Errata

3 downloads 0 Views 9MB Size Report
the image of the real axis under the transformation m. Az + B. - Cz + D '. (3.3.24) where A = ¢ (b , A) ...... [1] APOSTOL, T., Mathematical Analysis, Second Edition,.
Volterra Stieltjes Integral Equations and Generalized Ordinary Differential Expressions Preprint, 1983

Errata •

p.90: Proof of Theorem 2.3.1; delete the remark in parentheses “since -y(t) is a solution” (comment: clearly this is not assumed; if y(t) < 0 in (2.3.11) define g(t) by the negative of the right-side of (2.3.12)) • p.244, Statement of Theorem 5.2.2: Replace the word “precisely” by “contained in” (comment: as this is what is proved there).

Quest' opera

e

umilmente dedicata

ai miei cari genitori Giosafat Oliviana e al mio fratello Marco A.M.D.G.

e

PREFACE The aim of these notes is to pursue a line of research adopted by many authors (We Feller, M.G. Krein, 1.5.

Kac,

F.V. Atkinson, W.T. Reid, among others) in order to develop a qualitative and spectral theory of Volterra-Stieltjes integral equations with specific applications to real ordinary differential and difference equations of the second order. We begin by an extension of the classical results of Sturm (comparison theorem, separation theorem) to this more general setting.

In chapter 2 we study the oscillation theory

of such equations and, in Chapters 3,4,5, apply some aspects of it to the study of the spectrum of the operators generated by certain generalized ordinary differential expressions associated with the above-mentioned integral equations. In order to make these notes self-contained some appendices have been added which include results fundamental to the main text.

Care has been taken to give due credit to those

researchers who have contributed to the development of the theory presented herein - any omissions or errors are the author's sole responsibility. I am greatly indebted to Professor F.V. Atkinson at whose hands I learned the subject and I also take this opportunity to acknowledge with thanks the assistance of the Natural Sciences and Engineering Research Council of Canada for continued financial support.

My sincere thanks go to Mrs. Frances Mitchell

VI

for her expert typing of the manuscript. Finally, I am deeply grateful to my wife Leslie Jean for her constant encouragement and patience and I also wish to thank Professor A. Dold for the possibility to publish the manuscript in the Lecture Note series.

Angelo B. Mingarelli Ottawa, April 1980.

TABLE OF CONTENTS

x

INTRODUCTION CHAPTER 1 Introduction

1

1.1.

Comparison Theorems for Stieltjes IntegroDifferential Equations .. . . . . . . . . . . . . . . . .

1.2.

Separation Theorems

20

1. 3.

The Green's Function

25

4

CHAPTER 2 Introduction 2.1.

Non-Oscillation Criteria for Linear Volterra-Stieltjes Integral Equations

28 29

2.1A. Applications to Differential Equations

52

2.1B. Applications to Difference Equations

60

2.2.

74

Oscillation Criteria

2.2A. Applications to Differential Equations

80

2.2B. Applications to Difference Equations

82

2.3.

An Oscillation Theorem in the Nonlinear Case . . . . . . . .. . . . . .. . .. . .. . . . . . . . . .. . . . . .

Addenda

87 113

CHAPTER 3 Introduction

118

3.1.

120

Generalized Derivatives

VIII

CHAPTER 3 (continued) 3.2.

Generalized Differential Expressions of the Second Order

123

3.3.

The Weyl Classification

129

3.4.

Applications

143

3.5.

Limit-Point and Limit-Circle Criteria

147

3.6.

J-Self-Adjointness of Generalized Differential Operators . . . . . . . . . . . . . . . . • .

156

3.7.

Dirichlet Integrals Associated with Generalized Differential Expressions

180

3.8.

Dirichlet Conditions for Three-Term Recurrence Relations ....•..•...•....•.•.

183

CHAPTER 4 Introduction

197

4.1.

Sturm-Liouville Difference Equations with an Indefinite Weight-Function .

199

4.2.

Sturm-Liouville Differential Equations with an Indefinite Weight-Function

212

CHAPTER 5 Introduction

225

5.1.

The Discrete Spectrum of Generalized Differential Operators . . . . . . . . . . . . . . . . . .

226

5.2.

The Continuous Spectrum of Generalized Differential Operators .. . . . . . . . . . . . . . . . .

242

1.1.

Functions of Bounded Variation

256

1.2.

The Riemann-Stieltjes Integral

258

1.3.

General Theory of Volterra-Stieltjes Integral Equations

264

1.4.

Construction of the Green's Function

273

APPENDIX I

IX

APPENDIX II ILL

Compactness in L

P

280

and Other Spaces

APPENDIX III III.l.

Eigenvalues of Generalized Differential Equations .

292

III. 2.

Linear Operators in a Hilbert Space

296

III. 3.

Linear Operators in a Krein Space

299

III. 4.

Formally Self-Adjoint Even Order Differential Equations with an Indefinite Weight-Function

.

303

BIBLIOGRAPHY

309

Subject Index

318

INTRODUCTION Let p,q: 1-+ IR, p(t) > 0

a.e.

Lebesgue measure) and lip, q E L(I)

(in the sense of [a,b]

where I

c

IR

Consider the formally symmetric differential equation

= 0,

(p(t)y')' - q(t)y

t

E

(1)

1.

By a solution of (1) we will mean a function y: I -+ C , Y E AC(I),

(i.e., absolutely continuous on I) such that

py'E AC(I) and y(t) satisfies (1) a.e. on I.

Let y E I.

Then a quadrature gives, for t E l , p(t)y' (t)

=

13 +

t

J

y(s)q(s)ds

Y

(py') (y).

where 13 a(t)

Jt a

Since q E L(I)

its indefinite integral

q(s)ds exists for t E l and a E AC(I).

Hence y

will be a solution of (1) if and only if y(t) satisfies a Stieltjes integro-differential equation of the form p(t)y' (t.)

=

13 +

Jt

y(s)da(s),

tEl,

(2)

y

where the integral may be interpreted, say, in the RiemannStieltjes sense. whenever

On the other hand (2) also has a meaning

a E BV(I)

is continuous on I.

(i.e., bounded variation on I) and y Hence equations of the form (2) may be

used to deal with differential equations (1). need not be continuous on I

Moreover a

(as long as we require a solution

of (2) to be continuous on I) and so (2) can be used to treat discrete problems, e.g., difference equations (or threeterm recurrence relations) as well as continuous problems

XI

as we have seen. 0 corresponds to

The

vet) non-decreasing and the

case of unrestricted ret) corresponds to V(t)E BV(I). In the former case the operator defined by the differential expression is formally symmetric (under suitable domain 2

restrictions) in the weighted Hilbert space L (I,dv).

In

the latter case the operator is J-symmetric in a Krein (Pontrjagin) space, since the measure induced by vet) is a signed measure.

XIV

Expressions of the form (5) were first considered by

w.

Feller [68J,[69J,[70J,[71J,[72J,[73J in the case when

aCt) :: constant on I, p(t) function on I, a

E

:: I, and

v a given non-decreasing

(cf., also Langer [41J).

The more general case

BV(I) was treated by 1.5. Kac [35J,[36J,[37J when v is

monotone, cf., [46,p.49J.

CHAPTER 1 INTRODUCTION: In this chapter we shall study the Sturmian theory of Stieltjes integro-differential equations; that is, equations of the form

p(t)y' (t)

c + It y(s)do(s) a

defined on a finite interval

I = [a, b)

and

(1. 0.0)

p,

0

are real

valued right-continuous functions of bounded variation on and

p(t) > 0

I

there.

Historical Background: The comparison and separation theorems of Sturm comprise what we call the Sturmian theory.

Comparison theorems

for the scalar equation

(p(t)y' (t))' - q(t)y(t)

o

(1.0.1)

were first obtained by Sturm [58, p. 135] in his famous memoir of 1836.

In that paper Sturm considered the equations

o

(1. 0.2)

2

a on a finite interval and showed that if G < G 1, 2

(1.

a

< K

2

0.3)

K1 '

equality not holding everywhere on the interval,

then between any two zeros of some solution of (1.0.2) there is at least one zero of any solution of (1.0.3).

This is the

result usually known as the sturm-Comparison Theorem.

Sturm's

proof depended upon the introduction of a parameter in the coefficients which allowed him to pass continuously from to

K 2

and from

G 1

to

G 2,

K1

as the parameter was increased,

and then he studied the location of the zeros of the solutions as the parameter varied. valid for all

t

1,

t

2

E

It also depended upon the identity I ,

which can be obtained by an application of Green's theorem [ 13, p. 291].

It seems [58, p. 186] that Sturm carne to the conclusion of the comparison theorem by first having shown it true for the case of a three-term recurrence relation or second order difference equation though the latter result was not published. A discrete analog of the comparison theorem was published by Fort [21, p.

whose method of proof was, in essence, that

of Sturm applied to difference equations instead of differential equations.

3

In 1909 Picone [48, p. 18] gave by far the simplest proof of the comparison theorem in the continuous case.

He

made use of the formula t

[:l ( K yz' - K y'

z

2

1

Z)] t

2

1

(1. 0.5)

commonly known as the Picone Identity.

The use of (1.0.5)

allows an immediate proof of the Sturm Comparison Theorem [33, p. 226]. (cf., also

[74]).

One important extension of the comparison theorem was that of Leighton [42, p. 604] who interpreted the theorem in a variational setting: Q[y] y

associated with (1.0.2-3) acting on functions 1

C (a, b)

E

He made use of a "quadratic functional"

and

= y (b) = 0

y (a)

termed 'admissible').

For such

(such functions were

y,

(1. 0.6)

The main result was that if some non-trivial admissible function y

had the property that

Q[y] < 0

then every real solution

of (1.0.3) would have to vanish at some point in

(a, b)

Swanson [59, p. 3] weakened Leighton's condition

Q[y] < 0

Q[y]

0

for

y

t

0

reaching the same conclusion provided

the solutions were not constant multiples of

y.

to

4

The Sturm-Separation theorem states that the zeros of linearly independent solutions of, say, separate one another.

(1.0.2) interlace or

A similar result holds for three-term

recurrence relations and in fact a more general result is known in the latter case.

(See section 2) .

In section 1 we shall give an extension of the aforementioned "Leighton-Swanson Theorem" to the class of integral equations (1.0.0) and give, as corollaries, the corresponding continuous and discrete versions of the comparison theorem. In section 2 we give a proof of the Sturm Separation Theorem for (1.0.0) and give some applications to both differential and difference equations.

We conclude this

chapter with a study of the Green's function for boundary problems associated with

(1.0.0)

and its application to the

problem of finding an explicit representation for the solution of the non-homogeneous problem.

§l.l

(See section 3).

COMPARISON THEOREMS FOR STIELTJES INTEGRO-DIFFERENTIAL EQUATIONS: Let

p. (t) 1

(J.

1

(t)

i

=

1 , 2,

functions of bounded variation over p. (t) 1

> 0,

t

E

[a, b]

We assume that

[a , b]

i = 1 , 2,

functions are right-continuous on

be real valued

and that all four

[a, b]

with each possess-

ing a finite number of discontinuities there. simplicity only.

(This is for

In the following chapters this hypothesis

can be omitted, in most theorems, without affecting the conclusions.)

We will, in general, assume that all these

5

functions are continous at lim a(t)

exists as

t

+

a, b ,

and if

b

00

then

00

Consider the equations

P1 (t)u' (t)

c + It u(s)da (s) 1 a

P2 (t)v' (t)

c '

+

(1.1.0)

C

(1.1.1)

v (s) do 2 (s)

where by a solution of (1.1.0), say, we mean a function u(t)

E

AC[a, b]

each point

t

E

with

P1(t)u'(t)

E

BV(a,b)

satisfying (1.1.0) at

[a, b]

Associated with the pair (1.1.0-1) is the quadratic

Q[u]

functional

{u

U

E

with domain

AC [a , b]

, P 2 U'

E

BV(a , b)

, u (a)

u(b)

= a} (1.1.2)

and where, for

u

E

D ' Q (1.1.3)

We can now state and prove an extension of the LeightonSwanson result.

THEOREM 1.1.0: Let

i

=1

, 2 ,

be defined as above and let

6 U

E

D , Q

u pO,

be such that

QIu l

(1.1.4)

< 0 •

Then every solution of (1.1.1) which is not a constant multiple of

u(t)

vanishes at least once in

Proof:

Assume, on the contrary, that

(a, b)

and let

a < s < t < b.

Then

v

(a, b)



does not vanish in P2v'/v

E

BV b) 10 c(a,

and so

(1.1.5)

exists.

Case 1:

v(a)

0,

v(b)

For

t

Js

0 •

satisfying (1.1.4),

u

dl-VI)

2 (P2 v-

(1.1.6)

(1.1.7)

where in passing from (1.1.6) to (1.1.7) we used the equation (1.1.1).

Integrating (1.1.5) by parts we find that

7

v'u'u

(1.1.8)

v

Combining (1.1. 7),

(1.1. 8) and adding

to both sides we obtain,

[

P2

u

V'

v

2] t +

Jt

s

s

P u' 2 dt - 2 2

[p

+

v' 2 t

Jt s

P v'u' 2

J: p,{U' _ t t

[P2 V' Js

+

v

'2

(1.1.9) for

a < s < t < b . Hence if we let

obtain, since

v(a)

,

s

a +0

-+

t

-+

b - 0

in (1. 1. 9) we

a ,

v(b)

(1.1.10)

Q[u]

The hypothesis on we must have [a , b l

u

implies

a

or that

which we excluded.

Q[u] u

=

0

but since

is a multiple of

v

1- a , v

This contradiction shows that

on v

8

must vanish at least once in

Case 2:

= v(b) =

v(a)

(a, b)

0 •

To settle this case it suffices to

that in

(1.1.9) , 2

lim t+b-O

u (t)P2 (t)v' (t) v(t)

o

(1.1.11)

o .

(1.1.12)

and 2

lim s+a+O

u (s)P2 (s)v' (s) v(s)

It is possible to show that solutions to the initial value problem (1.1.1),

v(a)

=

c

See Appendix I and [3, p. 341]. v' (a)

0 •

=

P2(a)v' (a)

1,

Thus since

c

are unique:

2

v(a)

0,

(The prime here usually represents a right-

derivative which is an ordinary (two-sided) derivative if is continuous at the point in question.) since

o

v(b)

P2 (b ) v ' (b)

O.

Similarly [3, p. 348],

Hence

2

lim s+a+O

u (s)P2 (s)v' (s) v(s)

P2 (a) v

provided the latter limit exists.

02

I

(a)

u

2

(s)

lim V""T""S) s+a+O

The hypothesis on

(1.1.13)

02

implies that it is continuous in some right-neighborhood of Thus

P2(t)v' (t)

Similarly

P2(t)

is continuous in such a neighborhood. is continuous in some, possibly different,

right-neighborhood of

a.

Hence

v' (t)

is continuous (i.e.

a.

9

is an ordinary derivative) in some right-neighborhood (a,a+o)

o

> 0 •

In the same way it can be shown that ordinary derivative in (a, a + n) 2(t))' (a, a+n) , (u = 2u(t)u'(t)

n > O. Since

u' (t)

is an

Thus in u, v

E

AC [a , b] ,

we can apply L'H;pital's theorem to the limit in the right of (1.1.13) to obtain

u

2

(s)

lim s-+a+O

lim V"TS) s-+a+O

2u(s)u ' (s) v' (s)

o since, as we saw above,

v' (a)

7

O.

Hence the limit (1.1.12)

exists and is zero. Similarly it can be shown that (1.1.11) holds. Combining (1.1.11),

(1.1.12) and letting

s-+a+O,

t-+b-O

in (1.1.9) we obtain (1.1.10) again and thus derive a contradiction.

Case 3:

v (a) = 0,

v (b)

7

0

or

v (a)

7

0,

v (b) = 0 .

This case is easily disposed of as it is simply a combination of Cases 1 and 2 leading to (1.1.10) via (1.1.9) and (1.1.11-12).

This proves the theorem.

Associated with (1.1.0) is the quadratic functional Q' [u]

with domain

10

{u:

u

AC [a , b]

E

, P 1u

I

E

BV(a , b)

, u (a )

u(b)

= o}

(1.1.14) and (1.1.15)

Q' [u]

COROLLARY 1.1.0:

Let u(a)

u(b)

u

=

(Swanson [59, p. 4], Leighton [42, p. 605, Cor. 1]). be a non-trivial solution of (1.1.0) with

0

Then every solution constant multiple of

u

v(t)

of (1.1.1) which is not a

must vanish at least once in

(a, b)

provided

o .

Proof:

Let

u

be a solution of (1.1.0),

u(a)

(1.1.16)

u(b)

o .

Then

[up 1u

I ]

ba _

I

b

a

p 1u I 2 d t

(1.1.17)

Using the equation (1.1.0) in the left-side of (1.1.17) we find that Q' [u]

[uPl u' ]

o .

b

a

(1.1.18)

11

(1.1.16) now says that

Q' [u] - Q[u] > 0

or, because of

(1.1.18) , Q [u]

Since

u

(1.1.19)

< 0 •

is not a constant multiple of

applies and hence

v(t)

v,

Theorem 1.1.0

vanishes at least once in

(a, b)

.

Swanson's extension [59, p. 4] of Leighton's Theorem [42] is obtained by setting

t

0. (t) 1

[a, b)

E

,

i

=1

, 2 ,

(1.1.20)

in (1.1.0-1) and in (1.1.16).

COROLLARY 1.1.1:

(Sturm Comparison Theorem)

Let

q.EC[a,b],

i

1

=1

, 2

and suppose that

If

and

uta)

for which

=

=

u(b)

v(c)

=

0

0,

o

(1.1.21)

o

(1.1.22)

then there is at least one

whenever

v

c E (a, b)

is a solution of (1.1.22)

which is not a constant multiple of

u

12

Proof:

Let

0. (t)

be defined as in (1.1.20).

l

follows from Corollary 1.1.0 on account that is non-decreasing on

[a, b]

The result now

°1 (t)

- 02(t)

by the above hypothesis.

