• No results found

Non-standard boundary conditions for Sturm-Liouville operators in the limit point case

N/A
N/A
Protected

Academic year: 2021

Share "Non-standard boundary conditions for Sturm-Liouville operators in the limit point case"

Copied!
37
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

AB CD EF

GH

Non-standard boundary

conditions for Sturm-Liouville operators in the limit point case

A.E. Sterk

Department of

(2)

Preface and acknowledgements

During the spring term of 2005 I participated in a small research project con- cerning non-standard boundary conditions for Sturm-Liouville differential equa- tions.This project eventually resulted in a small joint paper with Seppo Hassi, Henk de Snoo and Henrik Winkler.The present thesis is an extension and generalization of the results in that paper.

Several people supported me in the process of doing research and writing this thesis.First of all, I would like to express my gratitude to my supervisor Henk de Snoo for the interesting subject of this thesis and for giving me the opportunity to get involved in doing research.At times I got stuck again, his suggestions were of great importance to me.

I am also grateful to Henrik Winkler (TU Berlin, Germany).He was always willing to help me and during our discussions I obtained a better understanding of what I was doing.

In May 2006 I have visited Seppo Hassi (University of Vaasa, Finland).I thank him for his kind hospitality and the interesting discussions we had during the ten days of my visit.I am also grateful for his remarks which have improved the text of this thesis significantly.

I thank Jussi Behrndt (TU Berlin, Germany) for the many discussions we had and for organizing with Henk de Snoo the seminar Boundary tiplets and Weyl functions at the University of Groningen during the spring term of 2006.

Finally, I thank Ineke Kruizinga for taking care of the administrative tasks which came along with finishing this thesis.

Groningen, August 2006, Alef Sterk

(3)

Contents

Preface and acknowledgements i

Contents ii

1 Introduction 1

2 Preliminaries 3

2.1 Nevanlinna families ... 3 2.2 Boundary triplets and Weyl functions ... 4 2.3 Symmetric extensions with compressed Weyl functions ... 5 2.4 Selfadjoint exit space extensions and generalized resolvents . . . 5 2.5 Finite-dimensional relations . . . 7 2.6 Finite-dimensional restrictions of selfadjoint relations ... 8

3 A class of Sturm-Liouville operators 12

3.1 Singular Sturm-Liouville operators ... 12 3.2 Defect numbers: limit point and limit circle ... 14 3.3 The limit point case . . . 16

4 Selfadjoint boundary value problems 19

4.1 Standard boundary conditions ... 19 4.2 Non-standard boundary conditions ... 19 4.3 Examples . . . 21 5 The spectral matrix and generalized Fourier transform 23 5.1 The Kre˘ın-Naimark formula and generalized Fourier transform .23 5.2 The generalized Fourier transform onR . . . 27 5.3 An example . . . 30

A Linear relations in Hilbert spaces 32

Bibliography 33

(4)

Chapter 1

Introduction

Let q be a real-valued, locally integrable function on the interval [0,∞), such that the Sturm-Liouville differential equation

(1.1) −y(x) + q(x)y(x) = λy(x), x∈ [0, ∞),

is in the limit-point case at ∞.This means that for each λ ∈ C \ R the linear subspace ofH = L2(0,∞) of eigensolutions of (1.1) is one-dimensional. With a fundamental system (u1(·, λ) : u2(·, λ)) of (1.1), fixed by the initial conditions

u1(0, λ) = 1, u1(0, λ) = 0, u2(0, λ) = 0, u2(0, λ) = 1.

there exists a function m(λ) defined on C \ R such that for all λ ∈ C \ R the function

ω(·, λ) = u1(·, λ)m(λ) − u2(·, λ)

belongs to H.The standard selfadjoint boundary value problems associated with this equation are determined by boundary conditions of the form

(1.2) y(0) = sy(0), s∈ R ∪ {∞},

with the interpretation that y(0) = 0 for s = ∞.The corresponding spectral theory involving the so-called Titchmarsh-Weyl coefficient m(λ) is well known, see for instance [8], [16], and [22].An extension to λ-depending boundary conditions of the form (1.2) is also well known.

