p, means that the Euclidean
distance from f,, to fin E tends to zero.
This distance enjoys the properties of distance in ordinary space, including
the triangle inequality and the Schwarz inequality
The Schwarz inequality shows that fand g are orthogonal if and only if the angle
Ilr > Vide = (ff) - > [eer dp] + > [ce — Wor o,)]
Geometrically, the least-square approximation Lj. ¢,6, to fappears as the orthog-
onal projection of the vector f onto the subspace S of all linear combinations
N11 tess + YnOnof oy, %. This is because, in the orthogonal projection
onto a subspace S of a vector c issuing from the origin, the component ofc per-
pendicular to S$ is the shortest vector from S to c. The coefficients y, are given by
the direction cosine formulas of analytic geometry.
Completeness of the ¢, is defined as in §4, as the property that
"0, oo
f
k 1
+ The reader should familiarize himself with this notion; the space is not assumed to be finite-
dimensional.
362 CHAPTER 11 Expansions in Eigenfunctions
for every fin E. 1t has a simple geometric interpretation in any Euclidean vector
space E. The relation f = LP ¢,@, holds if and only if the distance ||f — 27 ¢¢,|
tends to zero as n — 00, That is, as in the proof of Theorem 3, the condition
for completeness is that we can approximate any f arbitrarily closely by finite
linear combinations ¢,¢, + + c,, of the orthogonal vectors ¢, whose
completeness is in question. This idea is most vividly expressed in terms of the
concept of a dense subset of a Euclidean vector space.
o,)"
k
> Cie — f
1
"eUGN-> | (f,
(Op oy)
k=1
|=o
must tend to zero as n — 00. Hence, if the sequence {¢,} is complete, then
(23’)
If Parseval’s equality fails, then strict inequality will occur in (24) for some f in
E. For such an f, we have by Theorem 1
for any choice of y,. Since 6 is independent of n, this shows that the ¢, cannot
be a complete set of orthogonal vectors. This gives another proof of Theorem
3, which we now restate for an arbitrary Euclidean vector space.
9 COMPLETENESS OF EIGENFUNCTIONS
ni(x — a) O(1)
u,(x) =
(b — a) | (b — a) | n
Since the series 1 +34 + $+ -°-°+> + 1/2 +- converges (to 1/6), this
implies the following lemma.
LEMMA. Let u,(x) be the nth normalized eigenfunction of any regular S-L system
in Liouville normal form, with a’B’ # 0, and let
n(x — @)
bn(x) =
(b
— a) | (b
— a) |
Then the , are an orthonormal sequence, and
Since the cosine functions are complete (by Corollary 3 of Theorem 3), it
follows from this lemma and Theorem 10 that the eigenfunctions of any regular
S-L system in Liouville normal form with a’6’ # 0 are a complete set of ortho-
normal functions.
As shown in Ch. 10, §9, the transformation to Liouville normal form, applied
to the (normalized) eigenfunctions, carries the inner product
(.¥) = J b(x)Y(x)p(x)dx
into the inner product
(u, v) = f a
u(x)v(x) dx
Therefore,t the change of variable that leads to a Liouville normal form carries
complete orthonormal sequences, relative to a weight function p, into complete
orthonormal sequences. Hence, the eigenfunctions of regular S-L systems not
in Liouville normal form are also complete.
Finally, since similar arguments cover the case a’6’ = 0, we have the following
result.
THEOREM 11. The eigenfunctions of any regular S-L system are complete in the
Euclidean vector space of square-integrable continuous functions, on the interval
a=x Sb, relative to the weight function p.
+ Since distance and convergence are defined in terms of inner products in any Euclidean vector
space.
10 Hilbert Space 365
The set of real numbers differs from the set of rational numbers by the com-
pleteness property that every Cauchy sequence of real numbers is convergent.t
This property of completeness has an analog for Euclidean vector spaces (and,
more generally, for metric spaces).
The space E is called complete when, given any Cauchy sequence { f,}, there exists
a vector fin E such that |/f, — {|| ~ 0 as n > oo. A complete Euclidean vector
space is called a Hilbert space.
Any finite-dimensional Euclidean vector space is complete, but the Euclidean
vector space of continuous square-integrable functions defined in §8 is not com-
plete, as will appear presently.