We now interpret these results for a three-term recurrence relation.

Let

(1.1.23)

b

be a fixed partition of the interval

[a , b]

and let

c

, c m- 1 be a given positive real sequence. -1 ' Co ' c 1 ' Let b ' b ' , b m- 1 be an arbitrary real sequence and o 1 define a function p(t) by setting on [a , b l

p(t)

for

c n- 1 (nt - tn- 1)

n= 0,1,2, ... , m.

if

Then

(1.1.24)

t E [ t n- 1 , tn)

p(t)

is a positive right-

continuous function of bounded variation with jumps, if any, at the

{t.} . l Now define

o(t)

on

[a, b]

by requiring that it be

a right-continuous step-function with jumps at the

{t

i}

of

magnitude o(t ) - o(t n n where With

0)

-b

n

(1.1.25)

n = 0 , 1, ..• , m-l . p(t)

, o(t)

as defined above consider (1.0.0).

On

13

[a, to)'

=

o(t)

It

constant, hence,

Ydo _ 0

(1.1.26)

a

and so (1.1.0) implies that

p(t)y' (t) = c

But

p(t)

=

p(a)

on

(1.1.27)

p(a)y' (a)

[a, to)

because

p(t)

is also a step-

function, hence (1.1.27) implies that

y' (t)

In fact, letting

y(t

(1.1.28)

y' (a)

n)

yn'

n = ­1 , 0 , 1 , ... ,m,

then

(1.1. 27) gives

y' (t)

(1.1.29) Hence p(t)y' (t)

p(a)y' (a)

(1.1.30)

14

Now let

t

[t

E

n_ 1,

from (1.1.30) that

t

,

n)

1

p(a)y' (a) +

p(a)y' (a) +

p(a)y' (a) +

p(a)y' (a) +

since

a

is constant on

constant on there,

[t

y' (t)

m.

When

p(t)y' (t) = c_ (Yo - y-1) 1

Thus

p(t)y'(t)

n

r a

n = 0, for

t

we know [a, to) .

E

yda

n-1

t.

L f

i=O

n-1

L

i=O n-1

L

i=O

yda +

l

t i- 1 t.+O

r. l

rt

Jt

yda n-1

yda + ft yda t n_ 1+O

y(t.) (a(t.) - a(t. - 0)) + 0 l

l

l

[t

is n- l ' t n ) . Hence p(t)y' (t) and since p(t) satisfies (1.1.24)

n- l ' tn) is also constant so that

y' (t)

Yn - Yn-1 t - t n n-1

t

[t n- l ' t n ) .

E

Consequently, p(t)y'

(t)

t

E

[t

n- l ' t n ) . (1.1.31)

This is true for each

n

in the range considered.

(1.1.0) gives

If

15

p(t -O)y'(t -0) + It ydo n n t -0

p(t)y' (t)

n

+0

t

p(t

n

-O)y'(t -0) + n

It

ydo

n n

-0

c n- 1 (Yn - Yn-1) + y(t n) (o(t n) - o(t n -

0») (1.1.32) (1.1.33)

where we have used (1.1.31) and (1.1.25) in obtaining (1.1.32), (1.1.33) respectively. By (1.1.31) we find that t

E

[tn' t

n+ 1)



p(t)y' (t) = c n(Yn+1 - Yn)

if

Combining this with (1.1.33) we obtain

(1.1.34) or c y + c y n n+1 n-l n-1

(c + n

C

n-l

-

b )Y n n

o

(1.1.35)

which is equivalent to

o where

(1.1.36)

represents the forward difference operator,

n = y n+1 - y n· Summarizing then, we see that when

p(t)

, o(t)

are

defined as in (1.1.24-5) respectively, the Stie1tjes integro-

16

differential equation (1.1.0) has solutions which are polygonal curves whose "vertices" are the points the

(tn' Yn)

and

satisfy the three-term recurrence relation (1.1.35)

or the second-order difference equation (1.1.36) for n

=0

, 1 , 2 , ••• , m-l .

An argument similar to the one above shows that if, instead of (1.1.25), we require

a(t)-a(t-O) n n then (1.1.0), with the same

(1.1.37)

p

as in (1.1.24), will give rise

to the recurrence relation

c n y n+l + where

C

n-l Yn-l - b n Yn

o

(1.1.38)

n = 0 , 1 , 2 , ••• , m-l .

The initial conditions

y(a)

=

a

p(a)y' (a)

(1.1.39) =

S

(1.1.40)

associated with (1.1.0) become, in the case of a recurrence relation, =

(1.1.41)

a

s

(1.1.42)

17

on account of (1.1.30). A fundamental solution, i.e. one in which

a

=

0,

B

1,

will then become

o

(1.1.43)

(1.1.44)

in the case of a recurrence relation (1.1.38).

(In this

respect see [3, p. 97]). With the

t

n

defined as in (1.1.23) we suppose given

two arbitrary real finite sequences and two positive sequences Let

C5. 1

(t)

,

i = 1 , 2,

c

b

n

qn' n = 0, 1, ... , m-l n, n = -1 , 0 , 1 , ... , m-l

be step-functions on

[a, b]

with

saltus b

n

(1.1.45)

(1.1.46) n

=0

Let

, 1 , •.• , m-l .

p. (t) 1

,

i

=1

, 2,

be defined by

c n- l(t n - t n- 1)

t

E

[t n- l ' t n )

(1.1.47)

r n- l(t n - t n- 1)

t

E

[t n- l ' t n )

(1.1.48)

18

where

o,

n

i=1,2,

1

...

,

, m-l .

Then the

and, along with the

0. (t) J.

for ,

continuous and of bounded variation on

i=1,2,

are right

[a, b]

Consider now (1.1.0-1) with the above choice of 0i' points

i=1,2. t.

J.

Pi'

The solutions of (1.1.0-1) evaluated at the

will then satisfy the recurrence relations

c n u n +1 + c n- 1 u n- 1 -

(c n +

n- 1 + b n ) u n

0

r n v n +1 + r n- 1 v n- 1 -

(r n + r n- 1 + q n ) v n

0

C

and

where

n = 0 , 1 , '"

, m-l

respectively.

The latter are

equivalent to

i'l (c n- li'lu n- 1) - b n u n

o

(1.1.49)

i'l (r n- li'lv n- 1) - q n v n

o

(1.1.50)

n = 0 , 1, ... , m-l.

We can now state a discrete analog of

the Sturm comparison theorem one form of which was proven by Fort [21, p.

] .

COROLLARY 1.1.2: > 0 and b n n equality not holding for every n

Let

c

n

> r

i'l(c n- li'lu n- l ) - b n u n

for

n=O,l,

u

o

m

=

0

. .. , m-l , and

(1.1.51)

19

then there is at least one node of

o

lI(r n- lllv n- 1) - q n v n in

(a,

(1.1.52)

b)

REMARK: We note that the condition lent to

u(a)

u(b)

=

0

when

u -1

u

=

um

=

0

is equiva-

is considered a solution

of (1.1.0). By a node we mean a point on the abscissa where the "polygonal curve" defined by the finite sequence

v

crosses

n

the axis.

Proof:

The condition

implies that

c

along with (1.1.47-8)

> r > 0 n = n

P1 (t) > P2 (t)

> 0

.

Moreover, since

b

n

> q = n

we find from (1.1.45-6) that

(J

Since

1

(t

n - 0) -

°2 (t n -

(1.1.53) implies that

for

t

E

[a, b]

°1 (t)

-

O

(1.1.53)

for each

are step-functions on

n,

0) •

2

(t )

is non-decreasing

This, along with the above Remark, shows

that Corollary 1.1.0 is applicable and hence the equation (1.1.1) has at least one zero in (a, b) to the required conclusion.

which is equivalent

20

Note:

In general, a comparison theorem for equations of the

form

c n Yn+l +c n-l Yn-l -b n Yn

under the assumptions For example let r

n

that

q

= c/2

n c

n

> r

c

n

c

= c > 0

for each b

n

in this case,

n

> r

n

> qn

(1.1.54)

o

(1.1.55)

for all n=O,l

n

,

,

and

...

, m-l

b

n

,

= 3c

We see then

but a simple computation shows that,

has no nodes eventually while (1.1.55)

will have nodes for large

§1.2

(1.1.54)

is not available.

n

,

n

o

n.

SEPARATION THEOREMS: In this section we prove the classical Sturm separation

theorem, on the separation of zeros of linearly independent solutions, as a consequence of the results in section 1.

In

the case of finite differences this result was also probably known to Sturm [58, p. 186] as one can gather from the remarks at the end of his memoir. If

c

n

> 0

for all

n,

then the nodes of solutions

of

c n Yn+l +c n-l Yn-l -b n Yn

o

(1. 2.0)

21

separate one another if these are linearly independent.

(The

proof of this result will follow below.) The Sturm separation theorem is not valid in general in the case of a general three-term recurrence relation

o .

(1. 2 .1)

Bacher [6, p. 176] points out that the separation property for solutions of (1.2.1) holds if

P

for all

n

n

R

n

> 0

(1. 2.2)

in the range considered.

false, in general, if (1.2.2) fails. an example the case where

P

n

= 1 ,

The result is however He gives [6, p. 177] as

Q = R =-1 n n

for all n.

The nodes of the linearly independent solutions corresponding to the initial values Yo

=

6

y-l = 0,

Yo = 1

and

y-l = -10 ,

do not separate one another. One proof of the separation property of (1.2.1) under

the hypothesis (1.2.2) was given by Moulton [45, p. 137].

We

note that the condition (1.2.2) is the analog of the condition p(t) > 0

for the equation

p(t)y" + q(t)y' + r(t)y

If

p(t) > 0

o .

(1.2.3)

then the zeros of linearly independent solutions

22

of (1.2.3) separate one another.

(One way of seeing this is

that (1.2.3) can then be transformed into an equation of the form (P(t)y')' + Q(t)y

where

P(t) > 0

o

(1. 2.4)

and the result follows from the separation

property of the zeros of (1.2.4).)

THEOREM 1. 2 . 0 : The zeros of linearly independent solutions of

c + It y(s)dcr(s) a

p(t)y' (t)

(1.2.5)

separate one another.

Proof:

(1.2.5) has two linearly independent solutions

which generate the solution space [3, p. 348]. and to, say,

u

find that

If we now set

in (1.1.16) we can apply Corollary 1.1.0 when

v

u, v

u

vanishes at two consecutive points to

must vanish in between since

stant multiple of

v

is not a con-

u.

In particular if

o

E

C' (a , b)

the classical Sturm separation theorem.

we immediately obtain

23 COROLLARY 1.2.0: If

C5

C I (a , b)

E

and

C5 '

(t)

= q (t)

t

E

[a, b)

then the zeros of linearly independent solutions of

(p(t)y')

I

-

q(t)y

o

(1.2.6)

separate each other.

Porter [49, p. 55] showed that two linearly independent solutions of (1.2.0) generate the solution space and considered the limiting process which takes a difference equation to a differential equation. Defining

C5,

P

as in (1.1.24-25) we obtain the discrete

analog

COROLLARY 1. 2 . 1 : If n

c

n

> 0,

= 0 , 1, ••. , m-l

n

= -1

, 0 , ... , m-l

and

b

n

is any sequence and

n-

n- 1)

- b n Yn

o

(1. 2.7)

then the nodes of linearly independent solutions separate one another.

As an application of Corollary 1.2.1 to the recurrence relation (1.2.1) we state the following [45, p. 137].

24 COROLLARY 1.2.2: Let

P

Q ,R n n

n,

be real finite sequences and

o for

n

=0

(1.2.8)

, 1, ... ,m-l .

If

P

> 0

R

n n

n

=0

, 1 , ••• , m-l

(1.2.9)

then the nodes of linearly independent solutions of (1.2.8) separate each other.

Proof:

The idea is to show that (1.2.8) under the hypothesis

(1.2.9) can be brought into the form (1.2.7) after which we simply apply Corollary 1.2.1. Let

c_

l

> 0

P

and consider the recurrence relation

n

n

R c n- l n

(1.2.9) implies that

c

n

> 0

=0

for

, 1 , ••• ,m-l .

n

=

0 , 1 , ••• , m-l

(1.2.10)

since

If we now set

(1.2.11)

for

n

= 0 , 1, ... ,m-l,

then a simple computation shows that

25

with the substitutions (1.2.10-11),

(1.2.7) reduces to the

three-term recurrence relation (1.2.8).

Hence the result

follows:

§1.3.

The GREEN'S FUNCTION: In Appendix I to this

work

we have shown the exis-

tence of a Green's function for the inhomogeneous problem

lji(t)

CI.

+ 13

Ja t

1 p

+

Jt

IS

1

P (s )

a

ljido ds + Jt f p

a

a

o

(1.3.0)

(1.3.1)

where

{M .. lji ( j -1)

j=l

1J

(a)

+ N .. P (b) lji ( j 1J

-1) ( b) }

i

=

1 , 2 ,

(1.3.2)

and the

M.. , N .. 1J

1J

are real constants, under the hypothesis

that the homogeneous problem (with

f

conditions (1.3.1) is incompatible.

=0

)

a.nd the boundary

(By this we mean that the

homogeneous equation with homogeneous boundary conditions has only the zero solution.)

If

f

=

0

in (1.3.0) then the

resulting integral equation is of the form (1.0.0). If ous on with

0

[a, b) o'

=

q.

E

C' (a, b)

and

p(t)

then (1.3.0) with

is positive and continuf

=

0

reduces to (1.2.6)

In this case the "derivative" appearing in

(1.0.0) is continuous and the Green's function reduces to the

26

usual one.

(See Appendix I, p. 278 .)

On the other hand if then (1.3.0) with

o

f

recurrence relations.

p(t)

,

a(t)

are step-functions

can be made to include three-term

In this case and, more generally, for

difference equations of higher order the Green's function seems to have been first constructed by Bacher [5, p. 83]. Another treatment was given by Atkinson [3, p. 148]. We showed in Appendix I that if (1.3.0-1) with is incompatible then the unique solution of (1.3.0-1)

f

0

is

given by

e

for t

E

x

E

[a, b]

[a, b]

and

same points where

In the particular case when f(t)

p(t) = 1 ,

is a step-function with jumps at the

a(t)

has its jumps and if we denote by

f(t.)

f.

- f(t. - 0)

l

l

where, as usual, the

(1. 3.3)

G(x, t)df(t)

1j!(x)

t. l

(1.3.4)

l

represent the jump points of

f

,

then a simple computation shows that

t a

m-l

L

i=O

G(t

n

G(t

, t)df(t)

n,

til • (f(t

i)

- f(t

i

- 0)) (1.3.5)

27

and if we write

o

G . _ G(t , t.) nl. n l.

< n,

i < m- 1

we find

that m.... l ljin

This

ljin

I

i=O

G .f. nl. l.

(1.3.6)

then represents the solution to the corresponding

inhomogeneous difference boundary problem.

Usually (1.3.6) is

derived directly using methods of finite differences. for example [3, p. 149] and [5, p. 84].)

(See

For further details

see Appendix I, section I.4. We note that when

p(t)

, a(t)

functions of bounded variation on

are continuous

[a, b]

then the derivative

appearing in (1.0.0) is continuous everywhere and so, from Appendix I, the discontinuity in the first derivative of the Green's function is given by

G (t + 0 , t) - G (t - 0 , t) x

x

1 p(t)

(1.3.7)

which is the usual measure of discontinuity of the Green's function associated with a second-order linear differential equation of the form (1.2.6).

CHAPTER 2 INTRODUCTION: There is a very extensive literature dealing with the subject of oscillation and non-oscillation of real second order differential equations on a half-axis (see, for example, [59]).

On the other hand there is little known about

establishing criteria for the oscillatory and non-oscillatory behaviour of solutions of difference equations.

In the

particular case of three-term recurrence relations some results can be found in [23, pp. 126-128] and more recently in [32, p. 425]. [12],

Other results are more or less scattered:

[21],

[20]. In this chapter we shall be concerned with obtaining

some oscillation and non-oscillation criteria for linear and non-linear Stieltjes integral equations on a half-axis.

It

will be noted that if one makes an hypothesis on the integral of the potential

q

in

y" - q(t)y

o

t

E

[a, (0)

(2.0.0)

which will guarantee the existence of oscillatory or nonoscillatory solutions, then a certain discrete analog will

29

exist for a three-term recurrence relation. In section 1 we give some non-oscillation criteria for Stieltjes integral equations and their applications to second order difference equations.

In section 2 we give some results

on the oscillatory behaviour of solutions and in section 3 we extend a result of Butler [8, p. 75] and state a necessary and sufficient condition which guarantees that all continuable solutions of a non-linear equation are oscillatory.

As a

corollary we shall obtain the discrete analog of Atkinson's theorem [2, p. 643].

Various examples are included which

should help visualize the theorems stated.

§2.l

NON-OSCILLATION CRITERIA FOR LINEAR VOLTERRA-STIELTJES INTEGRAL EQUATIONS: In the following, we shall usually be considering

equations of the form

y' (t)

where

0

c + It y(s)do(s) a

t

E

[a, 00)

(2.1. 0)

is a right-continuous function locally of bounded

variation on

[a, 00)

Because of the applications we shall

assume, in addition, that the number of discontinuities of remains finite in finite intervals.

0

The theorems proved here

can also be extended to equations of the form

p(t)y' (t)

c +

r a

y(s)do(s)

(2.1.1)

30

in the case when

p(t) > 0 ,

(2.1.2)

00

p

satisfying the usual conditions stated in Chapter 1.

every equation of the form (2.1.1), where

p

For

satisfies

(2.1.2), can be transformed into an equation of the form (2.1.0) by the change of independent variable

t t---»

which will take

[a ,00)

T

(t)

into

=

It 1. a

(2.1. 3)

p

(See Appendix I,

[0,00).

equation (1.3.14).)