With the underlying Sturm-Liouville operator−D2+q one can also associate boundary value problems different from (1.1) and (1.2). In the present paper it is shown for the case of the Sturm-Liouville operator how the spectral theory for non-standard selfadjoint boundary value problems can be obtained via the abstract extension theory of not necessarily densely defined symmetric operators or relations, in particular, via the theory of boundary triplets and their Weyl functions, as originally worked out by V.A. Derkach and M.M. Malamud, cf.

[12], [13], and [9].

The main idea is to consider non-standard boundary value problems as graph perturbations of a fixed standard boundary value problem determined by the boundary condition (1.2). The use of finite-dimensional subspaces to restrict symmetric relations to enlarge the classes of boundary value problems goes back to [6], [7], and [14].Restricting selfadjoint relations is more general: the symmetric restriction need not be part of the minimal operator associated with the Sturm-Liouville operator.

The spectral theory is given by means of generalized Fourier transforms.The corresponding spectral matrix will be given explicitly in terms of the two Weyl functions associated with the fixed standard boundary value problem and its perturbation.The present work is a continuation of [18] where only canonical extensions of one-dimensional perturbations were considered; now, more gener- ally, exit space extensions of finite-dimensional perturbations will be considered.

(5)

The contents of this report are as follows.Chapter 2 contains some prelim- inary results.Boundary triplets and their Weyl functions are introduced and the Kre˘ın-Naimark formula for generalized resolvents will be stated.Finally, the framework for finite-dimensional graph perturbations of selfadjoint relations will be given in detail.In Chapter 3 the formal Sturm-Liouville operator −D2+ q will be realized as a symmetric operator in the Hilbert space H = L2(0,∞).

Since the Titchmarsh-Weyl coefficient forms the basic motivation for the con- cept of a Weyl function, Weyl’s alternative will be proved in detail along the lines of [8].Chapter 4 shows how non-standard boundary value problems can be obtained by means of graph perturbations and examples will be given.The spectral matrix will be derived in Chapter 5.Furthermore, some properties of the generalized Fourier transform will be given.For the convenience of the reader some facts concerning linear relations in Hilbert spaces will be stated in Appendix A.

A paper based on the results in this report has been submitted, see [19].

(6)

Chapter 2

Preliminaries

2.1 Nevanlinna families

A family of linear relations M (λ), λ ∈ C \ R in a Hilbert space H is called a Nevanlinna family if:

(i) for every λ∈ C+ (resp. λ∈ C) the relation M (λ) is maximal dissipative (resp.accumulative);

(ii) M (λ)= M (¯λ), λ∈ C \ R;

(iii) for some (and hence for all) µ∈ C+(resp.C) the operator family (M (λ) + µ)−1∈ [H]

is holomorphic for all λ∈ C+(resp.C).

If the multi-valued part of M (λ) is nontrivial, then it is independent of λ C \ R, c.f., for instance, [9]. Hence, each Nevanlinna family M(·) can be decom- posed as

M (λ) = Ms(λ)⊕ M, M={0} × mul M(λ), λ ∈ C \ R,

where the orthogonal operator parts Ms(λ) form a Nevanlinna family of densely defined operators in the Hilbert spaceH mul M(λ).

Denote by R(H) the class of all Nevanlinna families in a Hilbert space H.

The following subclasses of R(H) will be useful:

R[H] = { M(·) ∈ R(H) : dom M(λ) = H for all λ ∈ C \ R}, Rs[H] = { M(·) ∈ R[H] : ker Im M(λ) = {0} for all λ ∈ C \ R}, Ru[H] = { M(·) ∈ Rs[H] : 0 ∈ ρ(Im M(λ)) for all λ ∈ C \ R}.

The Nevanlinna families that belong to R[H] are called bounded Nevanlinna functions; the Nevanlinna functions in Rs[H] and Ru[H] are called strict and uniformly strict, respectively.