Example 2. Let (£;) denote the Euclidean vector space of all infinite
sequences a = {a,} = (a), a9, 43, .. .) of real numbers which are square-sum-
2
mable, that is, which satisfy Xa, < +00, The vector operations on these
sequences are performed term-by-term, so that a + b is the sequence (a, + 4,
dy + do, dg + bs, ...). Inner products are defined by the formula
Proof. Our problem is to prove completeness. To this end, let {a"} be any
Cauchy sequence of square-summable sequences. That is, let
For each fixed k, the sequence of real numbers aj (n = 1, 2,.. .) (the Ath com-
ponents of a”) is a Cauchy sequence and, therefore, converges to some real num-
ber a,. Let a = {a,} (k = 1, 2, 3,...). We must prove that the sequenceais
square-summable and that a” > a in the Euclidean vector space.
Since ||la"|| — |la"|}| <= lla” — a™|| by the triangle inequality, it follows that
the sequence ||a"|| is bounded. Let \/M be an upper bound. Then, for all
integers n, we have Dj-1 (az)? <= M. Letting n — oo in this finite sum, we get
D1 (a,)* = M. SinceN is arbitrary, it follows that al? = D2, (@,)? < M.
m wo
»
nt+1 | (bis bn)
By the Bessel inequality (24), the series D714 [(f ¢))°/(¢s, $)] of positive num-
bers is convergent; hence, the last sum in the preceding display tends to zero as
n — 00. That is, the sequence {g,} is a Cauchy sequence.
It follows that #, being complete, contains a vector g to which the g,
converge; let h = f — g = lim, (f — gn). Then, since (f — gn, 6) = 0 for
all m = k, we have in the limit, as m — 00, (h, @,) = 0 for all k. By Theorem 9
(Parseval’s equality), h = 0 for all f if and only if {,} is complete. This completes
the proof.
f, g) = | (6
~
| f "flxg(x)dx = f .fs) at
Hence, the correspondence f — f also preserves inner products (up to a constant
normalizing factor). Therefore, it also preserves lengths (f, f 2, In particular,
= 0 implies that {7 f() dt = 0. Hence, since f is continuous, it also implies
that f(t) = 0. In conclusion, we have proved the following result.
LEMMA 2. The Euclidean vector space @{a, b] (—0 <a <b < +00) can be
embedded in the Hilbert space (€.), with preservation of vector operations and inner
products.
(f, o1)
for fin E, defines an embedding of E as a subspace of (€9). In the same way as
before we obtain the following lemma.
We are now ready to prove Theorem 10. But to bring out more clearly the
idea of the proof, we first treat a special case.
We define a sequence of orthonormal vectors of a Hilbert space to be an
orthonormal basis if and only if it is complete.
LEMMA. Let {@,} be an orthonormal basis in the Hilbert space Ff. Let {W,} be an
orthonormal sequence in #f satisfying the condition
Proof. If the sequence y, were not a basis, then we could find a nonzero
vector h-orthogonal to every y, by Theorem 12. The inner product of this vector
with @, would be given by
co
for some integer N. Then every element h of 7 orthogonal to $y, Gy and toWys1,
Vn+2, .. must vanish.
Proof. Any such element h satisfies the inequality (29) for k > N. Indeed,
summing over all k, we have, since (h, ¢,) = 0 fork = 1, 2,...,N,
eo
all?
=
=
must vanish.
oo
converge to f(x) in the mean square. That is, the metric completion of the space
@[a, 6] is precisely the Hilbert space £,[a, 6] of all functions on [a, 5] whose
squares are Lebesgue integrable. It is this space that is really appropriate for
the theory of expansions in eigenfunctions.
EXERCISES E
{Hint: Show that linear combinations of the functions f(x) = |x — c| form a dense
subset, of @[a, 6], and apply Ex. 2.)
> (f aoa) ae = OH
(Hint: Let g(x) = x — a — tay (S% (8) dt)®, and show that q(x) = 0 by establishing
q(x) = 0 and f® q(x) dx = 0, ¢ continuous, applying Ex. 3.]
Assuming the equality Df2,1/k? = 2/6, infer from Ex. 4 that the trigonometric
functions cos kx, sin kx, k = 0, 1, 2, . , are complete on (—7, =).
n=
f; x"f(x) dx
Show that if all the moments m, vanish, f(x) = 0. [HINT: Use the Weierstrass Approx-
imation Theorem.]
*7 Prove the completeness of the eigenfunctions of any regular S-L system with bound-
ary conditions u(a) = u(b) = 0.
*8 Let G(fi, fa .»f,) = det (@,), where a, = (f;, f). Show that the minimum of
If — © afill is equal to GU, fi, fo .? td /C des . » f). (Hint: Interpret the
determinants as volumes.]
Show by a counterexample that Theorem 10 does not remain valid if the y,, are
allowed to be nonorthogonal unit vectors.