DEFINITION 2.1.1: A solution of (2.1.0) is said to be oscillatory if it has, to the right of

a,

an infinite number of zeros and is

non-oscillatory if there is some

zeros when

t

to

E

m

such that it has no

to

From the Sturm separation theorem, Theorem 1.2.0, we see that if one solution is oscillatory (non-oscillatory) then all solutions are oscillatory (non-oscillatory). Equation (2.1.0) is said to be oscillatory (nonoscillatory) if all of its solutions are oscillatory (nonoscillatory) .

Unless otherwise stated we shall, in the following,

31

assume that

o(t)

,

appearing in (2.1.0), has a limit at

00

Le. lim

(2.1.4)

0 (t)

t-+ oo

exists and is finite.

0(00)

Denoting this limit by

assume it is zero (for if we let has the same properties as

,(t)

(2.1.0) remains unchanged if

o(t) - 0(00)

=

,(00)

and

0

=

0

then

,

Moreover,

is replaced by

0

we can

,).

The first result is an extension of a well-known theorem of Hille [31, p. 243] which relates the non-oscillatory behaviour of (2.1.0)

to the existence of solutions of a

certain non-linear integral equation. THEOREH 2.1.1: Let

0

be right-continuous and locally of bounded

0(00)

variation satisfying (2.1.4) with

and sufficient condition for (2.1.0)

= o.

Then a necessary

to be non-oscillatory is

that the integral equation

v(t)

=

o(t) +

C

v 2(s)ds

have a solution, for sufficiently large integrable at infinity

Proof:

(cf.,

(2.1.5a)

t,

which is square

[80]).

To show that the condition is sufficient assume that

(2.1.4) has a solution then implies that

v(t)

V E

some

(2.1.4)

is right-continuous, locally of

32

bounded variation and

v(oo)

O.

Put y(t)

exp Jt v(s)ds .

(2.1.5b)

a

Then

y(t)

is locally absolutely continuous and so

=

y' (t)

v(t)exp Jt to

v(s)ds

(2.1. 6)

everywhere, as a two-sided derivative, except possibly the jump points of Letting

v(t)

h > 0,

t

which are the same as those of

o(t)

arbitrary,

e x p [(t+h Y (t+h) - Y (t)

y(t) .

h

v (s)

dS) (2.1. 7)

{

h

Now

t +h Jt

for each

h

>

0,

v-I

)

fixed

use Theorem H of Appendix

(2.1. 8)

t I

Hence we can let

+

0+

and

to find that

I It+h lim 11 t v(s) ds

h+O+

h

v (t)

(2.1. 9)

33

while the other terms are zero by virtue of (2.1.9) and the continuity of the integrals. Hence letting

h

+

0+

in

we obtain, from above,

y' (t)

(2.1.10)

y(t)v(t)

where the derivative is in general understood as a rightderivative which is locally of bounded variation.

r r

Thus if

dy' (s)

to

d (y (s) v (s) )

to

I

t

v dy +

Y dv

to

to

where we have

I

t

by equation (2.1. 10) ,

(2.1.11)

Theorem K of Appendix I now implies that the first integral in (2.1.11) vanishes for all

t

and hence

34

y' (t)

so that

y(t)

equation. for

t

> =

is a positive solution of the above integral

This implies that (2.1.0) has a positive solution t

0

and hence is non-oscillatory.

To prove the necessity we suppose that (2.1.0) has a non-oscillatory solution positive for For

y(t)

t > to . t > to

we set

v(t)

Then

v(t)

which we can suppose is

z.'.J!.L

(2.1.12)

y(t)

is locally of bounded variation on

[to' 00)

is right-continuous.

Hence, for

t > to '

r

to

y(lS) dy' (s) -

r [?) 0

2

ds

and

35

Hence

v(t)

for

t

to'

(2.1.13)

Since

o(t)

has a limit at

that the same must be true of Suppose, if possible, that square-integrable at

v(t)

v(oo)

y' (t) < 0

there is a

t

for

(2.1.13)

shows

. a

O.

Then

v

cannot be

and so (2.1.13) implies that

00

lim v(t) t-reo

Hence

=

00

t

> =

t

1

(2.1.14)

_00

because of (2.1.12).

Moreover

such that when

2

(2.1.15)

If we let

t3

=

(2.1.13) with

max{t o ' t 1 ' t 2 }

t , to

then using (2.1.15) in

replaced by

T, t

3

respectively, we

obtain

v(T) < -1 +

JTt 3

whenever

T > t

=

3

.

y(s)

(2.1.16)

We now use Gronwall's inequality [9, p. 37,

Exercise 1] in (2.1.16) to obtain

36

v(T)

« -1 -

I:, IY;(;:/ I {JJy; I exp

-exp JT t

}dS

y(s) 3

Thus, by (2.1.12),

(2.1.17)

for all

T

t

(2.1.17) implies that



3

positive which is a contradiction.

y(t)

cannot remain

v(oo) = O.

Hence

We can

now rewrite (2.1.13) as

v(T) - v(t)

where

T

that

v(t)

t

t

o (T) -

Now letting

3.

satisfies (2.1.4) for

0 (t)

T

-

J

v

2

(2.1.18)

in (2.1.18) we find

00

t

T

t

t

3.

This completes

the proof.

THEOREM 2.1.2: Let

0

1 '

O

2

bounded variation on 0;(00)=0,

i=1,2

be right-continuous functions locally of [a, 00)

satisfying (2.1.4) with

37

Assume that

(2.1.19) If

(2.1.20)

has a solution for

t > to

then

(2.1.21)

v(t)

has a solution for

Proof:

t > to .

We shall make use of the Schauder fixed point theorem

(Appendix II, Theorem 2.1.1).

With the Banach space

and the usual norm we consider the subset

x where

{v

E

vi (t)

For

L2

is as in (2.1.20). v

E

X

we define an operator

a E [0, 1]

and

T

on

X

by

(2.1.23)

(Tv) (t)

If

(2.1.22)

(to ' co)

x , Y EX,

38

I ax

+

(1 - a) y

I

< a Ix

I

+

(1 - a)

Iy I

(2.1.24) and hence

X

For

is convex. v EX,

I (Tv)

(t)

I


to.

- (Tx) (t) [2


t

Tx

n

- Tx I 2

Hence

0

To show that of Appendix II. ITxl

vi

T

-+

0

as

n

-+

(2.1.31)

00

is continuous.

TX

is compact we use Corollary 11.1.2

(11.1.4) is satisfied since if

x EX,

and so

(2.1.32)

= {t: to

E A

choose

sufficiently large so that

A

A

t < oo}

If we let

then given

>

0,

we

(2.1.33)

This will then imply that

(2.1.34)

for all

x E X

by virtue of (2.1.32).

This proves (11.1.5).

TO prove (11.1.6-7) we need some additional information. since

is a solution of (2.1.20),

2

vi

(t)

>

1° 1 (t) I 2

(2.1.35)

41

and so

01

L

E

2

(to' (0)



By the same argument,

{C

(t)

and so

(2.1.36)

The following theorem [24, p. If

f

E

LP[t

o'

(0)

,

P > 1,

Ilf(x+h) -f(x)ll

Since

01

L 2 [t

E

o

L

2

, (0)

[to ' (0)

] will also be useful. then

p

as

0

-+

h

0 .

-+

we have from (2.1.19)

(2.1.37)

that

and thus

110 2(t+h) -0 2(t)11

-+

0

h

as

0

-+

(2.1.38)

on account of (2.1.37) . Similarly if we set

V(t)

C

2

Vi

then I!V(t+h) -V(t)11

-+

0

as

h

-+

0

(2.1.39)

42

because of (2.1.36). Thus if

x EX,

>

0

(2.1.40)

if

Ihl < 8 ,

by the continuity of the integral.

This proves

(11.1.6) •

For

x EX,

> 0

II (Tx) (t+h) - (Tx) (t)11

Ilo 2(t+h) -02(t) +

From (2.1.38) we can choose

h

+

II

Jt+h

x

2

ds II

2

t

< Ilo 2(t+h) -o2(t)11

t

Jt+h x

so that if

(2.1.41)

II • jhl < 8 1

then

(2.1.42)

Similarly there is a

8

such that whenever

> 0

2

Ihl < 8

2

(2.1.43)

Ilv(t+h) -V(t)11 < Thus

II

Jt

t+h

x

2

11


to

and

(2.1.45) If v

has a solution for

1

(t)

t > to

v (t)

has a solution for

Proof:

01 (t)

+

"" v 12 ds ft

then

= ± o 2 (t)

+

f:

v

2

ds

t > to .

This follows immediately from the theorem.

THEOREM 2.1.3: With

(2.1.46)

as above and

(2.1.47)

44 (2.1.48)

suppose that y' (t)

is non-oscillatory.

z'

(2.1.49)

Then

(2.1.50)

(t)

is non-oscillatory.

Proof:

This is immediate from Corollary 2.1.2 and Theorem

2.1.1.

THEOREM 2.1. 4 :

Let

0(t)

satisfy the conditions of Theorem 2.1.1.

If

tl0 (t) I
to > 0

(2.1.51)

then (2.1.0) is non-oscillatory.

Proof:

Let

01 (t)

Theorem 2.1.3.

=

1/4 t

and

02 (t)

==

0(t)

and apply

This is permissible since (2.1.49) is then

equivalent to y" +

1 4t

2

Y

o

(2.1.52)

45

which is a non-oscillatory Euler equation [59, p. 45].

The

result now follows.

COROLLARY 2.1.3: (2.1.0) is non-oscillatory if

lim sup tlo(t) t++ oo Proof: t

I


0

or

y(t) < 0

for

t;; t

1



If

y(t)

> 0 ,

46

Theorem 2.1.1 implies that (2.1.46) has a solution t ;;: t* for

maxi to ' t i}

vi (t)

Hence (2.1.47) has a solution

for

v(t)

t > t* (because of Corollary 2.1.2) which corresponds to

some non-oscillatory solution suppose

z(t) > 0

for

z(t)

t > t*

of (2.1.50).

We can

Since the proof of Theorem

2.1.2 guarantees that

Iv(t)

I

t > t*

< vi (t)

we can recover the non-oscillatory solutions

(2.1.56)

y, z

to find

that z (t) < Y (t)

If

z(t) < 0

for

other hand if

t;;: t*

y(t) < 0

t > t* .

(2.1.57)

the last line is clear. for

t;;: t

i

then

On the

-y(t) > 0

the above argument shows that there is some solution

and z(t)

such that z(t) < -y(t)

t > t* .

(2.1.58)

This completes the proof.

THEOREM 2.1.6: Let

o(t)

satisfy the hypotheses of Theorem 2.1.1.

If (2.1.59)

47

then (2.1.0) is non-oscillatory.

Proof:

By Theorem 2.1.1 i t suffices to show that (2.1.4) has

a solution for sufficiently large

We shall again make

t

use of the Schauder fixed point theorem. Let

X

be a subset of

defined by

Iv (t)

X

-

0 (t) I

10 (t) I ,


to}

. (2.1.60)

For

v

X

E

we define a map

(Tv) (t)

If

a E [0, 1]

1

au

and

T

by

I:

o (t) +

Ia (u -

01

0)

I

< a u - 01

X

2

(2.1.61)

ds .

u, v EX,

+ (1 - a) v -

This shows that

v


0

in Theorem 2.1.7 and then state the

converse.

2.1. 8 :

THEOREM

Let With

O

o(t)

satisfy the hypotheses of Theorem 2.1.1.

defined as in (2.1.65) suppose that (2.1.64) is non-

2

oscillatory. Then (2.1.0) is non­oscillatory and for each non­trivial solution

z

of (2.1.64)

there is a solution

such that

o for

t > t

Proof: space

X

where

< Y (t)


to .

is no longer required to be

Equality in (2.1.84) is attained in the case of

the Euler equation (2.1.52).

THEOREM 2.1. 7A: Let

a(t)

satisfy the hypotheses of Theorem 2.1.6A

along with (2.1.84).

If

A(t) > 0

y" + 4A 2 (t)y

for large

t

then

o

(2.1.85)

is non-oscillatory.

Proof:

Refer to Theorem 2.1.7 with

o(t)

- A(t)



Whether (2.1.68) being non-oscillatory implies that (2.1.85) is, appears to be an open question [59, p. 93] which we shall discuss in section 2.2.

COROLLARY 2.1.4A: Let

A(t)

> 0

and suppose that (2.1.84) is satisfied.

Then (2.1.68) has a non-oscillatory solution

Iy (t.) I

< exp {2

r t"

A (s) ds } .

y(t)

such that,

(2.1.86)

59

Proof:

This follows from Corollary 2.1.4 with

o(t) - A(t)

.

THEOREM 2. 1 . 8A : Let

A(t)

be defined as in (2.1.68) and suppose that

z" + is non-oscillatory.

4A

2

(t)

o

z

(2.1.87)

Then

o

y" + a(t)y

(2.1.88)

is non-oscillatory and for each non-trivial solution (2.1.87) there is a solution

o for

t

Proof:

< y(t)

y(t)

t

sufficiently large, say,

of

of (2.1.88) such that

{r

< [z Lt.)

z(t)

t

IA(S) IdS}

(2.1.89)

1

t

1

.

This is an application of Theorem 2.1.8. The first part of the theorem is identical with a

theorem of Hartman and Wintner [27, p. 216] though the estimate (2.1.89) is stronger than the corresponding estimate in [27] where the absolute value sign about not appear.

A(t)

in (2.1.89) does

Thus the first part of Theorem 2.1.8 extends the

Hartman-Wintner result cited

above

to equations of the type

(2.1.0), while the second part extends the corresponding result only when

o(t) > 0

in Theorem 2.1.8.

60

2.1B

APPLICATIONS TO DIFFERENCE EQUATIONS: In this subsection we apply the theorems of section

2.1 to recurrence relations of the form

o

c y + c y + b y n-l n-l n n n n+l where

n

=

-1 , 0 , 1 , ... ,

(2.1.90)

is any given real

(b )

n

n = 0 , 1 , ••• •

sequence,

We shall assume, unless otherwise specified, that

(2.1.91)

00

be satisfied as an extra condition upon the We saw in Chapter 1 that if

o(t)

c

n

is a step-function

with jumps, at a fixed increasing sequence of points where

t -1

=

a

(t )

n

and (2.1.92)

n=O,l, ... , of magnitude

o(t) - o(t -0) n

for

n

=0

, 1 , 2 , •.. ,

n

-b

n

- c

n

- c

(2.1.93)

n-l

then (2.1.0) gives rise to solutions

of some "extended" recurrence relation in the sense that the resulting solution is a polygonal curve [a , 00)

y(t)

which has the property that if we write

defined on y

n

:: y(t ) n

61

then the sequence

(Yn)

is a solution to the three-term re-

currence relation (2.1.90) for

n

We note that, whenever for given sequences

m

> 0

b

,1, . . . •

o(t)

,

n

=0

is defined by (2.1.93)

then for

, m

o (t)

o (a)

-

L o

(b

n

+c

n

+c

n- 1)

This follows from (2.1.93) and the relation

(2.1.94)



o

(t

n

- 0)

= o(tn- 1) . THEOREM 2.l.1B: Let

o(t)

be defined as in (2.1.94) and assume that

exists and is zero.

0(00)

Then a necessary and sufficient

condition for (2.1.90) to be non-oscillatory is that

v(t)

where L2

o(t)

o (t) +

I:

v

2

ds

(2.1.95)

is given by (2.1.94), have a solution which is in

at infinity.

Proof:

This follows immediately from Theorem 2.1.1 and the

results of Chapter 1.

REMARK: A solution

(Yn)

of (2.1.90) is said to be oscillatory

62

if the sequence exhibits an infinite number of sign changes and non-oscillatory if, for constant sign.

n

N,

the sequence retains a

The discrete version of the Sturm separation

theorem shows that if a solution is oscillatory (nonoscillatory) then all solutions inherit the same property. Moreover the transition from (2.1.0) when

°

to (2.1.90), in the case

is given by (2.1.94), shows that a given solution of

(2.1.0) is oscillatory (non-oscillatory) if and only if the corresponding solution of (2.1.90) is oscillatory (nonoscillatory) . Thus, with

°

defined as in (2.1.94),

c + It y(s)do(s) a

y' (t.)

t

E

[a, (0)

(2.1.96)

is oscillatory (non-oscillatory) if and only if

o

c n y n+l + C n-l Yn-l + b n Yn

n=O,l, ..•

(2.1.97)

is oscillatory (non-oscillatory). The latter theorem thus gives the discrete version of Hille's theorem [31]. For given sequences functions

01

'

02

on

c

n

[a, (0)

> 0 ,

n

we define step

by setting

m I' L.

b

(b

n + c n + c n- 1)

(2.1.98)

63

m > 0,

and m

Io

02 (a)

(g

n

+

c

n

+c

n-

We recall that the (2.1.92) a fortiori, so that With

°1

'

°2

t

n

-->-

00

as

c

(2.1.99)

1)

also satisfy

n

n -->-

00

so defined we obtain the discrete

analogs of Theorem 2.1.2 and Corollary 2.1.2 denoted by Theorem 2.1.2B and Corollary 2.1.2B respectively.

Since the

latter two results can be stated in the same way as the former two, we shall omit them and it shall be understood that when we refer to either of Theorem 2.1.2B or its corollary we shall mean Theorem 2.1.2 or its corollary with 01

'

02

THEOREM

given by (2.1.98-99).

2.1. 3B : Let

c

n

>

0

and satisfy (2.1.91).