If M (·) ∈ R[H], then it admits an integral representation of the form M (λ) = A + Bλ +



R

 1

t− λ− t t2+ 1

 dΣ(t),

where A = A ∈ [H], 0 ≤ B = B ∈ [H], and the [H]-valued family Σ(·) is nondecreasing and satisfies



R

1

t2+ 1d(Σ(t)h, h) <∞,

for all h ∈ H.Moreover, the family Σ(·) can be obtained from M(·) via the Stieltjes-Livˇsic formula:

Σ(t2)− Σ(t1) = lim

ε→0

1 π

 t2

t1

Im M (s + iε)ds.

For a proof, see, for instance, [20].

(7)

2.2 Boundary triplets and Weyl functions

Let S be a closed symmetric relation with defect numbers (n, n) in a Hilbert space H with inner product (·, ·).Let Nλ(S) = ker (S− λ) be the defect subspace of S, and define

Nλ(S) :={ fλ={fλ, λfλ} : fλ∈ Nλ}, λ ∈ C, so that Nλ(S)⊂ S.

A boundary triplet Π ={H, Γ0, Γ1} for S consists of an auxilliary Hilbert spaceH with dim H = n and two boundary mappings Γ0 and Γ1 from S toH with the following properties:

(i) the abstract Green’s identity (f, g)− (f, g) =



Γ1f , Γ 0g

H

Γ0f , Γ 1g

H, holds for all elements f ={f, f}, g = {g, g} ∈ S;

(ii) the mapping Γ : f → {Γ0f , Γ 1f} from Sinto H × H is surjective.

When S is densely defined, then the mappings Γ0and Γ1are interpreted as being defined on dom Sinstead on S, i.e., in that case one speaks of Γ0f and Γ1f when f ={f, f} ∈ S.

The mapping f → {Γ0f ,−Γ1f} induces a one-to-one correspondence be- tween the closed extensions H of S which are intermediate (i.e., which satisfy S⊂ H ⊂ S) and the closed linear relations τ inH, via

(2.1) H :={ f ∈ S: 0f ,−Γ1f} ∈ τ },

cf.[3], [12], and [13]. Moreover, (2.1) establishes a one-to-one correspondence between all selfadjoint extensions A = A of S and all selfadjoint relations τ in H.In particular, the special choices τ = {0} × H and τ = H × {0} show that A0:= ker Γ0and A1:= ker Γ1define two selfadjoint extensions of S.

Associated to the boundary triplet Π are two operator functions: the Weyl function M (λ), defined by

(2.2) M (λ) ={ {Γ0fλ, Γ1fλ} : fλ∈ Nλ(S)},

the graph of a bounded linear operator inH, and the γ-field γ(λ) defined by (2.3) γ(λ) ={ {Γ0fλ, fλ} : fλ∈ Nλ(S)},

the graph of a bounded linear operator fromH to Nλ(S).For all λ, µ∈ ρ(A0) the γ-field satisfies the identity

(2.4) γ(λ) = (I + (λ− µ)(A0− λ)−1)γ(µ),

which, in particular, shows that γ(λ) is a holomorphic operator function for λ ∈ C \ R.The Weyl function M(λ) and the γ-field γ(λ) are related via the identity

(2.5) M (λ)− M(µ)

λ− ¯µ = γ(µ)γ(λ),

which holds for all λ, µ ∈ ρ(A0).Moreover, it can be shown that each Weyl function belongs to the class Ru[H], cf.[13].

(8)

2.3 Symmetric extensions with compressed Weyl functions

Let S be a closed symmetric relation with equal defect numbers and let Π = {H, Γ0, Γ1} be a boundary triplet for S.Denote by M (·) and γ(·) the associated Weyl function and γ-field respectively.Assume that the spaceH is decomposed as H = H1⊕ H2 and denote by Pi, i = 1, 2 the orthogonal projection ontoHi. Decompose the Weyl function M (·) as

M (·) = (Mij(·))2i,j=1, where Mij(·) = PiM (·) Hj.