11 Show that the first-order complex DE iu’ + (q(x) + Alu = 0, a = x < 3b, with the
boundary condition u(a) =
=
u(b) and i = —1, has a complete sequence
of eigenfunctions for any real continuous q(x).
12 Show that linear combinations of the trigonometric functions cos kx, sin kx are dense
in —a < x <7, relative to any continuous positive weight function.
APPENDIX A
LINEAR SYSTEMS
1 MATRIX NORM
where x(f) is a column vector of length n and A(#) ann X n matrix, both depend-
ing on ¢. In Ch. 6, we proved the existence and uniqueness of solutions of (1) in
any interval of continuity of A(/), for any initial x(0) = c. In this appendix, we
derive some properties of solutions of linear systems (1) whose proofs depend
on deeper properties of matrices.
As before, x’ (é) is the limit
where the limit is taken separately in each component Ax,/At. Here we extend
this definition, calling the norm of any matrix A the real number
| |
(2) All = sup
| |
where
+ We use standard notation here: A” signifies the transpose of A, and x" the inner product of x
with itself; see Birkhoff-MacLane, Chs, 8—10 for the facts assumed in this Appendix.
371
372 APPENDIX A Linear Systems
The equation ||J|| = 1, the triangle inequality |A + Bl] = ||Al| + |B], the
multiplicative inequality |ABl| < ||Al| - |B}, and |tAf}' = [2] - ||All all follow
quite directly from (2). From these relations, in turn, one can prove the conver-
gence of the exponential series
for any matrix A and any scalar ¢. It suffices to copy the usual proof for real or
complex exponential series, replacing absolute values by norms throughout.
2 CONSTANT-COEFFICIENT SYSTEMS
d (ey
tA uA
Because of the commutativity of all terms, we have, as in the scalar case, ee
= et +uyA
and, hence,
co
(4 _ De'A —_ ( At*A®
Je
et tAnAa _ eA
>
=
=
k=1
kl!
Dividing the series in parenthesis through by Ai, we get the identity matrix J
plus a series of matrices whose norms are bounded by Af”'a*/(kl), k = 2,
3, , where a =
=
|| All. Hence, the norm of the sum tends to zero with Aj,
proving (4).
From (4), there follows a very beautiful result.
Complex Solutions. The preceding results hold for complex as well as for
real solution vectors, coefficient-matrices, and independent variables, provided
that we define the norm of a complex vector z = (z,, , Z,)' as the square root
of the Hermitian inner product
A,
Zz = x, a tatty = ant
of z with its conjugate transpose z” = (z*, , z%). Hence, we can apply them
to complex eigenvalues and eigenvectors of real matrices.
dizg/‘dt =
bz, + AZo
dt ( }=(
22 b a
I 29
where the variables #, z,, zg may be real or complex, but the coefficients a and b
a —b
are real. The characteristic polynomial of the coefficient-matrix A =
b a
vector-valued functions
ib
“( \] and “(7): c=at ib, ch=a—
are a basis of complex solutions of the system (5).
In general, every matrix A with distinct eigenvalues has a basis of real or com-
plex (column) eigenvectors v,, with corresponding eigenvalues A, (j = 1,. ,n).
Aye
The vector-valued functions ¢,(i) = ¢ v, then form a basis of solutions of the
constant-coefficient linear system z’(t) = Az in the entire complex #-plane.
The Secular Equation. We now generalize the concept of the secular equa-
tion from second-order linear systems with constant coefficients to nth-order
systems. Namely, if p,(A) is the characteristic polynomial of the matrix A, we
define the secular equation of the constant-coefficient system x’(/) = Ax to be the
DE p,(D)u = 0, where D denotes d/dt, and p,(D) is the scalar differential oper-
ator obtained from the characteristic polynomial |A — AZ| = p,(A) of A by sub-
stituting D for X. We first prove Theorem 3 of Ch. 5.
374 APPENDIX A Linear Systems
THEOREM 3. Any component x,(t) of any vector solution x’ (t) = Ax satisfies the
secular equation p,(D) = 0.
Proof. Rewrite the given system as Dx = Ax, and let C(D) = ||C,,(D)|| denote
the matrix of cofactors of the entries ap
J
DéJ , of A — DI. Premultiplying the
equation (A — D)x = 0 by the matrix C(D)", we get the vector DE
But (6) asserts precisely that for eachj = 1,. » 0, pa(D)x, = 0, q.e.d.
dx,
di
= AX, or
di
7 = AX, + X41
Aj 1 0
0 A, 1
J
0 0 A,
0 0 90 A,
1 t P72) 8/3!
eh = ov
0 1 t 7/2!