Suppose that

m lim I (b + c + c 1) n n nm-->-oo 0

(2.1.100)

m lim I (g +c +c 1) n n nm-->-oo 0

(2.1.101)

both exist and are finite

(so that the series need only be

conditionally convergent) . Suppose further that

64 00

L

m for

(c

m > mo'

n

+C

n-

1 +b )

>

n

I

I

m

(c n

+ C n- 1 + g n )

I

(2.1.102)

If

c n y n+l + c n-l y n-l + b n Yn

o

(2.1.103)

o

(2.1.104)

is non-oscillatory then

is non-oscillatory.

Proof:

Define

01

'

02

by (2.1.98),

(2.1.99) respectively.

(2.1.100-101) are then equivalent to requiring that both 01 (00)

,

exist and be finite.

02(00)

Since we can alter these o (00)

by an additive factor, we can assume that

1

= o 2 (00) = 0



This then implies that 00

01 (a)

L

(c n + C n- 1 + b n )

L

(c n + C n- 1 + g n )

0

o 2 (a)

Hence, for

t

E

0

(2.1.105)

.

(2.1.106)

[tm- l ' t m) , 00

o 1 (t)

L

m

(c n + C n- 1 + b n )

(2.1.107)

(c n + C n- 1 + g n )

(2.1.108)

00

o 2 (t)

L

m

65

Thus the requirement that (2.1.48) be satisfied for large

t

is equivalent to the requirement that (2.1.102) hold for large m.

From the remark we see that (2.1.49) must be non-

oscillatory.

Hence Theorem 2.1.3 applies and hence (2.1.50)

is non-oscillatory.

Consequently (2.1.104) is also non-

oscillatory and this completes the proof. The latter theorem is therefore the discrete analog of the Taam result [60].

Simultaneously i t provides an

extension of the discrete version of the theorem of wintner [63]

and Hille [31]

(see Theorem 2.1.3A).

Thus for example,

if > 0

n=O , 1 ,

...

(2.1.109)

gn > 0

n=O , 1 ,

...

(2.1.110)

b

n

and co

L

m

co

b

n

>

L

m

gn

o ,

m > m

(2.1.111)

then (2.1.104) is non-oscillatory if (2.1.103) is.

This would

be the formulation of the discrete analog of Hille's theorem [31] •

THEOREM 2.1. 4B:

c

Let sequence

(b )

n

n

> 0

and satisfy (2.1.91).

For a given

assume that (2.1.100) exists.

If

66

(2.1.112)

then (2.1.90) is non-oscillatory.

Proof:

We define

by (2.1.93).

G

Then, for

t

E

[tm- l ' t m) ,

we shall have

G

L

(t)

m

(c

(2.1.113)

n + C n- 1 + b n ) •

For (2.1.51) to hold for large

t

it is necessary that

(2.1.114)

for all

t t

we let

-+

m

t_

l

m

=

a

mo

=

0

.

m

is sufficiently large.

Thus

in (2.1.114) and use (2.1.92) to obtain

tm

{a +

for

when

[tm- l ' t m)

E

I

0

_1 c n- l

}IIm (cn +

C

n- 1 + b n )

I
0

t

E

(2.1.131) then the differential equation

z"

2

o

+ 40 (t)z

(2.1.132)

is non-oscillatory.

Proof:

This follows immediately from Theorem 2.1.7.

COROLLARY 2.1.4B: Let

c

n

, b

n

satisfy the hypotheses of Theorem 2.1.7B

and suppose that (2.1.131) holds. oscillatory solution

(Yn)

Then (2.1.90) has a non-

such that for

2t

t

Iy n I

< ex p {

n

0 (s)

dS}

n

N , (2.1.133)

73

where

Proof:

a

is as in (2.1.131).

Follows from Corollary 2.1.4.

THEOREM 2.1. 8B : Let

a(t)

be as in Theorem 2.1.7B except that

need not be non-negative.

Assume that

2

o

z"+4a(t)z is non-oscillatory.

a

(2.1.134)

Then

c n y n+l +

C

n-l Yn-l + b n Yn

o

is non-oscillatory and for each non-trivial solution (2.1.134) there is a solution

(Yn)

(2.1.135)

z

of

of (2.1.135) such that

t

where

Proof:

o

< yn
0

is fixed and

(2.2.26)

tA(t)

then (2.1.70) is oscillatory.

Proof:

This is a consequence of Theorem 2.2.1 where

a(t)

:: A(t) The first part of this theorem is due to Wintner [63, p. 260] and the second part follows almost immediately from this result (see [44, p. 131],

[63, p. 259]).

81

THEOREM 2.2.2A: Let

a(t)

and suppose that

satisfy the conditions of Theorem 2.1.6A A(t)

> 0 .

If

(

2 A (s)ds


0 ,

(2.2.28)

then (2.1.70) is oscillatory.

Proof:

This follows from Theorem 2.2.2. The above theorem is due to Opial [47, p. 309].

THEOREM 2. 2 . 3A : Let

a(t)

be continuous on

[a, 00)

and suppose that

( 2 . 1. 6 8 ) exi s ts . If

JOO a(s)ds a

then (2.1.70) is oscillatory.

00

(2.2.29)

82

Proof:

We let

= -

o(t)

Jt a(s)ds a

in Theorem 2.2.3.

The latter theorem was proven by Fite [19, p. 347] in the case when

a(t)

0

p. 115] for general

§2.2B

and was extended by wintner [61,

a(t)

APPLICATIONS TO DIFFERENCE EQUATIONS:

THEOREM 2.2.1B: Let the

c

n

, b

n

satisfy the hypotheses of Theorem

2.1.4B and assume further that

defined in the proof

G

m

m > m

of Theorem 2.1.6B, is non-negative for

= a

.

If

m

{ Ia

1

} mI 00

c n-l

(c n + C n- 1 + b n )
m a (2.2.31)

where

Proof:

> 0

is fixed, then (2.1.90) is oscillatory.

The first part is a consequence of Theorem 2.1.4B.

An argument similar to the one used in the proof of the latter theorem shows that (2.2.3) is equivalent to (2.2.31).

83

As a consequence of this we obtain in particular,

COROLLARY 2.2.1B: Let

(b)

be any sequence whose series is conditionally

n

convergen t and 00

(2.2.32) If 00

m

I

m

bn
my

>

>

for some

E >

0

if

m

C+l

m

-2

dx

y

m+l 1 4" +

x

(2.2.36)

E

is sufficiently large since

y >

1:-4

Consequently the above corollary implies that (2.2.35) is oscillatory. Using the discrete Euler equation (2.2.35) as a comparison equation we can deduce the following result. (b) n

is a positive sequence such that, for fixed

E

If

> 0 ,

(2.2.37)

then (2.1.116) is oscillatory.

[4"1+ ) E

then

bn

gn

0

For if we let

(n

+

1) -2

(2.2.38)

and Theorem 2.1.3B would lead immediately

to a contradiction if (2.1.116) was assumed non-oscillatory. Similarly it can be shown that if


0

(2.3.4)

n=1,2, ... ,

and continuous, in terms of an integral condi-

tion on the coefficient.

This has recently been generalized

[8 ] to equa tions y" + p(t)f(y)

where

p(t)

o

is unrestricted to sign and

into a "superlinear" equation.

(2.3.5)

f

turns (2.3.5)

The result which we shall

prove later on will give, in particular, a necessary and sufficient condition for the difference equation

2

6. Y n­ 1 + b n f (y n )

to be oscillatory.

o

As a corollary we shall obtain the dis-

crete analog of Atkinson's theorem [2], i.e. positive then, for

(2.3.6)

n=O,l, ...

If

k > 1 ,

o

n=O,l, ...

(b

n)

is

89

has a non-oscillatory solution if and only if 00

I

o

nb

n


0

for all

y

7-

o.

f'(y) > 0

and

oo

dt

1

f (t)

J b)

T

J

T

J

t

dt

< 00

f (t)


- 00


0

Y (t)

We let

g (t)

=

f

f

(t)

Y (t)

J

(2.3.12)

where the prime represents in general a right-derivative. Then

g{t)

variation on

shall be right-continuous and locally of bounded (to' 00)

An integration by parts shows that, for

to < t < T ,

91

I

T 1 t f(y(s») dy' (s)

T f'(y)(y,)2 t ds f(y)2

J

g(T) - g(t) +

(2.3.13)

where we have omitted the variables in the integrand for simplicity.

Moreover, an application of the integral equation

(2.3.1) shows that

T 1 It f (y)

dy I = -

ITt

(2.3.14)

de .

Hence combining (2.3.13-14) we find

=

g(t)

whenever

g(T) + er(T) - er(t) +

to < t < T.

I:

f'

ds

(2.3.15)

Our basic assumption leads us to two

cases: I)

lim sup T-+oo

II)

Case I: t

=>

t

0

(

lim sup ( T-+oo

P(s)ds

some

+00

P (s) ds
to

t > to.

(2.3.16)

(2.3.17)

(2.3.16) implies that the relation is valid for all .

If there is a sequence

T

n

t

00

such that

(2.3.18)

then for

n

sufficiently large we shall have, for

92 T

n

Tn

g(t)

g (T ) + 0 (T ) -

n

G (t)

n

+

f

f

I

(y) g

2 ds.

(2.3.19)

t

Hence

It

T

g(t)

so that letting lim sup

>

n

Tn

It

n do +

00

we obtain

f' (y)g

2

g(t)

ds

(2.3.20)

P(t)

Taking the

of both sides of the latter we obtain a contradiction

on account of

(a).

If no such sequence exists then we must have

g (t)

< 0

(2.3.21)

y' (t)

< 0

(2.3.22)

and so

(2.3.15) now implies

g(T) < g(t) -

Moreover (2.3.16) implies that large of some

t

I:

(2.3.23)

do .

P(t) > 0

for arbitrarily

(not necessarily all such) which shows the existence t2

t

1

such that

(2.3.24) if we assume that

P(t)
t

'

2

g(T) < g(t ) - -K < 0 •

(2.3.25)

2

Replacing

t

by

t

in (2.3.15) and using (2.3.24 - 25) we

2

obtain

+

g(T)

If we write t > t

=

2

r t

f' (y (t) )

¢(t)

fl (y) Iy' f (y)

g(T)

-K +

g ds .

(2.3.26)

2

Iy

1

fly (t)}

and so

I

(t)

r t

I

then

¢(t) > 0

for

(2.3.27)

¢(s)g(s)ds .

2

An application of the Gronwall inequality to (2.3.27) then gives

g(T) < -K

f (y (t

2)

)

f(y(T))

T > t2

(2.3.28)

T > t

(2.3.29)

Le. y' (T)


0 t

P (s) ds

I

such that

T > t

< Mt

We now proceed as in Case I.

(2.3.31)

.

If there is a sequence

such that (2.3.18) holds for large

n

T

n

t

00

we find from (2.3.20)

that g2 (t) >

y (t)

i)

Now either

ii)

,;

i)

Let

y(t)

>

s > 0

there is Y (t ) n cS

> 0,

(2.3.32)

(t)

t > t3 '

(t)

such that

n

or t

n

too,

4- 0

t > t

c = inf {r : (u)



3

cS

Since

u < oo} > 0

(2.3.33)

(2.3.20) then implies that

g(t) > P(t) + c (g2 d S

(2.3.34)

95

g ( 1:) > P (t) + c (

Integrating both sides over

T +

[t, T)

we get a contradiction to

ro

(2.3.35)

P (r) dr

and taking

(a)

lim sups

as

since by hypothesis the

integral of the right side of (2.3.35) is divergent. ii)

Let

For large

t

n n

t

ro

and y(t )

0 >

t(t)n

fixed,

t

ds f(s)

+

y(t ) n

be such that

t > t

3

t

0

.

, ro

> It n (p (s) + I

s

f'

because of (2.3.20) , in the limit, and (2.3.32) . Thus Y (t )

••

ds f(s)

t (2.3.36)

> It n P(s)ds

so that (

lim inf n+ ro l

y(t )

••

ds f(s)

t

1

> lim inf It n P(s)ds n+ ro

and so

o because of But

(b).

I

y (t )

o

ds f(s)


T+n.

f(y)dcrds

[T, T + n )

,

[T,T+n].

E

AnY

(2.3.41)

(2.3.42)

has a right-derivative at

given by

(A y) n

1

(t) =

Joo t

f(y(s))dcr(s)



then integration by parts shows that

(

f(y(s))dcr(s)

f(y(t))

r r{r t

de +

s

t

do } f' (y)y'ds •

Hence

f(y)dcr = f(y(t))P(t)

(

Thus for

T::; t

+ (

P(s)f ' (y(s))y' (s)ds •

::; T + n ,

and proceed to show as in [8] that

I (AnY)' if

T

(t)

I

is so large that

< Q(t)

t

E

[T, T

+ n)

(2.3.43)

100

2b

If

t

T

+n ,

(A

(I n

y)

I

P ( s) ds

I

(t)

=

hence

0

> T .

t

A

n

(B )

B

c

n



T so large that

If, in addition, we require

(

< 1

Q(s)ds < 1

> T

t

(2.3.44)

then o < (An y) (t)

< 2

t

> T ,

(2.3.45)

since we can estimate the inner integrals in (2.3.41-42) by (2.3.43) and (2.3.44)

then gives

also follows from (2.3.42)

(A

If

T < t

1

< t

n

2

y) (t)

(2.3.45).

For

y

E:

B

n

it

that

t>T+n.

constant

< T +n ,

t 2

I ft 1

f

(y (s)

) do (s)

t

f' (y)y'

If necessary we can restrict

T

J s

I

2

do ds

further by requiring that

101

t

b

IJ

Q (s)

2

do

s

I

< 1

Substituting this in the former equation we obtain

Thus

A

n

(B

n

)

B

c



n

There remains to show that

A

n

is continuous:

This

can be done as in [8, p. 82] with the appropriate modifications in the definitions of

a(o)

, b(o)

there.

Le. a(o)

sup{lf(y) -f(x)!

0 < x,y < 2,

b(o)

sup{lf'(y) -f'(x)!

0 < x,y < 2, Iy-xl < o} .

From the above definitions we see that as b(o)

+

0

since If

given

E;,

f

E

C'

c (0)

0

+

0

both

.

0

is chosen sufficiently small so that, for

II x

- yll < 0 ,

2

where

Iy-xl < o}

c (0)


n=m c n n

N+oo

.

-00

Q.J- 1 - max{-P j_ l

I

o}

and 00

R.J- 1 00

Then

1 c. R.J
0

for

n

N.

Then a necessary and

105

sufficient condition for (2.3.48) to be oscillatory is that 00

00

00

+

Proof:

00



(2.3.52)

This follows immediately from Theorem 2.3.2 because

(2.3.51- 52) are equivalent since

P

> 0 .

n

THEOREM 2.3.3: Let

0

satisfy the basic hypotheses of Theorem

2.3.1. a)

If

lim

(2.3.53)

0 (t)

t-+oo

then (2.3.1) is oscillatory. b)

If

o(t)

is non-decreasing then a necessary and

sufficient condition for (2.3.1) to be oscillatory is that

r

tdo (t) =

(2.3.54)

00

to

Proof: P(t)

=

a)

follows immediately from Theorem 2.3.1 since for all

00

t.

Hence (2.3.9) is identically satisfied.

To prove b) we must show that (2.3.9) is equivalent to (2.3.54) . If

0

is non-decreasing and (2.3.9) is finite then (

P(s)ds
0 ,

108

so that (2.3.58) is equivalent to (2.3.54). In particular we can choose all

n

when

-1,0,1, '" b

n

.

a = -1

and

c

n

1

for

We then obtain from (2.3.48) that,

0 ,

>

/:,.2 Y

+ b f n

n-l

o

(y )

n

(2.3.60)

is oscillatory if and only if 00

mb

(2.3.61)

m

This is the discrete analog of Atkinson's theorem [2] which follows from the previous corollary.

Example 1: b

n

Let

=l/(n+l)

and if we choose m

L{ Lc

00

0

0

c

= n + 2, n n=O,l, a = 0,

n = -1 , 0 , 1 , ...

and let

Then (2.1.91) is satisfied

then 00

L {I +1.+ 2

1 }b m i- l

0

1 ••• +_l_} m + 1 m+l

00

>

1 L m+l

00

0

Hence Corollary 2.3.2(a) implies that all solutions of (2.3.48) are oscillatory where (a) of Theorem 2.3.1.

f

is any function satisfying

109

Example 2:

If we let

b

be as in Example 1 above and

n

1 (n + 1) 1+8

n

o

> 0

then

_l_} bm

Yo { I o ci - l




t.

But (2.3.74) implies that

t

f(s)ds

Hence the latter holds for all

t

E

(2.3.76)

[tm- l ' t m )

and thus applying Corollary

2.1.2 again we find that (2.3.73) has a solution and thus (2.3.64) has a non-oscillatory solution which implies that it is non-oscillatory.

This completes the proof.

As a consequence we immediately obtain that the discrete Euler equation 2

/:; Yn-l +

is oscillatory when

Y >

41

Y

+ 1)

(n

2 Yn

o

(2.3.77)

and non-oscillatory when

y

< ! = 4

(see example I, section 2.1, and Example 1 of section 2B). Furthermore we shall have [30, p. 30],

o non-oscillatory when

A S 0

n=O,l, ...

and oscillatory when

A > 0

117

because of the analogous property for

y" +"Ay

o.

CHAPTER 3 INTRODUCTION: The purpose of this chapter is to provide a basic framework for the theory of operators generated by the Volterra-Stieltjes integral equations encountered in the preceding chapters.

The method used here will show that

these integral equations can be thought of as defining generalized differential operators. undertaken by I.S. Kac

[35]

though the application there was

only to differential equations. been used by H. Langer

Such a formalism was

A different formalism has

[41] to deal with the notion of an

operator defined by a Stieltjes integral equation of the form (2.1.0).