Let τ ={0} × H2, then it is clear that τ=H1× H so that τ is a symmetric relation in H.It follows by (2.1) that the relation H1 defined by

(2.6) H1={ f ∈ S : 0f ,−Γ1f} ∈ τ} = { f ∈ S : Γ0f = P 1Γ1f = 0 } is a symmetric extension of S.The next proposition gives a boundary triplet for H1 such that the corresponding Weyl function coincides with M11(·).

Proposition 2.1. A boundary triplet for H1 is given by Π1={H1, Γr0, P1Γr1} where Γrj = Γj H1 and the γ-field and Weyl function associated with Π1 are given by

γ1(λ) = γ(λ) H1, M1(λ) = P1M (λ) H1. Moreover, the identity ker Γ0= ker Γr0 holds.

The proof can be found in [3], see also [10].

2.4 Selfadjoint exit space extensions and generalized resolvents Let S be a closed symmetric relation with equal defect numbers in a Hilbert space H.A selfadjoint relation A in a larger Hilbert space H ⊕ K is called a selfadjoint exit space extension of S if S ⊂ A.The space K is called the exit space; whenK = {0}, the relation A is called a canonical selfadjoint extension of S.The compression of the resolvent (A− λ)−1 to the Hilbert spaceH defined by

Rλ:= PH(A− λ)−1 H, λ ∈ ρ(A), is called a generalized resolvent of S.

Theorem 2.2. Let S be a closed symmetric relation with equal defect numbers and let Π ={H, Γ0, Γ1} be a boundary triplet for S, and let M (·) and γ(·) be the corresponding Weyl function and γ-field. The generalized resolvents of S are in one-to-one correspondence with the Nevanlinna families τ (·) ∈ R(H) via the Kre˘ın-Naimark formula:

(2.7) Rλ= (A0− λ)−1− γ(λ)(M(λ) + τ(λ))−1γ(¯λ), where A0= ker Γ0.

Conversely, it can be shown that for each generalized resolvent defined by (2.7) there exists a selfadjoint exit space extension Aτ of S such that Rλ= PH(Aτ λ)−1 H.Moreover, Aτ can be chosen to be minimal, i.e.

H ⊕ K = span { (I + (Aτ− λ)−1)H : λ ∈ ρ(Aτ)}.

(9)

In special cases, the exit spaceK and the extensions Aτ can be given explic- itly.The following example gives an explicit formula in the case when τ (·) is a rational Nevanlinna function.

Example 2.3 (Rational Nevanlinna functions, c.f. [2]). Let S be a closed symmetric relation in a Hilbert spaceH and let {H, Γ0, Γ1} be a boundary triplet for S.For j = 0, . . . , m pick operators Aj, Bj ∈ [H] such that Aj = Aj and Bj> 0.Define

τ (λ) = A0+ λB0+

m j=1

Bj1/2(Aj− λ)−1Bj1/2,

then it readily follows that τ (·) is a uniformly strict Nevanlinna function.In particular, it follows that τ (·) is the Weyl function of a symmetric relation H in a Hilbert spaceK, see [9].In this particular example this can be made completely explicit.Let K = Hm+1 and let

H =



















 0 h1

... hm



,



 h0 A1h1

... Amhm













∈ K2, : h0=

m j=1

B0−1/2Bj1/2hj







 .

Clearly, H is a symmetric relation inK and its adjoint is given by the relation

H=



















 h0 h1 ... hm



,





h0

A1h1+ B11/2B0−1/2h0 ...

Amhm+ B1/2m B−1/20 h0













∈ K2 : h0, h0, . . . , hm∈ H







 .

Define the mappings χ0, χ1: H→ H by χ0h0⊕ · · · ⊕ hm:= B0−1/2h0,

χ1h0⊕ · · · ⊕ hm:= A0B0−1/2h0+ B1/20 h0

m j=1

Bj1/2hj.

It easily follows that {H, χ0, χ1} is a boundary triplet for H and the corre- sponding Weyl function is given by τ (·).Finally, the relation

Aτ =

f⊕ h0⊕ · · · ⊕ hm∈ S⊕ H : Γ0f− χ0h = Γ1f + χ 1h = 0

is a selfadjoint exit space extension for S and the generalized resolvent Rλ = PH(Aτ−λ)−1 H is given by (2.7), see [11]. The question whether this realization is minimal will not be treated.