0 0 1 t
0 0 0 1
Since negative exponentials die out faster than any power, it follows that e >
0 as ¢ > 00 if and onlly if every eigenvalue of A has a negative real part. Finally,
since e4 = P™'e!P, it follows that the same is true of e“. (The matrices A and
J = PAP™' have the same eigenvalues.)
3 THE MATRIZANT
We next consider the general case of A(é) variable but continuous in (1). If
Xx) , , X,,(¢) is any list of m (column) vector solutions of (1), the nm X m matrix
X() composed of these columns will, trivially, satisfy the matrix DE X’() =
A(X. In particular, we can construct in this way an n X n matrix solution of
The matrix function M(/) so defined is called the matrizant of the system (1). By
(4), the matrizant of the constant-coefficient system x(t) = Ax is just ¢.
Since the multiplication of matrices is associative, the vector-valued function
Mic = x(?) will satisfy x(0) = M(0)c = Ic = c for any given c. Since
it will also satisfy (1). By the uniqueness theorem proved in Ch. 6, it follows that
every solution of (1) has the form M()c.
The determinant det M(t) of the matrizant also has a remarkable property.
We know by (7) that
or
Now, passing to the limit as At > 0, we get [using the alternative notation | M(i)|
for det M(O):
d|M@|
= Tr A|M(@|
d
This is the case, for example, with systems (1) that come from nth-order DEs
having the normal form
Next, we define the adjoint of the system (1) to be the first-order linear system
~ mu nd)
We consider the solution curves of (9) as lying in the dual space of linear func-
tionals on the space of x(d), given by the linear forms
(10) f(x)
= Xf, x, = (@, x) = fx (inner product)
THEOREM 5. In order that [{(), x(é)] be constant for every solution x(0) of the
vector DE (1), tt is necessary and sufficient that (9) hold.
4 Floquet Theorem; Canonical Bases 377
4 b soso
| = Xposs + Tsjoso
= 2 HiOanlbaild + 2 fiOn(
Since there exists a solution of (1) with x(é) = c for any é) and c, the condition
that (£(, x(é) be constant for all solutions of (1), and hence that the derivative
above vanish identically, is precisely that
We next consider periodic linear systems, that is, systems of the form (1) with
/
U = 0, v’ = (~A + 16d cos 2x)u = 0
u” + piu = 0, pli + x) = pl
THEOREM 6. [If the matrizant M(T) of (1) has r independent eigenvectors and
(12) holds, then (1) has r linearly independent solutions satisfying
Cyt,
(14) x(t) = ey, @, where y.@ + T) = y,0
fp(idt = Sip@dt = 0
In particular, referring back to Ch. 10 (Example 4 in §1, Ex. A5, and §8), we
can conclude that there exist infinite sequences of even and odd Mathieu func-
tions, and that the functions in each sequence have periods 27 and z alternately.
Finally, let ¢ = a@ be an isolated singular point of the matrix A(), considered
as a function of the complex independent variable ¢. By this we mean that all a,,(é)
are analytic in some punctured disk 0 < | — a| < p. Much as in Ch. 9, we can
consider the effect of making a simple loop ¢ = @ + és around ¢ = aon some
basis of solutions of the linear system (1); we do not assume periodicity. If the
basis of solutions has for initial values x,(f)) = e,, where e, = 6, 6,,,) is the
ith unit vector, the vectors resulting from going once around the loup are
the column vectors of the “matrizant’’ of (1) for the given loop. In any event,
we can write
THEOREM 7. [If the circuit matrix of A(t) in (1) for the isolated singular point
t =a has r independent eigenvectors, then (1) has r solutions
where the £,(t) are holomorphic in 0 < |t — a| < p for some p> 0.
4 Floquet Theorem; Canonical Bases 379
EXERCISES
00 1 0 1 0 A 1 0
(a) 1 0 0 (b) 001 (c) 0A 1
0 1 0 000 0 0A
2
0 0
te
(a (_° t
0
} of 0 t
) (c) 2
0
e
0
0 0
dx, dx
— = —2x, + Xs ——
= Xx. nn 2x.
dt
dx,
—= X, 1 — 2x, + x, 1 G
=
,n— 1)
dt
Show that the matrizant M(T) of any DE of Hill’s type has determinant |M(T)| = 1.
[HinT: Show that, in the phase plane, TrA(/) = 0.]
APPENDIX B
BIFURCATION
THEORY
1 WHAT IS BIFURCATION?
d x a b x
(1)
dt ( }=( I
J c d J
(2) ‘ ES 2 + Prolx)—glx)lu = 0
provides another classic illustration, to which Chapters 10 and 11 were devoted.