The method which we shall use here is a natural

extension of that used by Kac

[35] and its applications will

include differential equations and in particular, SturmLiouville problems with indefinite weight functions and difference equations. In section 2 we shall proceed to define the generalized differential expression

.2 [f]

l

[f (x) -

r a

f(s)do(s)

J

(3.0.0)

119

where

v, a

are real right-continuous functions of bounded

variation, after having given the background material in section 1. In section 3 we shall study the Weyl classification (limit-point, limit-circle) of singular generalized differential operators with an application to the particular case

_y" + q(t)y

h(t)y

where the weight-function interval.

r(t)

t

E

(3.0.1)

[a, 00)

vanishes identically on some

Other applications will include the three-term

recurrence relation

-c n y n+l - c n-l y n-l + b n Yn where

c

n

>

o.

Aa y

(3.0.2)

n n

These will be discussed in section 4.

In section 5 we give some criteria which can be used to determine whether a certain equation is in the limit-point or limit-circle case.

In section 6 we shall be considering

the self-adjointness and, more generally, the

J-self-

adjointness of such generalized operators. In section 7 we discuss the finiteness of Dirichlet integrals associated with (3.0.0) and consider the chain of implications [39]

DI => CD => SLP => LP

120

where these abbreviations stand for Dirichlet, Conditionally Dirichlet, Strong Limit-Point, Limit-Point respectively. Finally, in section 8 we define these notions for a three-term recurrence relation and give some examples.

§3.1

GENERALIZED DERIVATIVES:

v

Let

be two real right-continuous functions

locally of bounded variation on at each interior point

y

lim

E R

a > _00.

Then

,

u (x )

x->-y±O

[a, 00)

lim

x->-y±O

v (x)

both exist and are finite. Associated wi th u (or v ) defined on intervals

(a, (3]

m (a, (3]

u

is a set function

and

-

[a , (3]

in

m

u

[a, 00)

(3.1.0)

u to)

(3.1.1)

[a , (3]

When

is non-decreasing then

a-fini te Borel measure on function

by

[a, 00)

induces a

[55, p. 262].

Since every

of bounded variation is a difference of two non-

decreasing functions such a function will induce a signed Borel measure on

[a, 00)

a-finite

[55, p. 264, ex. 11] which is

fini te if the original function is bounded on

[a, 00)

We

121

will denote such a measure by set

Then, for every Borel

m

u

E ,

(3.1.2)

m (E)

u

where jl

+ , mjl mjl

are the positive and negative variations of

obtained by its Jordan decomposition [24, p. 123].

each function

jl

right-continuous and locally of bounded

variation induces a

a-finite signed Borel measure on

satisfying (3.1.2).

The measure

[m

where

+

mjl , mjl

u

1m I u

defined by

(3.1.3)

are as in (3.1.2) is called the total varia-

jl , v

jl.

are signed measures we say that

absolutely continuous with respect to every measurable set If

jl(x)

is

[a, 00)

I (E)

tion (or total variation measure) of If

Thus

E

for which

[m v

v

if

I (E)

u

is

[m u I (E) = 0 0

for

[24, p. 125] .

v-absolutely continuous there exists a finite-

valued measurable function

m (E) u for every Borel set

E

¢

such that

J

E

¢dm

(3.1.4)

v

[24, p. 131, ex. 4].

The function

¢

appearing in (3.1.4) is called the Radon-Nikodym derivative of jl

with respect to

v

It is unique in the sense that if

is another measurable function with this property then

¢ =

122

v-almost everywhere (that is they are equal everywhere except possibly on a set When and

¢

E

with

[rn v

I (E) = 0) .

is non-decreasing and right-continuous

is a non-negative Borel measurable function the

Lebesgue-Stieltjes integral of

¢

with respect to

is

defined by

J ¢ (x) If

¢

(x) -

J

(3.1.5)

.

is both positive and negative it is integrable with

respect to

When

if it is integrable with respect to

is of bounded variation

t a

¢ (x) dlJ (x)

agrees with the ordinary Riemann-Stieltjes integral whenever the latter is defined [55, p. 261]. When

f1, v

are of bounded variation and

is

v-absolutely continuous,

dv exists

(x)

lim

- MO

v (x

+ h) -

mv-almost everywhere, in particular when

point of discontinuity of

v

it exists and

v(x+O) -v(x-O) ¢ (x)

(3.1.6)

v (x - h)

x

is a

123

where to

v

¢

is the Radon-Nikodym derivative of

defined in (3.1.4).

with respect

(For general information on these

derivatives, see [24, p. 132].) From (3.1.4-5), for

a,S

E

S±O

I IS±O

-

a±O

[a, (0)

,

[18, p. 134],

¢ (x ) dv (x)

(x) • dv (x)

a±O d v (x)

When

v

are of bounded variation and have no common points

of discontinuity then

I

S±O

13+0

a±O

-

IS±o a±O

.

This is the general formula for integration by parts. is a single point i ts (or

§3.2

When

x

u (or v ) measure shall be denoted by

It is defined by (3.1.1).

v{x}).

GENERALIZED DIFFERENTIAL EXPRESSIONS OF THE SECOND ORDER: In this section we shall essentially pursue the

approach of Kac [35] in the definition of a generalized differential expression of the second order on some interval I,

i.e. for

x

E

I,

a

R,[y] (x)

==

d {y+(x) , -dv(x) -

E

I

fixed,

Ja+O+O y(s)do(s) } X

(3.2.0)

124

where and

was assumed to be locally of bounded variation on

G

I

was non-decreasing.

v

It shall be convenient to assume that addition, right-continuous and bounded variation.

v

G

is, in

an arbitrary function of

We shall see below that it is still

possible, in the latter case, to define (3.2.0).

Basic assumptions: Throughout the remainder of this chapter we shall assume that:

v,

a)

both have at most a finite number of discontinuities

G

in finite intervals (see the hypotheses of §l.l). b)

If

I

is a finite interval then both

continuous at the end-points of infini te interval then both

v,

If

I G

v,

G

shall be

I

is a semi-

shall be continuous

at the finite end, and that neither has a discontinuity at infinity, i.e.

lim v(x)

lim

X-HO

X-+oo

G

(x)

both should exist (may be infinite) . Thus both a), b) shall be assumed in addition to the usual hypotheses of right-continuity and bounded variation for

v ,

G



125

Let

v

be two right-continuous functions of bounded

variation. induces a

As we saw in the previous section each of these a-finite signed Borel measure on

addition, we assume that respect to

m

m

u

I.

If, in

is absolutely continuous with

then the Radon-Nikodym derivative

v

dm u

(3.2.1)

¢ - dm

v

exists

v-almost everywhere and we have relation (3.1.4).

Moreover (3.2.1) agreees with (3.1.6) Let

v, a

v-almost everywhere.

be two right-continuous functions of

bounded variation on A function

[a, b] . f

is said to belong to the class

i)

f

is absolutely continuous on

ii)

f

has at each point

derivative

,

f+(x)

x

E:

[a, b]

[a, b]

,

a right-

and the function

_ f:(X) - JX f(s)da(s) a

is

v-absolutely continuous on

We note that (ii) necessitates that variation on

[a, b]

an "associated number"

The quantity [35, p. 212].

[a, b] be of bounded can be termed

126

Thus the preceding discussion shows that if

f

E

Vv

the

quantity

.Q,

exists

{< (x) -

[f] (x) - - dv

v-almost everywhere on

r a

(3.2.2)

f (s) do (s) }

[a, b]

(i.e., it has meaning

everywhere except possibly on a set on which the total variation of

v

is zero).

A particular case of a generalized differential expression is (2.1.0). m v

To see this we let

is Lebesgue measure and let

(2.1.0).

y(x)

v(x)

=x

so that

be any solution of

It is clear that

(x) - JX y(S)dO(S)}

0

a

so that equations of the form (2.1.0) can be brought into the form (3.2.2). By a soZution of the generalized differential equation

(x) -

is meant a function

f

E

V

r a

v

f(S)dO(S)}

, l/J

(3.3.20)

- l/J (x) cj>' (x)

(3.3.21)

1 .

are real for real

A

and satisfy

cj>(a, A)COS a + cj>' (a, A)sin a

0

l/J(a, A)sin a - l/J'(a, A)COS a

= 0

Every solution

cj>

of (3.3.9)

is, up to a constant multiple,

of the form

e where

m

a < b
+ ml/J

(3.3.22)

is some number which depends on

A.

Now let

and introduce a real boundary condition at

b

by

requiring that g(A)

for

S

E

[0,

TI)

- y(b, A)COS S + y'(b, A)sin S



o

(3.3.23)

The eigenvalues of (3.3.9) are then the

zeros of the entire function

g(A)

Since these eigenvalues

must necessarily be real (Appendix III, Theorem 111.1.2),

137

(3.3.23) does not vanish identically and consequently the zeros have no finite point of accumulation. We now seek

m

such that the solution

satisfies the boundary condition (3.3.23).

e

above

A simple computa-

tion shows that cot S ¢ (b , A) + ¢ , (b , A) cot S 1jJ (b , A) + 1jJ' (b , A)

m

Thus as

m = m (A

vary

A

are entire functions of

1jJ'

m

S

A, b ,

z

= cot S and fix

S varies from

0

to

TI,

b, A

I

,

1jJ ,

A.

for the moment, then as

varies over

Z

¢, ¢

(Appendix III, Theorem 111.1.0),

A and real for real

is meromorphic in

If we let

Since

b , S)

,

(_00, 00)

so that

the image of the real axis under the transformation Az + B +D

m

A = ¢ (b , A)

where D

=

1jJ'

(b, A)

m-plane. lies on

Thus C b•

and

e

- Cz

B=¢'(b,A)

AD - BC

C

=

1jJ (b , A)

C in the b will satisfy (3.3.23) if and only if m 7

0,

(3.3.24)

'

is a circle

From (3.3.24) we have

z = _B+Dm

(3.3.25)

A+ Cm

so that the circle is given by

C b,

which is the image of

1m z

0,

138

(A + Cm)

(B

+

Dm)

(A

Since every circle with center

+

Cm)

o .

(B + Dm)

and radius

y

r

(3.3.26)

can be

described by

o

(3.3.27)

we see on comparing coefficients of (3.3.26-27) that center of

is given by

AD - BC CD - DC and its radius

(3.3.28)

is given by lAD - BC I

(3.3.29)

ICD - DC I A , B , C , D

Substituting the values for

into (3.3.26) we

obtain the equivalent equation [ee] (b)

0

.

(3.3.30)

In the same way we find that

1

Hence,

[1/1] (b)

AD

[1/Il/J] (b)

DC - CD

(b)

AD

BC

BC

139

mb

[ 0,

if and only if

b

is any sequence and an 0 n [3, p. 129, Theorem 5.4.2] for any A 1m A 0 ,

Aa y

-c n y n+l - c n-l y n-l + b n Yn has at least one nontrivial solution

then

n n

w

=

(w) n

in

£2(V

I).

147

§3.5

LIMIT-POINT AND LIMIT-CIRCLE CRITERIA: In this section we shall give some conditions on

and

v

which will enable us to establish the limit-point or

a

limit-circle classification of

{Y'(X)-r Yd a}

in the space

v

L

2(V;

I)

where

over the interval

(x )

V(x)

[a, x ]

and

(3.5.0)

AY

a

is the total variation of I = [a, (0)

,

a >

-00



This space was defined in (3.3.3).

LEMMA 3.5.0: Let

a, v

be right-continuous functions locally of

bounded variation on AO x

Em a,

such that

[a, (0)

Suppose that there exists

a(x) - AOV(X)

is non-decreasing for

say.

Then (3.5.0), with

A

AO'

y(x) > 1

Proof:

Let

y(x)

has a solution x > a •

y(x)

with (3.5.1)

be the solution of (3.5.0) satisfying the

initial conditions y(a)

Then, by Theorem 3.2.0,

o •

1

s ' (a)

y(x)

is a solution of the integral

(3.5.2)

148

equation y(x)

for

x

a.

an interval for

x

E

+

1

Since

r(Xa

8 > 0

[a, a + 8]

(3.5.3)

AO\)(S))

then by continuity there exists

1

y(a)

[a, a+ 8]

s)y(s)d(a(s) -

y(x) > O.

in which

Then,

the integral in (3.5.3) is non-negative

and so y (x )

Since in

y (a + 8) > 1

[a + 8 , 8

y(x)

> 1

X

E

[a, a + 8]

there exists

8

Consequently for

1]

y(a+ 8) +

JX

a+8

such that

> 0

1

x

(3.5.4)

y(x) > 0

in such an interval

(x- s)y(s)d(a(s) -

AO\)(S))

and so y (x)

> 1

Repeating this process we obtain an increasing sequence of real numbers.

n

)

It is then necessary that

lim 8 n+ oo

otherwise if

(8

lim 8

n

8*

n

then

(3.5.5)

00

y(8*)

1

so we could repeat the above process past diction proves that (3.5.5) holds and thus

by continuity and 8*.

This contra-

149

y(x) > 1

THEOREM 3.5.1: 3.5.0.

Let

a, v

x

> a



satisfy all the hypotheses of Lemma

Suppose further that

('ldV(t) a

I

(3.5.6)

00

Then (3.5.0) is limit-point at

Proof:

It suffices to show that, for some

solution of (3.5.0) which is not in

L

2(V;

where the latter exists by hypothesis. there exists a solution

y(t)

A,

there is a

I)

From Lemma 3.5.0

of (3.5.0) such that (3.5.1)

holds. Then for such a solution,

CldV(t) a

hence

y

is not in

L

2

(V; I)

I

00



COROLLARY 3.5.1 : Let

(a ) n

be a sequence such that 00

L 0

Let

(b ) n

sequence.

la n I

00

be any given sequence and

(c ) n

another positive

150

AO such that

If there exists a real number

b

n

- c

- c

n

n-

1 + A a 0

>

n

n=O, 1, ...

0

(3.5.7)

then cy +c Y -by n n+l n-l n-l n n is limit-point at solution

(Yn)

Aa

y

(3.5.8)

n

A there corresponds a

i.e. for some

00

n

such that

(3.5.9)

00

Proof:

We note here in passing that Lemma 3.5.0 extends to

equations of the form (3.4.11) when

p(x) > 0

right-

continuous and of bounded variation satisfying locally

L (a ,

The proof is similar with minor changes.

00)

We define a step-function (t)

n

p(t)-1

v(t)

with jumps at the

by v (t ) n

and require that n=O,l, ... has solutions

v

v (t

-a

- 0)

be constant on

We define y(t)

n

[t n- l ' t

n) as in (3.4.7).

o(t)

such that

(3.5.10)

n

y(t

n)

= Yn

Then (3.5.0)

satisfies the

recurrence relation (3.5.8). (3.5.10) and the hypothesis imply that (3.5.6) is satisfied.

Moreover for

A = An '

(3.5.7) implies that

151

o - 1..

0

is non-decreasing.

V

Thus Theorem 3.5,.1 applies and so

j""ly(t)1 a

2!dV(t)[ 00

which implies (3.5.9) In this form, Corollary 3.5.1 is a minor extension of [3, p. 135, Theorem 5.8.2] where the case

a

n

> 0

is considered.

THEOREM 3.5.2: Let

0, V

be right-continuous functions locally of

bounded variation on

[0,

r t

and

00)

I do (t) I


-00 by

V.

As usual we denote the total The operator

L

generated by the

"formally self-adjoint" generalized differential expression

[y] (x)

-

y' (x) -

r

y(S)da(S)}

is defined as in section 3.2 of this chapter.

V of

L consists of all functions

f

E

(3.6.0)

Thus the domain

L 2(V; I)

such that

157

i)

f

is locally absolutely continuous on

ii)

f

has at each point

iii)

The function ].1

is iv) For

x

L

V

E

r

(x) :: fl (x) -

00)

a right-derivative

f(s)da(s)

a

V-absolutely continuous locally on

t[f] (x)

f

[a,

E

I .

E

L

2(V;

I .

I)

is defined by

Lf

(3.6.1)

t[f]

The notions of "regularity" and "complete regularity" of the expression (3.6.0) are defined in [35, p. 249] in the case when

v

is non-decreasing. In general we shall say that the end

if the set of "points of growth" of

V(x)

a

is regular

and the set of its

values is bounded from below and if the set of points of growth of

is bounded below and, in addition

a

bounded variation in some right-neighborhood of

a

a

is not regular then it is said to be singular. is completely regular if

a

I

= [a, a

The end

a a

00)

[35]. It is then clear from the

latter definitions and the basic assumptions on section 3.2 that the end

If

belongs to the interval concerned.

These definitions are due to Kac In our case

is of

v, a

is completely regular.

of If

158

I

[a , b]

then the ends are both completely regular.

We note that since

v,

are continuous at

0

a, b ,

in the case of a finite interval, the left and rightderivatives of a solution and be equal. a, b

y(t)

of (3.5.0) will exist there

This can also be seen by extending

by setting each equal to

v(a)

, o(a)

respectively on some interval containing

and

v, v(b)

0

past , o(b)

[a, b]

THEOREM 3.6.1: Let

Q, ['J

on

I

I

=

[a, b]

Let

g

be a finite interval and consider 2(V,I) be any function in L The

equation Q,

has a solution

y(x)

[y]

(3.6.2)

g

satisfying y(a) y' (a)

if and only if the function

y(b)

0

y' (b)

g(x)

We note that

f

is

(3.6.4)

0

is

solutions of the homogeneous equation

Proof:

(3.6.3)

J-orthogonal to all

Q,[y]

J-orthogonal to

=

0 .

g

if and only

if b

Ja

f(x)g(x)dv(x)

o .

(3.6.5)

159

(The

J-orthogonality stems from the

2

L (Vi I)

,

J-inner product in

see Appendix 111.3.)

This theorem can be eroved exactly as in [46, p. 62, Lemma 1].

For by Theorem 3.2.0, and Theorem 1.3.1 the

equation (3.6.2) has a unique solution which satisfies y (a)

=

0,

0

z1' z2

Let

0

=

Y I (a)

be a fundamental system of solutions of

which satisfy

1

o

o

1 .