It will be convenient to consider only generalized resolvents determined by Nevanlinna families τ (·) such that mul τ(λ) = {0}.The following observation will be useful.

Lemma 2.4. Let τ (·) ∈ R(H) and let P be the projection from H onto dom τ(λ).

Then the identity

(M (λ) + τ (λ))−1= P (P M (λ)P + τs(λ))−1P holds for any Nevanlinna function M (·) ∈ Ru[H].

(10)

For a proof, consult [21].

In particular, it follows from Lemma 2.4 that the generalized resolvent de- termined by τ (·) can be written as

(2.8) Rλ= (A0− λ)−1− γ(λ)P (P M(λ)P + τs(λ))−1(γ(¯λ)P ).

The right hand side of (2.8) can be interpreted as the generalized resolvent of a symmetric relation H1 extending S.To see this, let H1 = dom τ (λ) and H2= mul τ (λ) so thatH = H1⊕H2.Let H1be a canonical symmetric extension of S induced by τ ={0} × H2 via (2.6). By Proposition 2.1 it follows that the generalized resolvent of H1determined by the Nevanlinna family τs(·) coincides with the right hand side in (2.8). Therefore, by extending the symmetry S in the way indicated, it is always possible to consider only the generalized resolvents induced by Nevanlinna families with a trivial multi-valued part.

The next proposition will be useful in connection with boundary value prob- lems.

Proposition 2.5. Let the assumptions be as in Theorem 2.2, then for every h ∈ H the vector f = Rλh is a solution of the following ‘abstract boundary value problem’ with the paramater τ (·) in the ‘boundary conditions’:

(2.9)

f− λf = h, f = {f, f} ∈ S, 0f ,−Γ1f} ∈ τ(λ).

For a proof, see [12], [13], and [10].

2.5 Finite-dimensional relations

LetH be a Hilbert space and let Z and T be linear relations in H with dim T = dim Z = n <∞.Fix a basis for the n-dimensional spaces T and Z by

T = span{ {e1, e1}, . . . , {en, en} }, and

Z = span{ {ϕ1, ψ1}, . . . , {ϕn, ψn} }.

Lemma 2.6. Define the n× n matrix G by

Gij={ej, ej}, {ϕi, ψi}, i, j = 1, . . . , n.

Then the kernel of G is given by

ker G =



(c1, . . . , cn) ∈ Cn :

n j=1

cj{ej, ej} ∈ Z



. In particular, ker G ={0} if and only if T ∩ Z={0, 0}.

Proof. A vector (c1, . . . , cn)∈ Cn belongs to ker G if and only if

n j=1

{ej, ej}, {ϕi, ψi}cj= 0, i = 1, . . . , n,

or, equivalently, if 

j=1

cj{ej, ej} ∈ Z. This completes the proof.

(11)

2.6 Finite-dimensional restrictions of selfadjoint relations

Let A0= A0be a selfadjoint relation in a Hilbert spaceH and let Z be a finite- dimensional subspace ofH × H such that A0∩ Z = {0, 0}.Define the relation S = A0∩ Z.Clearly, S is a closed symmetric relation inH and the adjoint S is given by the componentwise sum inH × H:

(2.10) S= A0 + Z.

Fix a basis for Z by

Z = span{ {ϕ1, ψ1}, . . . , {ϕn, ψn} },

so that dim Z = n.The defect spacesNλ(S) of S can be expressed in terms of the elements of Z.

Lemma 2.7. The defect space Nλ(S) of the symmetric relation S is spanned by the elements

(2.11) χj(λ) = ϕj+ (A0− λ)−1(λϕj− ψj), j = 1, . . . , n.

In particular, the symmetric relation S has defect numbers (n, n).