This appendix will introduce a typical concept associated with this parameter
dependence: that of bifurcation.
A striking example of bifurcation is provided by the van der Pol equation
(3) ¥—-— 21 — yy +y = 0
+ See the books by Hale and Chow-Hale, Guckenheimer-Holmes, and Thompson-Stewart listed in
the bibliography.
380
2 Poincaré Index Theorem 381
(4) X+ px =C
in the stable case 4 < 0, all solutions approach the same equilibrium solution
x = C/p; when p < 0, they all diverge from it.
By fixing one parameter (e.g., g), one can also apply the concept of bifurca-
tion to DEs depending on two parameters, such as
(5) ¥+ px + qx = 0
Setting A = 1, the phase portrait of (6) has a stable focal point for B << 2 =
1 + A’, and an unstable focal point for B > 2; hence B = 2 is a bifurcation
point for B in (6), if A = 1.
Let (X(x,y),Y(x,y)) be a plane vector field, and let be a simple closed curve
which does not go through any critical point of the vector field. By the Jordan
curve theorem,t the interior of ¥ is a well-defined, simply connected region; we
will assume the vector field to be of class @? in y and its interior.
(7) Ky) = = dy —1
2a Y
LEMMA I. Let ’ and y” enclose domains D’ and D", whose union is a simply
connected domain D with boundary y, as in Figure B.1. Then
The proof depends on the fact that y consists of y’ U y” with their common
segment y’ M ” deleted; since this segment is traversed in opposite directions
by y’ and y”, the contributions from it cancel.
D Y Y D
In other words, /(y) is the segn of the determinant |a,| a,d, - b,c, of the
matrix of the linearization of the DE % = X(x,y),j = Y(x,y) unless this critical
point (x,,y,) is degenerate, in the sense that |A,| = 0. Getting down to cases, we
see that focal, nodal, and vortex points have index +1, while saddle points have
index —1.
The proof of Lemma 2, in the case of nondegenerate critical points, is
straightforward but tedious. One first observes that, since the vector field
X(x) € @?, there is some neighborhood of (x,,,) in which not only is there no
other singular point, but the direction of X(x) differs by less than 7/4 radians from
that of the linearized vector field
One then takes up individually each of the non-degenerate cases of Figure 5.5;
see the exercises at the end of this appendix.
The reason why the degenerate case |A,| = 0 has been excluded in Lemma
2 is easily explained by examples. First, the degenerate field Z(z) = z"(n > 1)
has index I(y) = n for any contour ¥ containing the origin. Thus (x? — y*, 2xy)
has index 2, (x* — 3xy", 3x°y — y*) has index 3, and so on. And again, the vector
field (x,x* sin(1 /*)) has infinitely many critical points where the angle W is unde-
fined, in any neighborhood of the origin. However, using Lemmas 1 and 2, we
can prove the following Poincaré index theorem:
THEOREM 1. Let ¥ be any simple closed curve not containing any degenerate crit-
ical point of the plane vectorfield X(x). Then y contains only a finite number of critical
points x, = (x, y,), and
OX/dx OY/dy
(10) Ky) = > L=
» “Bn
J
J
|aXx/ay a¥/ay|,
is the sum of the (Poincaré) indices of these critical points.
3 HAMILTONIAN SYSTEMS
oH oH
(11) 4G = Op,’
—
=o
t
0q;’
q = dH/dp = p/m
(which checks), and m¥ = p =
=
— 0H/dq = — V(x) = F(x), where F(x) =
— q(x) is the force.
Likewise, for the pendulum discussed in Chapter 5, §3, letting the (general-
ized) position coordinate be @ = q, and p = £6, we have the energy function or
Hamiltonian
x = O0orx
=
=
— 2u; when x = py = 2 + Hu? + y*) = Owheny = + V3p.
Hence, if » # 0, the system (13) has four bifurcation points: one located at
the origin, and the other three at the vertices (— 2p, 0) and (u,+V3n) of an
equilateral triangle. (When pz = 0, there is only one critical point. This is at the
origin, and is degenerate.)