Applying Theorem 3.2.1 to b

Ja

g (x)

Ja

]

]

conditions above,

b

[z . ] (x)

=

]

Jab

(x)

d v (x) -

[y I

]

b

z.

and

z. (x) dv (x)

Ja y (x) By noting that

y

0

j

=

we find

z.] (x) dv (x) -

Z • -

1,2,

]

-, b y z .] • ] a

(3.6.6)

and using the boundary

(3.6.6) reduces to

g (x) Z . (x) d v (x) ]

{ -y' (bl

j = 1

(3.6.7) y(b)

j

=

2

Thus '3.6.4) is satisfied if and only if (3.6.7) vanishes for

160

j

=1

, 2,

solutions

i. e.

Of

f

if

t[z]

=

J-orthogonal to a fundamental system of

0

and thus the conclusion follows.

v,

Now since the measure induced by

in (3.6.0), is

absolutely continuous with respect to the measure induced by V

the quantity (section 3.1)

dv(x) dV(x)

exists

[V]

(3.6.8)

.

Consequently, the expression

{y'(X) -

J:

Yda}



{Y' (x) -

r a

Yda} (3.6.9)

V-almost everywhere by [24, p. 135, Ex. 1, and Theorem A] . Thus if we denote by

t t [Y]

the expression defined by

YE V

v

(section 3.2),

tt[y] (x)

we see that related to

tt t

-

{Y' (x ) -

r a

(3.6.10)

Yda}

is another generalized differential expression by (3.6.9), i.e.

for

Y

E

Vv

dv dV (x) • t [Y] (x )

(3.6.11)

and both of these are defined on the same domain (3.6.10) gives rise to an operator

Lt

on

V,

v

\)

where

Thus D

is

161

the domain of

L

y ED,

defined earlier, such that for

dv dV· Ly

or, in terms of the Gram operator

(3.6.12)

J

defined in Appendix

III. 3,

JL .

(3.6.13)

v

If, in Theorem 3.6.1, we assume that then we can replace

£[y]

in (3.6.2) by

conclusion will then follow with usual orthogonality in

L

is non-decreasing

2(V

i

£t[y]

and the

J-orthogonality being the

I)

since

v

=V

in this

case. We now define a new operator, denoted by

, Vo

domain

= [f

E

V

defined by

f

=0

with

[46, p. 60]

outside a finite interval

[a, S]

c

(a, b l ] (3.6.14)

The restriction of the operator Thus for

y

L

to

v'o

defines

L' o

E

Ly

Similarly we can define

,

(L n)

(3.6.15)

,Q,(y]

v'o (3.6.16)

162

THEOREM 3.6.2: a)

If

Y

,

E

VO '

Z

V

E

then (3.6.17)

[y, Lz]

where

[,]

is the

J-inner product defined by the left hand

side of (3.6.5). Moreover, the operator

L'

o is

J-hermitian, i.e.

y,ZEV

b)

If

v' YE:o

Z

E

,

(3.6.18)

O'

V then writing

we

have

where

is the inner product in Again, the operator

L

1

2

(V; I)



is hermitian, i.e.

y,

Proof:

L

Z

E

V •

(3.6.19)

Both a), b) can be shown as in [46, p. 61] making use

of Theorem 3.2.1 so we omit the details.

We now proceed as in [46, §17] in defining the operators Suppose that the interval i , R- t

are both regular on

[a, b]

[a, b] .)

is finite.

(Then

163

V

We define the domain

{y

and, for

y

E

y(a)=y(b)=y'(a)=y'(b)=O}

V:

E

L

o of the operator

o by (3.6.20)

Vo ' (3.6.21)

Ly

(3.6.22)

THEOREM 3.6.3: For any

y

Vo '

E

Z

E

[LoY, z] ==

t

( L oY , z)

and the operator

i. e.

for any

L

y, z

o E

is

V0

V (3.6.23)

[y, Lz]

t

(3.6.24)

(y , L z)

J-hermitian while

Lto

is hermitian,

'

(3.6.25)

(3.6.26)

Proof: similar.

We refer to [46, p. 62, I, II] since the proofs are

164

LEMMA 3.6.1: t Ro =

Let

range

0

Lto

f

and let

be the set of all

M

Q,t[z] = 0 .

solutions of the equation Then

t

H

Proof:

Ro

Since all solutions of the homogeneous equation are

continuous functions on so

M

c

H.

[a, b]

H

Hence

lies in H

Hand

is a finite

2.

replaced by

Q,t[y]

implies that

M

of dimension

solution of (3.6.2), with

Vo •

they all belong to

It is also readily seen that

dimensional subspace in

in

(3.6.27)

+ M .

Y

If

Q,t

is a

then

y

Thus the existence of

y

Theorem 3.6.1 then states that

M.

if and only if it is orthogonal to

is

g

Since

is a Hilbert space the decomposition (3.6.27) follows.

THEOREM 3.6.4: The domain

Proof:

LO

o of the operator

Since the domain

and

gonal to have then solution of 3.6.3,

V

Vo

L

o

is the same for the operators

it suffices to show that every element is zero. (h , y)

=

0

Q, t [z] = h

Letting for all For

H.

is dense in

h

h

ortho-

be such an element, we Let

z

be any

we have, by Theorem

165

(L

and so

z

t z, y)

is orthogonal to

previous lemma, and so

o

(h , y)

Consequently

£ t [z]

-=

E

M

by the

h = 0

i.e.

0,

z

We recall that a set is dense in a Krein space if it is dense in the Hilbert norm topology.

Thus the latter

Va

theorem expresses the fact that the domain

La

operator

of the

H.

is dense in the Krein space

THEOREM 3.6.5: The operators

are

and

J-symmetric and

symmetric respectively.

Proof:

Note:

This follows from Theorems 3.6.3 - 4.

The rest of the results in [46, §17]

can be similarly

shown to be true in this more general setting. in the regular case the operator operator whose

LX

J-adjoint

a

La

Thus, e.g.,

is a closed

is equal to

L

J-symmetric

[46, p. 66,

Theorem 1].

In the singular case, i.e. a > -00,

I = [a, 00)

we follow the approach outlined in [46, §17.4]

where we begin with the operator We recall that domain

when

,

Vf") •

I

La

and

L1

L'

a

defined in (3.6.14-15).

are both defined on the same

166

THEOREM 3.6.6: The domain of definition

L'

and

is therefore a

o

Proof:

is dense in

of

o

H

J-symmetric operator.

An argument similar to that of [46, p. 68] shows that

Vo ,

when viewed as the domain of

Thus

L

I

V'

L

1

H

is dense in

L'

is a symmetric operator, by Theorem 3.6.2, and

1

0

J-symmetric.

is

We now take the closure of

L

, o

-L ,

in the Hilbert

o

space topology and define

Lo

it then follows from the preceding theorem that closed

is a

J-symmetric operator. We now proceed to find a property of the domain

Lt

of L

(and so

L)

when

£t

V

is in the limit-point case in

2(V;I)

LEMMA 3. 6 • 2 :

[ 14] •

For any set of six functions {g

q

:

1

q

[0 , 00) point

3}

{f p:

1

P

3}

each being locally absolutely continuous on

and each having a finite right-derivative at each x

E

[0, 00)

167

det {[f

p

q ] (x) } q

o

X

E

[0, 00)

where [fg] (x)

Proof:

See [14, p. 374].

LEMMA 3.6.3: Let

be a Borel measure on

II

and let

[0, 00)

2

L (u )

be the space of square integrable "functions" with respect to ll.

Suppose that

f, g

are complex-valued

u-rnea surab Le

functions which satisfy

2 gEL ( ll i [ 0 , X » )

for all

X > 0

and that

Then

o .

lim

Proof:

This result can be proven in exactly the same way as

in [15, p. 42] with the necessary modifications.

Let

L

t

be the operator defined by (3.6.12),

Then by Theorem 3.3.2, for

1m A

0,

the problem

(3.6.10).

168

"Ay

on

[0,

has at least one nontrivial solution in I

=

[0, (0)

(3.6.28)

(0)

L 2(V; I)

where

Using this result and Theorem 3.3.1 along with

Lemmas 3.6.1-2 we can show, by adapting the argument of Everitt [15, pp. 42 - 45] to our situation, that whenever (3.6.28) is limit-point there follows

lim

x+ oo

{f (x) g

(x) -

f: (x) g (x)

}

o

(3.6.29)

Conversely if (3.6.28) is limit-circle, then it must be so for "A

=

o.

In this case it is possible to find two real linearly

independent solutions

¢,

of

o which satisfy (3.3.18-19) say. L

2(V;

I)

and consequently in

(3.6.30)

By hypothesis these are in D.

Moreover, by (3.3.18-19)

hence lim

(x)

0 .

x+ oo

Summarizing, we obtain

THEOREM 3.6.7: A necessary and sufficient condition for (3.6.28) to

169

be limit-point is that for all

f, g

E

V

(3.6.29) be

satisfied.

THEOREM 3.6.8: Let

a

satisfy (3.5.11).

Then (3.6.28) is limit-

point if and only if

try]

is limit-point (in the

(3.6.31)

"Ay

2

L (V ; I)

sense)

where £

and

are

related by (3.6.11).

Proof:

For Theorem 3.5.2 implies that (3.6.28) is limit-

circle if and only if

(3.6.32)

where

V(t)

However the latter is equivalent to

00

(3.6.33)

and Theorem 3.5.2 again implies that (3.6.31) is limit circle if and only if (3.6.33) and so (3.6.32) is satisfied.

The

result now follows.

COROLLARY 3.6.1: In order that (3.6.31) be in the limit-point case at infinity (in the space

2 L (V; I))

it is necessary and

170

sufficient that

o

lim [fg] (x) x-+ oo

where of

(3.6.34)

is defined in Lemma 3.6.2 and

[

L , Lt

a

La

of

V

where

a

is the domain

defined earlier.

La , Lat

We now define the operators

V

V

Let the domain

be defined by

{f

a

[0, 'IT)

E

V:

E

,

f(O)cos

and for

La t

Similarly

L a

0.-

f

,

.

f+(O)sln a

Lf .

f

f

Lat

in the limit-point case,

= JL a t La

First of all we note that if

(3.6.35)

Va

E

(3.6.36)

is defined on the same domain

or what is the same,

O}

E

Va

V a

and

(3.6.37)

We now proceed to show that, is self-adjoint. f, g

[fg] (0)

E

o .

Va

then (3.6.38)

Next the Lagrange identity (Theorem 3.2.1) shows that, for f,gEV

a

,

X>O

171

-[fg] (x) + [fg] (0)

Consequently if

x

+

00

is limit-point and

f , g

V

E

ex

, we let

in (3.6.39) and use Theorem 3.6.7 and (3.6.38) to find

(f,Ltg) ex

and so

(3.6.39)



L

t ex

(V

is symmetric

it contains the domain the singular case by

f , g

E

V

is dense in

ex

(3.6.40)

ex

L2

(V ; I)

since

V

o of the operator L o defined in L = The proof of this is similar o

to that in [46, p. 71, VI])!.

In the S3-me fashion it can be

shown that

[L

so that

Lex

ex

is

[f , L g]

f, g]

f , g

ex

E

V

(3.6.41)

ex

J-sYmmetric.

THEOREM 3.6.9:

V

In the limi t-point case , the domain

domain of self-adjointness of

if and only if

V

ex

ex

the following properties,

i) ii)

For all g

If

f

E

E

Vex ,

f, g

V

E

V

ex

satisfies then

g

E

Vex

[fg] (0)

o ,

[fg] (0)

o

for all

is a has

172

Proof: We note that this result is a particular case of a theorem of Naimark [46, p. 73, Theorem 1] and can be proven similarly. With

Va

defined as in (3.6.35) a simple computation

shows that both (i) and (ii) are satisfied in Theorem 3.6.9 and consequently, in the limit-point case, adjoint.

Lt

On the other hand, if

a

L

t a

is self-

is self-adjoint then

the deficiency indices [46, p. 26] of the operator are

(0, 0) .

Consequently the equation

(3.6.41)

AZ

has no non-trivial solution in

L

2(V;

Since

I)

Lat

is

self-adjoint (3.6.41) implies that the problem

Ltz

AZ

Z(O)cos a - z' (O)sin a

has no solutions in point.

L

2(V;

I)

0

Thus (3.6.28) is limit-

Hence we have proved

THEOREM 3.6.10: The equation (3.6.28) is in the limit-point case in if and only if the operator self-ad-joint.

Lt , a a

E

[0, 'IT)

is

173

In the following discussion

(L t)

a

*

,

will denote

LX, a

the Hilbert space adjoint and Krein space adjoint of the

Lat , La

operators

self-adjoint.

respectively.

Then by Theorem 3.6.10,

L

limit point case and so its

Let us suppose that

Let

exists.

LX a

J-adjoint f

0

E

a

a , g

[L

ct

E

is a

(3.6.28) is in the

J-symmetric operator and

We denote its domain by

oax

f , g]

[f , L aXg ]

(3.6.43)

Now since

L

t a

JL ct

Moreover

(3.6.44)

Substituting (3.6.44) into (3.6.43) we find that

g

E

* O(LaJ)

But

hence

g

0

E

J-symmetric

.

a

0

a

Consequently c

OX a

Hence

OX

c

ct

0

ct

0

and since

ct

L

a

is

vX a

and so

be

J-self-adjoint for

L

a

is

J-self-adjoint. On the other hand, let each

a

E

[0,

TI)

L

ct

so that we have

L

a

174 [f , L g]

f , g

[L f , g]

a

a

E

V

(3.6.45)

a

Using the Lagrange identity in (3.6.45) we find that for

o Since

lim JX{fIg - gL f}dV x..... oo 0 a a

f, g

V

E

g (0) f'

a

lim [gf' x..... oo

=

f, g

Va .

E

suffice to show that if f

EVa

'

holds.

For if

[fg] (x)

_ f (x) g' (x )

f

E

V

f

:l'

,

a

A

S

V

E

R

are real.

S

f, g

f = f + if R r

where

E

a

Thus let f

E

V

a

, g

g

E

E

V

f, g

E

E V

s ,

VS'

a,

S E [0, TI)

then (3.6.46)

f' (x)g(x)

,

g = gR + ig r

.

a

E

[0 ,

Hence for given

:l'

f

since

TI)

f

r

E V

s

R

(0)

,

where

The result now follows.

f , g

S

For this it would

are any two real-valued

for some

TI)

V.

f, g

similar result holds for [0 ,

(3.6.46)

We now wish to show that the latter

equality in fact holds for all

functions,

so that

o

lim {f(x)g' (x) - f' (x)g(x)} x ..... oo for all

o

g' (0) f (0)

(0)

.

be two real valued functions with a

We will show that under certain

(0)

175

hypotheses on

o , v

g*

we can find a function

Va

E

such

that [fg] (x )

[fg*] (x )

for all large

One such condition is the following: continuous so that for some

Let

v-absolutely ¢ ,

¢dv

in the sense of the measures defined by

0

and

v.

Let

be defined by

g*(x)

where

be

0

v-measurable function

do

g*(x)

x.

a, b

1

x

q Ix )

o

ax + b ,

are to be determined.

1


a .

x-""oo

Thus for

x > X ,

(3.7.9)

If

f

is uniformly bounded above on

inequality implies that zero.

Consequently If

f

g'

g' (x) L

2

(X ,

[X,

00)

then the latter

is uniformly bounded away from which is a contradiction.

00)

is not uniformly bounded then there exists an

increasing sequence

{x} n

with

x

n

-""

00

along which

183

f(x )

-+

n

n

00

(3.7.9) then implies that

If(x)

I

-+

00



> 0

for

x > Xi'

say,

and hence

If'

(x)g' (x)

I

>

!

lal

Integrating the latter over find

r

If' (x)

f(x) [Xi' x

g'

(x) I dx

and letting

n]

n

-+

00

we

00

Xi

a contradiction, by the Schwarz inequality, since both f' , g'

E

L

2

Hence the conclusion is that

(a, (0)

CD => SLP .

Again, in general, this implication is irreversible [17, p. 313].

Thus

DI => CD => SLP => LP .

(3.7.10)

We now interpret these results for three-term recurrence relations, the theory having been developed in the case of ordinary differential expressions.

§3.8

DIRICHLET CONDITIONS FOR THREE-TERM RECURRENCE RELATIONS: Let

c

n

> 0

(c)

for all

(b) n

n

n.

Let

be real sequences and suppose that (a) n

be a sequence of real numbers

184

where

a

0

(t) n sequence of real numbers defined by n

for all

n.

Let

be an increasing

n=O,l, ...

where t

n

+

c_

l as

00

and

> 0

n

+

t_

is fixed.

= a

l

v, a

We also assume that

00

Now define step-functions both

(3.8.1)

be constant on

v, a

-

0)

C

n = 0 , 1 , ...

[t n- l ' t n )

that these have discontinuities at the

a(t ) - o I t. n n

by requiring that

(t

and

only, given by

n)

n=O,l, ...

n +c n- I-b n ,

and v(t)

n

We

-

v(t

n

also suppose that

- 0)

v, a

-a

n

n=O, 1, . . . .

are both continuous at

(3.8.2-3)

a

and

that neither have a jump at infinity. Let

be summable and consider the differential

equation £ [y] (x )

(x)

X

E

[a,

00)

(3.8.4)

where

v, a

are defined above.

Rewriting the solution of the

above as the solution of a Volterra-Stieltjes integral

185

equation we see that, using the methods of Chapter 1, the solution

y(t)

then

satisfies the recurrence relation

-c

n

y

n+l

- c

is linear on

n-l

y

n-l

+b

Y

n

[t n- l ' tn)

and if

yn

= y(t n )

n=O,l, ...

n

(3.8.5)

where

= ¢(t n )

¢n

Thus the domain

V

of the operator

generated by the

L

generalized differential expression above consists of polygonal curves, i.e.

continuous and linear on Moreover the space space

.Q, 2 ( I a I )

V

each function in

is absolutely

for n == 0 , 1 , ... n- l ' tn) L 2(V; I) becomes, in this case, the

i. e.