Proof. Observe that

j(λ), λχj(λ)} = {ϕj, ψj}

+{(A0− λ)−1(λϕj− ψj), (I + λ(A0− λ)−1)(λϕj− ψj)} (2.12)

which shows thatj(λ), λχj(λ)} ∈ S for all j = 1, . . . , n.Hence, span1(λ), . . . , χn(λ)} ⊂ Nλ(S).

To prove the reverse inclusion, let f ∈ Nλ(S) so that {f, λf} ∈ S.By (2.10) it follows that

{f, λf} = {h, h} + c11, ψ1} + · · · + cnn, ψn}, where{h, h} ∈ A0 and c1, . . . , cn∈ C.Clearly,

h= λh +

n j=1

cj(λϕj− ψj),

so that{h, h} ∈ A0 implies that

h =

n j=1

cj(A0− λ)−1(λϕj− ψj)

Hence,

f =

n j=1

cjj+ (A0− λ)−1(λϕj− ψj)) =

n j=1

cjχj(λ).

This completes the proof.

(12)

Two finite-dimensional subspaces Z and Z of H × H are called equivalent (with respect to A0) if

(2.13) A0∩ Z = A0∩ Z ={0, 0} and A0 + Z = A 0 +  Z.

Obviously, (2.13) gives rise to an equivalence relation on the set of finite- dimensional subspaces inH × H and two equivalent subspaces necessarily must have the same dimension.The next lemma shows that each equivalence class contains at least one symmetric representative.

Lemma 2.8. Let the selfadjoint relation A0 and the n-dimensional space Z satisfy A0∩ Z = {0, 0}. Then there exists an n-dimensional symmetric space Z which is equivalent to Z.

Proof. Since dim A0/S = n there exists an n-dimensional relation T ⊂ A0 such that A0 = S + T and S∩ T = {0, 0}, or, equivalently, T ∩ Z ={0, 0}.Fix a basis for T by

T = span{ {e1, e1}, . . . , {en, en} }.

Define for i = 1, . . . , n the vectors b(i) ∈ Cn of which the components are given by b(i)j = (ϕi, ψj) and let α(i) ∈ Cn be the unique vector such that (i) = b(i), where G is the matrix of Lemma 2.6. Define the elements

(2.14) { ϕi, ψi} = α(i)1 {e1, e1} +· · · + α(i)n {en, en} +i, ψi}.

Clearly, the space Z spanned by the elements in (2.14) is equivalent to Z.There- fore it suffices to check that Z is symmetric.By a straightforward calculation it follows that

{ ϕi, ψi}, { ϕj, ψj} = (Gα(i)− b(i))j− (Gα(j)− b(j))i= 0, for all i, j = 1, . . . , n.This completes the proof.

It follows from Lemma 2.8 that in the definition of S = A0∩ Zwith A0∩ Z = {0, 0} the subspace Z can always be assumed to be symmetric.Clearly, Z is symmetric if and only if A1 := S + Z is a selfadjoint extension of S in which case the selfadjoint extensions A0and A1are transversal, i.e.

S = A0∩ A1, S= A0+ A 1.

By (2.10) it follows that each element{f, f} ∈ S= A0 + Z can be decom- posed as

(2.15) {f, f} = {h, h} + c11, ψ1} + · · · + cnn, ψn},

where {h, h} ∈ A0 and cj ∈ C for all j = 1, . . . , n.Moreover, it follows from the assumption A0∩ Z = {0, 0} that the decompostion in (2.15) is unique.

Define the mappings Γ0, Γ1: S→ Cn by

(2.16) Γ0{f, f} :=

 c1

... cn

 , Γ1{f, f} :=



{h, h}, {ϕ1, ψ1}

...

{h, h}, {ϕn, ψn}

 ,

where{f, f} is as in (2.15).