One easily verifies that the system (13) is Hamiltonian because, for
—-—-(x
(14) +9)
+ ny — 3
The first statement follows from a basic theorem of vector analysis, which
states that a velocity field X(x) defines a volume conserving flow x‘(t)= X(x) if
and only if its divergence is zero.t For (11), evidently
i=]
Ox; r=] 9q,0p, =] 94:
The second statement follows since
n n
oH
| oH
4 traso) = 3 Hs, = » =] 2 1=]
q, 1 »
=]
—
Op,
p,
n
0H
*. oH aH
» aH )-
( ~ 8a,
=
a x—& x E
)+
( }-| I
a1 a9
(17)
dt y— 7 ~ 1 — Ag9 y7 0
+ Courant and John, vol. II, p. 602. The fact that the Hamiltonian flows are volume-conserving is
called Liouville’s Theorem.
386 APPENDIX B Bifurcation Theory
The eigenvalues of the coefficient matrix of the linearization (17) at (x,,y,) are
the roots of the quadratic equation
x ay} a2
(18)
=
=
9° — &)dg9 = — det
Q19 Ag9
A(x,y)ishorizontalatcriticalpoints,det |a)
that since the surface z =
=
4 HAMILTONIAN BIFURCATIONS
V=y—pe”, re = x? + ¥?
(20)
+ Of course, this “‘time’’ parameter » is unrelated to the variable ¢ associated with the “time’’ deriv-
atives ¥ = dx/dt and y = dy/dt.
5 Poincaré Maps 387
2un. Bifurcation takes place when the line & = — 2yn in the (7,)-plane touches
the graph of § = e™. For larger slopes » > y,, there are two points of intersec-
tion and the surface z Vry-—
pe has corresponding critical points: a
=
=
The Poincaré Index. Theorem 3 explains why the critical points arising
from the preceding “‘bifurcation’’ are neither nodal points nor focal points. The
fact that they are paired and of opposite indices is, however, a simple conse-
quence of the Poincaré index theorem (Theorem 1), as we shall now explain.
For a general one-parameter family of plane autonomous systems,
consider the variation with » of the Poincaré index of a fixed simple closed curve
‘y, as defined by (7). Unless the curve passes throughacritical point of the system
(22) for some value of yu, the angle y(y,u) between the vector (X(x,9;4), Y(x,y3H))
and the horizontal will vary continuously with 4. Hence, the contour integral
Tye) = p dy(y,n)
will also vary continuously, and cannot jump by an integral multiple of 27. We
conclude
THEOREM 4. Inside any simple closed curve not passing through a critical point,
new nondegenerate critical points appearing or disappearing at bifurcation values p
must arise or be destroyed in adjacent pairs having indices of opposite signs.
5 POINCARE MAPS
The concept of a Poincaré map arises in three different contexts: (i) limit
cycles (Ch. 5, §13) and other periodic orbits of autonomous systems, (ii) periodically
388 APPENDIX B Bifurcation Theory
(26) ¥ + px
+ qx A cos wt
=
=
yield familiar examples (in the phase plane), and (ii) systems with periodic coef-
ficient-functions, like the phase plane representation
and has a periodic orbit of radius r = @ and period 27 whenever g(a) = 0. This
orbit is evidently stable (a ‘‘limit cycle’’) if g’(a) < 0, and unstable if g’(a) > 0.
When yz = 0, the origin is a stable equilibrium point toward which all orbits
tend. But when yu exceeds 1, the origin is unstable, and a limit cycle occurs when-
ever (sin r)/r = 1/y. The nearest limit cycle is locally stable; the next is unstable,
6 Periodically Forced Systems 389
the third is stable, and so on. Looking more closely at the intersections of the
curve s = (sin r)/r with s 1/u, we see that new periodic orbits arise in pairs
=
=
as 4s increases, one stable and the next unstable. The analogy with bifurcations
of equilibrium points is obvious!
Just as in Ch. 4, §7, this periodic orbit is strictly stable if an only if every eigen-
value A, of A has a negative real part, so that jer? <i.
considered in Appendix A, §4, also give rise to Poincaré maps. Indeed, every
such system has the trivial equilibrium periodic solution x(t) = 0, and the linear
transformation x() = M(T) x = x(t + T) defined by the matrizant (Floquet
matrix) referred to in Theorem 6 of Appendix A is the Poincaré map associated
with this equilibrium solution. Clearly, the equilibrium solution (‘‘equilibrium’’)
is strictly stable if and only if all the eigenvalues ), of this Floquet matrix have
magnitudes |A,| <1.
Example 5. Consider for example the Mathieu equation of Ch. 10, §1,
Example 4:+
(33) u”
+ (« + 16 dcos 2x)u =0
¢(0) = 1, $'(0) = 0
(33’)
(0) = 0, v’(0) = 1
determinea basis of solutions of (33), the first of which is an even function and
the second odd. Moreover, the associated Floquet matrix has déterminant one
(see Ex. 8 of Appendix A). Hence its eigenvalues are the roots of a real char-
acteristic equation of the form
(33”) \? — 2B.+1=0
LEMMA The Mathieu equation (33) is stable or unstable according as |B| < 1
or |B| > J.