[t

f

E

.Q, 2 ( 1 a

if

1 )

00

I [a n Ilf n 12

Since the domain


g

n

- g

n

c

n

l'>f

n

189

(3.8.19)

whenever either of (3.8.18-19) exists.

From the latter also

stems the relation

lim f (x)

x-H

g' (x)

lim c n+ co

O

(3.8.20)

f fig n n n

DEFINITION 3.8.3: The difference operator

L

is said to be in the

Strong Limit-Point case at infinity if for all

lim c

n+ co

L

all

n

f

n

fig

n

(=

exists

0)



f, g

E

V

(3.8.21)

is said to be in the Limit-Point case at infinity if for f, g

E

V

1 im c { f g- f -g 1 n n n+l n+l n ' n+ co

o .

(3.8.22)

The latter is consistent with the usual definition of limit(See for example,

point for a three-term recurrence relation. [3, pp. 498-99],

[32, p. 425, Theorem 2].)

We note, in passing, that the theory developed in section 3.6 also includes the difference operators as special cases.

Thus (3.8.22) holds for all

a certain difference operator

La

f, g

E

V if and only if

defined by

190

,Q, [f]

L f

ex

f

E

(3.8.23)

V

and

o is

J-self-adjoint in the Krein space

for all

n,

then

L

(3.8.24)

,Q,2(lal)

If

a

> 0

n

is self-adjoint and consequently every

ex

Lex

symmetric extension of

must coincide with

L

(This

ex

statement is also true in the Kreln space setting.)

When

(3.8.22) is satisfied it implies the self-adjointness of the "maximal" operator

L

[3, p. 499], it then follows that

Lex = L Moreover the implications in (3.7.10) are valid and generally irreversible.

EXAMPLE 3.8.1 : Let

c

n

n

1

In

Yn

Define

b

where

1 n

a

1

n

and let

if

n

2

if

n

2

m

some

m > 0

m

by

n

b

1

,

cnY n + l + cn-1Y n- l

n

Yn

o

say.

n

=1

, 2 , •••

A computation shows that

Y E,Q, n

2

and

191

if

zn

is a linearly independent solution then we must have

const n

n

=1

, 2 , •••

by the discrete analog of the Wronskian identity. z

n

Thus if

the Schwarz inequality applied to the latter identity

E

would produce a contradiction since the left side would be finite while the right side diverges.

c n Yn+l +c n-l Yn-l - b n y n

is

LP.

Thus

o

(3.8.25)

However

lim n y toy n n n-+ co does not even exist.

lim inf n y toy n n n-+ co

In fact,

-1

Thus (3.8.25) is not SLP.

lim sup n y toy n n n-+ co

o .

Other examples may be found to

show that, in general, the implications (3.7.10) are irreversible even for three-term recurrence relations.

The

next result follows from remarks 1, 2 of the preceding section.

THEOREM 3.8.1: A necessary and sufficient condition for

192

-c

to be

LP

for all

n

y

- c

n+l

in the f , g

E

Aan y n

y +b y n-l n-l n n

(3.8.26 )

£2(lal)-sense is that (3.8.22) should hold

V

(here

I

> 0 > 0

a

"0)

n

.

COROLLARY 3.8.1: Let for all

la

n

for all

n

and that

independently of the coefficient

Proof:

If we let

f

n

< M

n. Then (3.8.26) is limit-point in the

Thus

0 < c

+

n

0

as

f

n

V,

E

+

b

2

(I a I ) -sense

n

then

for every

00

£

bounded (3.8.22) holds for all

f

f, g

E

E

V. V.

Since the

care n The result now

follows. If we let

c

n

a

1

n

in (3.8.26) we find that the

equation

/:, 2 y

is always

LP

n-l

+b y n n

in the

Ay

n

n=O,I, •..

£ 2 -sense (see [32, p. 436] and [3,

p. 499]).

Unlike the results in Chapter 2, the limit-point,

193

limit-circle theory of difference equations differs substantially from the analogous theory for differential equations.

One reason for this appears to be related to the

general limit-point criterion (3.6.29) and its interpretation (3.8.22) for recurrence relations. and let

f

=

(f ) n

c

should exist and be zero. automatically satisfied.

n = an = 1

lim f

then it is necessary that

£2

E

For if we set

n

Consequently (3.8.22) is On the other hand if we consider

the differential equation

y" + b(x)y

"Ay

X

E

[0, (0)

then the maximal domain of the operator generated by the expression is a subset of

L 2 (0, (0)

Thus if

f

E

L

2

,

f

need not tend to a limit at infinity and can be essentially unbounded.

Hence (3.6.29) is far from being satisfied and

thus conditions have to be imposed upon that, say, if

f

E

V

then

f

and

f'

b(x)

to ensure

have limits at

infinity and (3.6.29) be satisfied. In the following

theorem we show that it is possible

to strengthen the conclusion of Corollary 3.8.1 under the same set of hypotheses.

THEOREM 3.8.2: Let

a

n

Corollary 3.8.1.

0,

and

c

n

satisfy the hypotheses of

194

Then

n = 0 , 1 ,

(3.8.27)

has the Dirichlet property at infinity.

Proof:

According to Definition 3.8.1 it is necessary to show

(3.8.15-16) . Let Since

la

n

I

f

E

£

2

(I a I)

be such that

then

> 0 > 0

f

E

£2

square-summable in the usual sense. £2

i.e. the sequence Since

f

=

(f ) n

f

n

is

is in

then

(M ) n

and the hypothesis

c

n

< M

implies then that

co

for

f

E

V •

The same argument shows that

I

[c n +c n- Illf n

2

1

0

on

[a, b]

q (x)

E

L

l oc

(a, b)

is such that

the problem

-(p(x)y')' + q(x)y

y(a)

y(b)

Ay

o

(4.2.1)

(4.2.2)

admits a denumerable number of eigenvalues having no finite point of accumulation. that both 8.4.6].)

p

-1

,q

E

(Conditions which guarantee this are

L(a ,b)

,

see [3, p. 215, Theorem

In this case the eigenfunctions form a complete

orthonormal set in

L 2 (a , b)

Associated with (4.2.1-2) is the "indefinite" boundary problem -(p(x)z')

I

+ q(x)z

Ar(x)z

(4.2.3)

213

z(a)

where

r(x)

such that

(4.2.4)

is a real-valued function defined on r(x)

measure.

o

z (b)

[a, b]

and

takes both signs on some subsets of positive

The "Indefinite case" is characterized by the fact

that both

q, r

have a variable sign in

[a, b]

(see [53,

p. 288] and not being equal a.e.

LEMMA 4.2.1: Let

f

eigenvalue

A

I: Proof:

be an eigenfunction corresponding to some of (4.2.3-4).

(p

Then

If' I 2 + q If I 2) dx

We multiply (4.2.3) by

over the interval

t a

[a, b]

(pf ' )' f dx + A

A

I

I

b

a

r If

1

2

(4.2.5)

dx.

and integrate both sides

to find

I

b

a

r If

1

2

dx =

I

b

a

q 1f

1

2

dx .

Integrating the first integral by parts and applying the boundary conditions (4.2.4) the result follows.

LEMMA 4.2.2: Let

A,

eigenfunctions

A f, g

be two non-real eigenvalues with respectively.

Then

214

b

Ja i.e.,

f, g

are

o

r(x)f(x)g(x)dx

(4.2.6)

J-orthogonal in the Krein space

L 2(lrl)

and

t a

Proof:

{p (x) f ' (x) g' (x) + q (x ) f (x) g ( x ) } dx

o .

(4.2.7)

We have - (pf')'

+ qf

Arf

(4.2.8)

- (pg ')

+ qg

flrg

(4.2.9)

along with

f(a) = f(b) = g(a)

(4.2.8) by

g

and (4.2.7) by

g(b) f

we obtain, upon integration over

t a

=

O.

Multiplying

and subtracting the results [a, b]

,

r(x)f(x)g(x)dx = fb{(pg')'f- (pf')' g}dx a

and integrating the latter integral by parts we find b [p(g'f - gfl)] a

o because of the boundary conditions.

A"



This proves (4.2.6) since

215

Multiplying (4.2.8) by

g

and integrating over

[a, b]

we obtain

\/\ fb

rfg dx .

a

Integrating the first term in the left by parts we see that

-fa

b

g

(pf')'

r a

pf'g' dx .

Thus

\ Jb

/\

rfg dx

a

o by (4.2.6).

This completes the proof.

Associated with (4.2.1) is the differential operator A

defined in

V(A)

{y E

where for

L

2

(a, b)

y , py'

L 2 (a , b)

y

E

V(A)

V(A)

E

V(A)

AC (a, b) l oc

,

Ay If we let

with domain

-(PY')'

be defined by

+ qy .

defined by

and

Ay

E

L 2 (a , b) }

216

{y

E

V(A)

: y(a)

o}

y(b)

(4.2.10)

and let

Ay then

- is a restriction of A

Y

E

A to

(4.2.11)

V(A)

V(A)

A is, in fact,

and

a symmetric operator [34, §4.11, Theorem 1]. The following lemmas are part of the theory of the regular Sturm-Liouville equation and can be found in [34], thus we omit the proofs.

LEMMA 4.2.3: a)

The regular Sturm-Liouville operator

A

above, is bounded below, i.e. there exists a constant

defined y

E

m

such that

(Af , f)

where

> Y(f , f)

,

(4.2.12)

is the usual inner product in b)

The operator

A

L

2

(a, b)



has at most a finite number of

negative eigenvalues.

Proof:

For part a) see [34, §5.17, Ex. 5.3 0 and §6.7,

Corollary].

For

Part b) is proved in [34, §5.8, Theorem 2].

f

E

V(A)

the expression

(Af, f)

defines a

217

quadratic functional with values

f

V(A)

E

This is immediate if we follow the argument leading to (4.2.5).

LEMMA 4.2.4: We define

{y

D (Q)

D(Q)

E

L

2

by

co

L I x . I I (y

(a , b)

o

co Q(y)

where

(le

j

(4.2.1-2)

)

,

L 1e·1 (y J

(epj)

2

J

2

(a, b)

, ep.)

J

I2


0

a.e.

of the quadratic functional D(Q)

=

D(Q') (b

Proof:

y

E

(Ay, y)

Q(y)

,

Q' (y)

are identical, i.e.

and

J {p I y' a where

the extensions

D(Q')

2

I

+ q Iy

2

I }dx

Io 1e·1 J

(y , 0

a.e.

on

[a, b]

,

q

L(a, b)

E

as in the hypotheses following (4.2.3-4).

and

The eigen-

value problem (4.2.3-4) possesses at most a finite number of non-real eigenvalues.

M

If we let

the number of pairs of distinct non-real eigenvalues of (4.2.3-4),

N

the number of distinct negative eigenvalues of (4.2.1-2)

(which we know is finite by Lemma

4.2.3(b)), then (4.2.15)

M < N .

Proof:

We let

A ' A ' ... , A 1 N_l o

be the negative eigenvalues

of (4.2.1-2) arranged in an increasing order of magnitude. Let

N . Then it is possible to choose the

is orthogonal (in the

2

L (a, b)-sense)

to

(e. )

J

¢0 '

For it is necessary that

(f , ¢.)

J

and so

o

j = 0 , 1 , ... , N-l

so that

f

221

M-l

L

o ,

e.(z., cp.)

i=O

1

1

]

=0

j

The latter constitutes a set of

N

, 1 , ... , N-l •

linear equations in

unknowns where

M > N.

solution

not all zero which we fix.

(e ) j

Thus this system has a non-trivial

necessary that, for such a choice of

Q' (f)

It is then

(e. ) ]

(4.2.19)

> 0

because of a preceding remark.

Q' (f)

M

Moreover,

Q'{Ie.z.} ]

b

]

z.)

, - -'

p (Ie. z .) (Ie. z,) + a (Ie. z .) (Ie. ]] 11 ]] 11

Ia

M-l

L

i, j=O

But since

e.e. ]

1

b

Ia

for all

{ p z ".

]

o

z.1

< i

+ q z . Z. }dx . ]

, j


0

I

+ qy

and continuous on

Consider the equation

[a, b]

ous and is negative in a subinterval of continuous on

[a, b]

L

(I r I i

[a, b l ) = H

functions

f

q(x)

[a, b]

with the property that

sign at least once in 2

(4.2.20)

"Ary

[a, b]

is continu-

and r (x)

r(x) changes

In this case the space

defined by those (equivalence classes of)

such that

is a Krein space with the indefinite inner product given by

f

E

H

(see Appendix III and the references therein for more discussion on these spaces). Let

U (y) i

is

U 2(y)

be the linear forms defined in

Appendix 1.4, equation (1.4.1).

U,,(y)

We denote the relationships

0

223

by

o .

U(y)

(4.2.21)

The problem

Ly

7T

Ary

Uy

o

then defines an eigenvalue problem, i.e., A

E

C

(4.2.22)

seek values of

such that (4.2.22) has a non-trivial solution satisfy-

ing (4.2.21). With some loss of generality, we shall say that the eigenvalue problem

7T

is formally J-self-adjoint if

[f , Lg]

f , g

for all

E

2

C (a, b)

U(f)

For a

[Lf , g]

which satisfy

U(g)

o .

J-self-adjoint problem, non-real eigenvalues mayor

may not exist but, in any case, if they do exist, their number appears to be finite for general boundary conditions also, because of the preceding theorem. results in [53, formulate

Combining the

§4] with the preceding theorem we can

224

THEOREM 4.2.2: The for.nally J-self-adjoint problem

Ly

>..ry

y(a)

y(b)

o

has a finite number of non-real eigenvalues, in some cases none at all, and on

IAI > A

has only real eigenvalues, with

no finite point of accumulation, clustering at minus infinity and plus infinity.

The second part of this theorem is due to Richardson [53, p. 301, Theorem VII].

It would seem plausible that

Theorem 4.2.2 remains true for arbitrary "J-self-adjoint" boundary conditions though we shall not go into this at the present time.

Theorem 4.2.1 extends, with appropriate

changes in the argument, to the general even order formally self-adjoint differential equation

(_l)n(p y(n)) (n) + (_l)n-l(p y(n-l)) (n-l) + ••. +p Y o 1 n y (j) (a)

where

Po > 0

[a , b] that

where Pk

E

y (j) (b)

and i

p. (x ) l

o

=0

, ••• , n-l

changes sign at least once in

is in the range

C (n-k) (a , b)

j

Ary

0 < i




T .

A.

239

Then tla(t) - AV(t)

I


T

and consequently Theorem 2.1.4 implies that (5.1.17) is nonoscillatory for such

A.

On the other hand let (5.1.15) hold

and suppose that (5.1.17) is non-oscillatory for all

A.

Suppose that, on the contrary,

lim t t+ oo

By our hypothesis

I v (t) I -

v(t) < 0

t

7-

0



and so

r c t t.) + nlv(t)

t(a(t) - AV(t))

We now choose

a

I .

so large that

to (t) >

a

-2"

t > T

and tlv (t)

Then, for

t > T,

I

2

t > T

A > 0 ,

t(a(t) - AV(t)) >

.

Since by hypothesis (5.1.17) is non-oscillatory for all

A we

240

can choose

A

so large that a 2

-(A - 1)

where

E >

0

1 > - + 4

is some fixed number.

E

Thus for such a choice of

t(a(t)-Av(t))

Thus

a(t) - AV(t)

t

if positive for

t

>

T .

T,

such A.

An

application of Theorem 2.2.1 shows that (5.1.17) is oscillatory for such

A

This is a contradiction and thus

a

=

O.

This

completes the proof.

In particular when

=0

a(t)

we obtain the result

of Kac and Kreln [38, p. 78, Proposition 11.9°]. original see p. 97, superfluous.

(2), of [38].)

(For the

Again (5.1.15) is not

The latter result had extended a theorem of

Birman [23, p. 93, Theorem 7] since we can let absolutely continuous and then, when

a(t)

= 0,

v

be (5.1.13) is

equivalent to -y"

where

p(x) > O.

Ap(X)y

X

E

(5.1.19)

[a, (0)

Glazman [23, §29] calls this case the

"polar" case though the latter is usually connected with the sign indefiniteness of

p(x)

in (5.1.19).

Because of

Theorem 5.1.0, other criteria for the finiteness of the

241

negative part of the spectrum can be obtained via the nonoscillation theorems of Chapter 2.1.

Moreover because of the

applications to recurrence relations, we therefore obtain some criteria for the finiteness of the negative part of the spectrum of difference operators.

Example 1: and let

c

If we let n

= 1

a (t)

for all

n

and define

0

-

t

or

n

= n

v(t)

for all

by (3.8.3) n

,

then

(5.1.0) includes the difference equation

-

where

a

n

/':,2

Yn-l

n=O,l,

AanYn

by hypothesis.

> 0

...

(5.1.20)

The discrete analog of Birman's

theorem (above) is that the spectrum of (5.1.20) is discrete if and only if co

lim n n-+co

whenever

La

n


0 ,

A(x) - /E"B(x)

r

> 0 •

Hence 2A(x)B(x)

Inserting this in (5.2.12) we obtain

II ox

2 f lido

I;

II

{ ( + 1] A2 (x) + EB 2 } • C (x)

248

where

C

is the quantity (5.2.1).

Replacing

s

by

siC

we

find 2

2

C(s)A (x) + sB (x) where 1 + C s

C (c )

This completes the proof.

When

is absolutely continuous the above lemma can

a

be found in [18, p. 339, Lemma 1] in the case when Our proof appears to be simpler than the case Consequently, if we choose f

E

s

p= 1 .

p= 1

of [18].

1 = 2" we find that for each

V , (5.2.13)

where

C'

C



LEMMA 5.2.3: For every

Proof:

f

E

V ,

f'

lim f (x)

lim

x-" oo

x-" oo

E

f

L

2

(0,00)

(x) f' (x)

and

a .

This can be shown as in Lemma 2 of [18].

For if

249

by (5.2.13).

Since

lim x-+ oo

f

E L

2

we must have

(0, 00)

x

00

J0

(5.2.14)

A simple calculation also shows that

((Lf) (t)f(t)dt

C

{!f'1

2dt+

IfI

2dCJ(t)}

(5.2.15a)

However, since

f,Lf

E

L 2 (0,00)

we must therefore have

lim f (x) f' (x) x-+oo

00

But by taking real and imaginary parts in (5.2.15a), and noting that

CJ

is real, the latter equation is clearly impossible.

This contradiction proves that

f'

E

2 L ( 0 , 00) •

Hence (5.2.13)

implies that

(5.2.15b)

Thus (5.2.15a) implies that

f(x)f' (x) -+ a ,

as

x -+ 00.

But

250

since that

If

(x)

12

-+ S,

I f(x) I f

Since

E

L

2

(0, (0)

as 2

If

r

1

f

f E

00

( 0)

If

I

:

2

(x) f

The following relation implies

(x

_

(5.2.16)

+ J 2 r e ( ff ' ) d t o The lemma is proved.

I

I

(x)

II do (x) I

(x) f' (x)

V.

x -+

o .

o

for all

O.

,

We also obtain Thus

=

a

ff'EL(O,oo)


0 •

The rest of the argument now follows that in Chapter 4, §2. For, an adaptation of Lemma 4.2.2 shows that, if

A,

are

non-real eigenvalues of

Af

Arf

f(i) (a)

o

i

=

0 , ••• , n-l ,

(III.4.6) where

r (x)

is, say, continuous on

[a, b]

and changes sign

306

at least once there, and

A

then

o

f(x)g(x)r(x)dx

(III.4.7)

o where

f, g

(III.4.8)

are the eigenfunctions corresponding to

A,

respectively. Thus we let (111.4.6)

0 '

u,

such that

eigenfunctions

be the non-real eigenvalues of

••• ,

¢0

'

J

1

¢1

'

••• ,

with

, Since

¢M-l

¢. (x) 1

E

D(A)

we

have P

for

n-r j

.

I

¢ j) 1

i = 0 , .•• , M- 1

1

2

dx

Thus we let M-l f (x)

L

j=O

e. ¢. (x)

(III.4.9)

J J

and, as in Chapter 4.2, we see that if e.

possible to choose the coefficients k=O, ... , N - l .

(Af , f)

> 0

M > N J

then it is

such that

This would then imply that



But by substituting (111.4.9) in the latter relation and

307

expanding the form, we shall find that

(Af ,f)

on account of (111.4.7-8).

0

This contradiction then proves

the result.

*Note:

The problem here is the following:

Richardson's

idea is to approximate the eigenvalues of the continuous problem, (py , )

,

+

(q

y(O)

+ Xk ) y

y(l)

0

0

by the eigenvalues of the discrete problem,

m

i

2

I1 (p. l1y. 1

= 0 , 1 , 2 , •••

points

i/m

,m

1-

1)

+ q.1 y.1 + Xk 1, Y 1.

where the values of

are denoted by

0,

y, p , q ,k k.

1

The claim appears to be that for large values of

at the

respectively. ill

the

eigenvalues of the discrete problem with

o are approximations to the eigenvalues of the above continuous problem.

However it is not at all clear that if the discrete

308

problem has non-real eigenvalues then these must necessarily approximate non-real eigenvalues in the continuous case. it is conceivable that these limits may be real.

For

It does not

seem as if enough information is provided in [53] to exclude the latter possibility.

In fact in some cases no non-real

eigenvalues may exist and so one. needs to establish some criteria on the coefficients which will guarantee their existence.

BIBLIOGRAPHY

[1]

APOSTOL, T., Mathematical Analysis, Second Edition, Addison-Wesley, Massachusets, 1974.

[2]

ATKINSON, F.V., On second order nonlinear oscillation, Pacific J. Math.

[3]

(1955), 643-647.

- - - - - , Discrete and Continuous Boundary Problems, New York, Academic Press, 1964.

[4]

LANGER, H., Sturm-Liouville Problems with Indefinite Weight Function and Operators in Spaces with an Indefinite Metric, Uppsala Conference on Differential

Equations 1977, Almqvist & Wiksell, pp. 114-124.

[5]

BaCHER, M., Boundary problems and Green's functions for linear differential and linear difference equations,

Ann. Math. [6]

(2),

(1911-12), 71-88.

_________ , B oun d ary pro b 1&ems



one d i

.

th 5 --

International Congress of Mathematicians, Proceedings, Cambridge U.P., 1912, Vol. 1, 163-195. [7]

BOGNAR, J., Indefinite Inner Product Spaces, Berlin, Springer, 1974.

[8]

BUTLER, G.J., On the oscillatory behavior of a second order nonlinear differential equation, Ann. Mat. Pura

ed Appl. Seria IV, Torno CV, [9]

(1975), 73-91.

CODDINGTON, E.A. and N. Levinson, Theory of Ordinary Differential Equations, McGraw-Hill, New York, 1955.

310

[10]

DAHO, K.

W.N.

and H. Langer, Some remarks on a paper by

Everitt, Proc. Royal Soc. Edinb. 78A,

(1977),

71-79. [11]

, Sturm-Liouville operators with an indefinite

weight-function, Proc. Royal Soc. Edinb. 78A,

(1977),

161-191. [12]

DERR, V. Ya, Criterion for a difference equation to be

non-oscillatory,

(Russian), Diff. Urav, 12,

(4),

(1976), 747-750 or Differential Equations, 12, (4), April 1976, 524-527. [13]

DUNKEL, 0., Some applications of Green's function in nd . . 8 one d i Am. Math. Soc. Bu1., 2 -- serles, _, (1901-02), 288-292.

[14]

EVERITT, W.N., Self-adjoint boundary value problems on

finite intervals, J. Lond. Math. [15]

]J...-, (1962), 372-384.

, A note on the self-adjoint domains of second

order differential equations, Q.J. Math., Oxford (2), 14, [16]

(1963), 41-45. , On the spectrum of a second order linear

differential equation with a p-integrable coefficient, App. Anal., [17]

1972, 143-160.

, A note on the Dirichlet condition for second

order differential expressions, Can. J. Math.

(2),

(1976), 312-320. [18]

EVERITT, W.N., M. Giertz, J. weidmann, Some remarks on

a separation and limit-point criterion of second-order ordinary differential expressions, Math. Annal. 200, (1973), 335-346.

311

[19)

FITE, W., Concerning the zeros of the solutions of certain differential equations, Trans. Am. Math. Soc.

!2.., [20)

(1918), 341-352.

FORT, T., Oscillatory and non-oscillatory linear difference equations of the second order, Quart. J.

Pure Appl. Math. 45, [21)

(1914-15), 239-257.

, Finite Differences and Difference Equations in the Real Domain, Oxford U.P., 1948.

[22]

GANELIUS, T., Un theoreme tauberien pour la transformee de Laplace, Acad. d. Sci. Paris Compt. Rend., 242,

(1956), 719-721. [23]

GLAZMAN, I.M., Direct Methods of Qualitative Spectral Analysis of Singular Differential Operators, Israel

Program for Scientific Translations (IPST), 1965. [24]

HALMOS, P.R., Measure Theory, Van Nostrand Reinhold, New York, 1950.

[25)

HARTMAN, P., Ordinary Differential Equations, New York, Wiley, 1964.

[26]

HARTMAN, P. and A. Wintner, Oscillatory and nonoscillatory linear differential equations, Am. J.

Math. 71, [27]

(1949), 627-649.

, On non-oscillatory linear differential equations with monotone coefficients, Am. J. Math.

Zi,

(1954), 207-219. [28]

HELLWIG, G., Differential Operators of Mathematical Physics, Springer, Berlin, New York, 1964.

[29]

HILBERT, D., Grundzuge einer allgemeinen Theorie der linearen Integralgleichungen, Gottinger Nachrichten,

1 und 2 Mitteilung (1904), 4 and 5,

(1906).

312

[30]

HILDEBRAND, F.B., Finite Difference Equations and Simulations, Prentice-Hall, New Jersey, 1968.

[31]

HILLE, E., Non-oscillation theorems, Trans. Am. Math. soc . , §.i,

[32]

(1948), 234-252.

HINTON, D., and R.T. Lewis, Spectral analysis of second order difference equations, J. Math. Anal. Appl. 63,

2, [33]

(1978), 421-438.

INCE, E.L., Ordinary differential equations, Dover, New York, 1956.

[34]

JORGENS, K., Spectral Theory of Second-Order Ordinary Differential Equations, Matematisk Institut, Aarhus

Universitet, 1964. [35]

KAC, I.S., The existence of spectral functions of generalized second-order differential systems with a boundary condition at a singular end,

Mat. Sb.

(110),

(1965), 174-227, or Translations

Am. Math. Soc., series 2, Vol. 62, [36]

(Russian),

(1967), 204-262.

, A remark on the article "The existence of spectral functions of generalized second order differential systems with a boundary condition at a singular end", Mat. Sb.

7!i., (118), (1968), No.1, or

Math. USSR. Sbornik, Vol. 5, [37]

(1968), No.1, 141-145.

, On the completeness of the system of eigenfunctions of generalized linear differential expressions of the second order,

Akad. Nauk. Armyan. LX, [38]

!,

(Russian), Dokl.

(1975), 198-203.

KAC, I.S., and M.G. Krein, On the spectral functions of the string, Translations Am. Math. Soc.,

VoL 103, 1974.

(2),

313

[39]

KREIN, M.G., On a generalization of investigations of Stieltjes, Dok1ady Akad. Nauk.

(1952),

881-884 (Russian). [40]

KREIN, M.G., Introduction to the geometry of indefinite J-spaces and to the theory of operators in those spaces, Second Math. Summer School, Part I, Naukova

Dumka, Kiev, 1965, 15-92 or Translations Am. Math. Soc. [41]

(2),

22,

(1970), 103-176.

LANGER, H., Zur Spektraltheorie verallgemeinerter gewonlicher differentialoperatoren zweiter ordnung mit einer nichtmonotonen gewichtsfunktion, University

of Jyvasky1a, Dept. of Math., Report 14, [42]

(1972).

LEIGHTON, W., Comparison theorems for linear differential equations of second order, Am. Math. Soc. Proc., 13,

(1962), 603-610. [43]

MASON, M., On boundary value problems of linear ordinary differential equations of second order,

Trans. Am. Math. Soc., [44]

2..

(1906), 337-360.

MOORE, R.A., The behavior of solutions of a linear differential equation of the second order, Pacific

J. Math., [45]

.!?,

No.1,

(1955), 125-145.

MOULTON, E.J., A theorem in difference equations on the alternation of nodes of linearly independent solutions,

Ann. Ma th ., [46]

( 2),

!l, (19 12),

13 7-13 9 .

NAIMARK, M.A., Linear Differential Operators, Vol. 2, Ungar, New York, 1968.

[47]

OPIAL, Z., Sur les integrales oscillantes de differentielle

u" + f(t)u

(1957-58), 308-313.

=

l'equation

0 , Ann. Pol. Mat.,

!,

314

[48]

PICONE, M., Sui valori eccezionali di un parametro da cui dipende un'equazione differenziale lineare ordinaria del second'ordine, Anna1i Scuo1a N. Sup.,

Pisa, Scienze Fisiche e Matematiche, Seria 1, 11, (1909), 1-141. [49]

PORTER, M.B., On the roots of functions connected by a linear recurrent relation of the second order,

Ann. Math., [50]

(2), Vol. 3,

(1901-02), 55-70.

REID, W.T., A criterion of oscillation for generalized differential equations, Rocky Mtn. J. Math.,

I,

(1977), 799-806. [51]

RICHARDSON, R.G.D., Das Jacobische Kriterium der Variationsrechnung und die Oszillationseigenschaften linearer Differentialgleichungen 2. ordnung, Math.

Ann., [52]

(1910), 279-304. , Theorems of oscillation for two linear

differential equations of the second order with two parameters, Trans. Am. Math. Soc.,!l,

(1912),

22-34. [53]

, Contribution to the study of oscillation properties of linear differential equations of the second order, Am. J. Math., !Q.,

[54]

(1918), 283-316.

RIESZ, M., Sur les ensembles compacts de sommables, Acta Sci. Math. Szeged.,

fonctions

(1933),

136-142. [55]

ROYDEN, H.L., Real Analysis, Second Edition, Macmillan, New York, 1968.

[56]

SANLIEVICI, S.M., Sur les equations differentielles des cordes et des membranes vibrantes, Anna1es de l'Eco1e Norm. Sup. Paris, t25,

(1909), 19-91.

315

[57]

SMART, D.R., Fixed Point Theorems, Cambridge U.P., London, New York,

[58]

, CT, No. 66.

STURM, C., Memoire:

Sur les equations differentielles

lineaires du second ordre, J. Math. Pures Appl.,

!, [59]

(1),

(1836), 106-186.

SWANSON, C.A., Comparison and Oscillation Theory of

Linear Differential Equations, Academic Press, New York, 1968. [60]

TAAM, C-T. , Non-oscillatory differential equations, Duke Math. J. , 19, (1952) , 493-497.

[61 ]

WINTNER, A. ,

criterion of osci l: latory stability, Quart. Appl. Math. , 2, (1949) , 115-117.

[62]

A

, On the non-existence of conjugate points, Am. J. Math.,

[63]

, On the comparison theorem of Kneser-Hille, Math. Scand.,

[64]

J...l, (1951),368-380. (1957), 255-260.

F.V. Atkinson, W.N. Everitt, m-coefficient of Weyl

K.S. Ong, On the

for a differential equation with

an indefinite weight-function, Proc. London Math. Soc.,

(3), [65J

(1974), 368-384.

W.N. Everitt, Some remarks on a differential expression with an indefinite weight-function, in Spectral Theory

and asymptotics of differential equations.

Math. Stud.,

13 (Amsterdam: North-Hall and, 1974). [66J

W.N. Everitt and C. Bennewitz, Some remarks on the Titchmarsh-Weigh m-coefficient, in\A Tribute to !ke

,

Pleijel, Uppsala Universitet,

(1980), 49-108.

316

[67]

W.N. Everitt, M.K. Kwong, A.

Zettl, Oscillation

of

eigenfunctions of weighted regular Sturm-Liouville problems, TO appear in J. London Math. Soc.

[68]

W. Feller, Diffusion processes in one dimension, Trans. Amer. Math. Soc.,

[69]

(1954),1-31.

__________ , On second order differential operators,

Ann. of Math., [70]

22,

(2), 61,

(1955), 90-105.

__________ , On differential operators and boundary condi tions, Comm , Pure Appl. Math.,

(8),

(1955),

203-216. [71 ]

__________,

On generalized Sturm-Liouville operators,

in Proceedings of the conference on differential equations,

(dedicated to A. Weinstein), University

of Maryland Book Store, College Park, Md.,

(1956),

251-270. [72]

__________ ,

Generalized second order differential

operators and their lateral conditions, J. Math.,

[73]

!.,

Illinois

(1957), 459-504.

__________ , On the intrinsic form for second order differential operators, Illinois J. Math.,

(1958), 1-18. [74]

Mingarelli A.B., Some extensions of the Sturm-picone theorem, C.R. Math. Rep. Acad. Sci. Canada.,

(4),

(1979), 223 - 22 6 •

!.,

317

[75]

______________, A limit-point criterion for a threeterm recurrence relation,

l,

Sci. Canada, [76]

______________ ,

C.R. Math. Rep. Acad.

(1981), 171-175.

Indefinite sturm-Liouville problems,

To appear in the Proceedings of the Conference

on

Differential Equations 1982, University of Dundee, Scotland, Lecture Notes in Mathematics/SpringerVerlag, N.Y. [77]

Some remarks on the order of an entire function associated with a second order differential \

I

equation, To appear in Tribute to F.V. Atkinson,

Proceedings of the Symposium on Differential operators, University of Dundee, Scotland, Lecture Notes in Mathematics, Springer-Verlag, N.Y. [78]

______________ , Asymptotic distribution of the eigenvalues of non-definite Sturm-Liouville problems, \

/

To appear in Tribute to F.V. Atkinson, Proceedings of

the Symposium on Differential Operators, University of Dundee, Scotland, Lecture Notes in Mathematics, Springer-Verlag N.Y. [79]

Reid, Generalized linear differential systems, Journal of Math. and Mech.,

[80]

(1959) 705-726.

_______________ , Generalized linear differential systems and related Riccati matrix integral equations,

Illinois J. Math., 10, [81]

(1966), 701-722

H. Weyl, uber gewohnlicher Differentialgleichungen mit

und die zugehorigen Entwicklungen

willkurlioher Functionen, Mat. Ann.,

(1910), 220-269.

Subject Index

Conditional Dirichlet,

180

Difference Equation, Dirichlet property,

15 180

Generalized derivatives,

120 ff.

Generalized ordinary differential expressions,

123 ff.

Generalized ordinary differential operators, 156 ff., 225 Green's function, 25-27, 273 ff. Indefinite weight-function, J-self-adjointness,

197 ff.

156 ff.

Limit-circle,

132, 147 ff.

Limit-point,

132, 147 ff.

Non-oscillatory equation,

30

Non-oscillatory solution,

30

Oscillatory equation,

30

Oscillatory solution,

30

Picone's identity,

3

Strong Limit-point, 180-181 Sturm comparison theorem, 10 ff. Sturm separation theorem,

4, 22 ff.

Three-term recurrence relation, 16 Volterra-Stieltjes integral equation, 29 Weyl classification,

129 ff.