For, since \jAg = 1, they are complex conjugate and on the unit circle if
|B| < 1, but real and one exceeding one if |B{ > 1. On the other hand, the
values B = +1 yield the eigenvalues ); = +1 associated with the Mathieu func-
tions of periods x and 27. There follows:
(34) K+ ck — x + x? = Acoswit.
In the unforced case A = 0, (34) evidently has three equilibrium solutions: x()
=0 and x(f) = +1. Of these, x = 0 is unstable, while x = +1 are stable. The
trajectories in the phase plane are easily drawn in the Hamiltonian case A = ¢
= 0: they are the level curves of the energy function H = x* — 2x” + 2uv”, with
“separatrices” v = +x* V1 — (x*/2) through the origin. These separatrices sep-
arate the periodic orbits corresponding to local oscillations about one stable
equilibrium point from the oscillations about all three, whose amplitudes exceed
2\/2. The tangents to all trajectories in the phase plane are horizontal where
they cross the vertical lines x = 0 andx = +1.
When A = 0 but c > 0, most trajectories in the (x,v)-plane spiral clockwise
from ‘‘infinity” into the two stable focal points at (+1,0). There are, however,
two special trajectories which originate (at £ = —0o) in the unstable saddle point
at (0,0), each of which spirals into one of the focal points at (+1,0). Likewise,
there are two special separating trajectories which spiral in from “infinity’’ but
6 Periodically Forced Systems 391
come to rest at (0,0). These separate the trajectories which spiral into (1,0) from
those which spiral into (— 1,0).
In the forced case A # 0, a rich variety of qualitatively different kinds of
behavior can arise, depending on the choice of the three coefficients ¢, A, and
w in (34). These can be explored most efficiently by using modern computers to
compute and display the sequences of points (x(n7),x(nT)) in the phase plane
arising by iterating the Poincaré map, applied to selected initial states (x,,%,) for
selected choices of c, A, and w.
Of particular interest here are the fixed points of the Poincaré map, that is,
points such that (x(T),x% (T)) = (x(0),% (0)), on periodic orbits of period T. For
example, with small A, there are three such fixed points for (34): two stable fixed
points (on stable periodic orbits) near (+ 1,0), and an unstable fixed point near
the origin. A family of trajectories asymptotic to the unstable fixed point forms
a manifold separating trajectories attracted to the two stable periodic orbits.
EXERCISES
1. Explain why, for fixed q > 0, the value p = 0 is a “bifurcation value”’ of p in (5).
2. (a) Show that for given A and B, (A, B/A) is the only equilibrium point of (6).
(b) Derive the system (6) of variational equations for perturbations of the equilib-
rium solution.
(a) Show that the system ¥ = 2y, ¥ = 3 — 3x leaves invariant the energy (Hamilton-
ian) function V = x° — 3x + yx.
(b) Show that this system has a saddle point at (1,0), a vortex point at (— 1,0), and
no other critical points.
(c) Show that the cubic curve x> — 3x + 9? = —2 through the saddle point is a
“separatrix.”
(d) Sketch the phase portrait of the system.
(a) Show that » =0 is a bifurcation value for the one-parameter family of Hamil-
tonian systems with Hamiltonians H = y* + 3yux? = 3x.
(b) Describe the qualitative change that takes place in the phase portrait when pu
changes sign.
1
(a) Solve this DE explicitly for the endpoint conditions u(0) = u(1) = 0, when
€ > 0 and whene
< 0.
(b) Contrast the increasingly irregular behavior of solutions as « | 0 with the smooth
“boundary layer” behavior as € t 0.
BIBLIOGRAPHY
GENERAL REFERENCES
Ahlfors, L. V., Complex Analysis, 2nd ed. New York: McGraw-Hill, 1966.
Apostol, Tom M., Calculus, 2 vols., 2nd ed. New York: Blaisdell, 1967, 1969.
Birkhoff, Garrett, and S. MacLane,
A Survey of Modern Algebra, 4th ed. New York: Mac-
millan, 1977.
Carrier, George F., Max Krook, and C. E. Pearson, Functions of a Complex Variable. New
York: McGraw-Hill, 1966.
Courant, Richard, and D. Hilbert, Methods of Mathematical Physics, Vel. J. New York:
Wiley-Interscience, 1953.
Courant, Richard, and F. John, Introduction to Calculus and Analysis, 2 vols. New York:
Wiley, 1965, 1972.
Dwight, H. B., Tables of Integrals and Other Mathematical Data, rev. ed. New York: Mac-
millan, 1947.
Fletcher, A.,
J. C. P. Miller, and L. Rosenhead, Index of Mathematical Tables, Vol. I and II,
2nd ed. Reading, Mass.: Addison-Wesley, 1962.
Hildebrand, Francis B., Introduction to Numerical Analysis, 2nd ed. New York: McGraw-
Hill, 1974.
Hille, Einar, Analytic Function Theory, Vol. 1. Waltham, Mass.: Blaisdell, 1959; Vol. 2,
1962.
Picard, E., Traité d’Analyse, 3 vols., 2nd ed. Paris: Gauthier-Villars, 1922-1928.
Rudin, Walter, Principles of Mathematical Analysis, 2nd ed. New York: McGraw-Hill, 1964.
Widder, David V., Advanced Calculus, 2nd ed. Englewood Cliffs, N.J.: Prentice-Hall,
1961.
392
Works on Ordinary Differential Equations 393
Cesari, L., Asymptotic Behavior and Stability Problems in Ordinary Differential Equations, 2nd
ed. New York: Academic Press, 1963.
Chow, S.-N., and J. R. Hale, Methods in Bifurcation Theory. New York: Springer, 1982.
Coddington, Earl A., and N. Levinson, Theory of Ordinary Differential Equations. New
York: McGraw-Hill, 1955.
Cronin, Jane S., Differential Equations: Introduction and Qualitative Theory. Dekker, 1981.
Davis, Philip J., and Philip Rabinowitz, Numerical Integration. Blaisdell, 1966.
Gear, C. William, Numerical Initial Value Problems in Ordinary Differential Equations. Pren-
tice-Hall, 1971.
Guckenheimer, John, and Philip Holmes, Nonlinear Oscillations, Dynamical Systems, and
Bifurcations of Vector Fields. New York: Springer, 1983.
Hale, Jack K., Ordinary Differential Equations. New York: Wiley, 1969.
Hartman, Philip, Ordinary Differential Equations. New York: Wiley, 1964.
Henrici, Peter, Discrete Variable Methods in Ordinary Differential Equations. New York:
Wiley, 1961.
Henrici, Peter, Error Propagation for Difference Methods. New York: Wiley, 1963.
Jordan, D. W., and P. Smith, Nonlinear Ordinary Differential Equations. Oxford: Oxford
University Press, 1977.
Kamke, E., Differentialgleichungen: Lisungsmethoden und Lésungen. Leipzig: Akademische
Verlag, 1943; Chelsea reprint, 1971.
Kaplan, Wilfred, Ordinary Differential Equations. Reading, Mass.: Addison-Wesley, 1958.
LaSalle, J. P., and S. Lefschetz, Stability by Liapounov’s Direct Method, with Applications.
New York: Academic Press, 1961.
Lefschetz, Solomon, Differential Equations, Geometric Theory, 2nd ed. New York: Wiley-
Interscience, 1963.
Magnus, Wilhelm, and S, Winkler, Hill’s Equation. New York: Wiley-Interscience, 1966.
McLachlan, N. W., Ordinary Non-Linear Differential Equations in Engineering and Physical
Sciences, 2nd ed. New York: Oxford University Press, 1956.
394 Bibliography
Pliss, V. A., Nonlocal Problems of the Theory of Oscillations. New York: Academic Press,
1966. (Russian ed., 1964.)
Poincaré, H., Les Méthodes Nouvelles de la Mécanique Céleste, Vol. I-III, New York: Dover,
1957.
Reid, W. T., Ordinary Differential Equations. New York: Wiley, 1971.
Sansone, G., and R. Conti, Non-Linear Differential Equations. New York: Pergamon, 1952.
Simmons, G. F., Differential Equations with Applications and Historical Notes. New York:
McGraw-Hill, 1972.
Stoker, J.-J., Nonlinear Vibrations in Mechanical and Electrical Systems. New York: Wiley-
Interscience, 1950.
Thompson,
J. M. T., and H. B. Stewart, Nonlinear Dynamics and Chaos. New York: Wiley,
1986.
Wasow, W., Asymptotic Expansions for Ordinary Differential Equations. New York: Wiley-
Interscience, 1966.
INDEX
395
396 Index
Weierstrass approximation theorem, 353 Well-posed problem, 24, 64, 170, 174
Weierstrass convergence theorem, 195 Well-set see Well-posed problem
Weight function, 302, 309, 352 Wronskian, 43