(13)

Proposition 2.9. Assume that A0∩ Z = {0, 0} and that Z is symmetric. Let S = A0∩ Z and let the mappings Γ0, Γ1: S→ Cn be as in (2.16). T hen:

(i) Π ={Cn, Γ0, Γ1} is a boundary triplet for S= A0 + Z;

(ii) the corresponding γ-field is given by the linear operator fromCn toH by

(2.17) χ(λ) =









 c1

... cn

 ,

n j=1

cjχj(λ)



 : c1, . . . , cn∈ C



,

where the χj(λ) are given by (2.11);

(iii) the corresponding Weyl function is given by the matrix function Q(λ) of which the components are given by

Qij(λ) = (χj(λ), ¯λϕi− ψi), i, j = 1, . . . , n.

Moreover, ker Γ0= A0 and ker Γ1= S + Z.

Proof. (i) Let f ={f, f}, g = {g, g} ∈ S then

{f, f} = {h, h} + c11, ψ1} + · · · + cnn, ψn}, {g, g} = {k, k} + d11, ψ1} + · · · + dnn, ψn}.

It follows from a straightforward calculation that



Γ1f , Γ 0g

Cn

Γ0f , Γ 1g

Cn=

n j=1

dj{h, h}, {ϕj, ψj}

+

n j=1

cj{ϕj, ψj}, {k, k}.

On the other hand (f, g)− (f, g) =

n j=1

dj{h, h}, {ϕj, ψj} +

n j=1

cj{ϕj, ψj}, {k, k}

+

n i,j=1

cidj{ϕi, ψi}, {ϕj, ψj} + {h, h}, {k, k}.

The facts that A0 is selfadjoint and Z is symmetric imply that the last two terms vanish.Hence, the abstract Green’s identity is satisfied.

To show that Γ = 0, Γ1} : S → C2n is surjective it suffices to show that Γ0 and Γ1 are surjective mappings from S toCn, since{h, h} ∈ A0 and (c1, . . . , cn)∈ Cn can be chosen independently in (2.15). Clearly, the map Γ0 is surjective.To show that Γ1 is surjective let (α1, . . . , αn) ∈ Cn belong to (ran Γ1).Then

n j=1

αj{h, h}, {ϕj, ψj} = 0,

(14)

for all{h, h} ∈ A0.This implies that

α11, ψ1} + · · · + αnn, ψn} ∈ A0= A0.

Since A0∩ Z = {0, 0} it follows that αj = 0 for all j = 1, . . . , n so that ran Γ1= Cn.

(ii) & (iii) Every element fλ∈ Nλ can be written as

fλ={fλ, λfλ} = c11(λ), λχ1(λ)} + · · · + cnn(λ), λχn(λ)}, from which it easily follows that

Γ0{fλ, λfλ} =

 c1

... cn

 , Γ1{fλ, λfλ} =



n

j=1cjj(λ), ¯λϕ1− ψ1) ...

n

j=1cjj(λ), ¯λϕn− ψn)



The proof is completed by applying the definitions (2.2) and (2.3).

(15)

Chapter 3

A class of Sturm-Liouville operators

3.1 Singular Sturm-Liouville operators

Consider the following eigenvalue problem on the positive half-axis:

(3.1) (L− λ)y(x) = f(x), x ∈ (0, ∞), λ ∈ C where L is the formal differential expression

L =− d2

dx2 + q(x).

Assume that the function q(x) to be real valued and locally integrable on the in- terval [0,∞).The formal operator L is said to be a singular differential operator with a regular left end point.

Let f be a measurable function on (0,∞).A function y is called a solution of (3.1) if y, y∈ ACloc(0,∞) and (3.1) holds almost everywhere in the interval (0,∞).The eigenvalue problem (3.1) is said to be homogeneous if f = 0;

otherwise it is said to be inhomogeneous.The eigenvalue problem (3.1) together with the Cauchy condition

y(b) = β1, y(b) = β2, 0≤ b < ∞,

has a unique solution; see, for instance, [23].In particular, there exists a funda- mental system{u1(·, λ), u2(·, λ)} of the homogeneous eigenvalue problem, fixed by the following initial conditions:

u1(0, λ) = 1, u1(0, λ) = 0, u2(0, λ) = 0, u2(0, λ) = 1.

The functions u1(·, λ) and u2(·, λ) are continuous as functions in (x, λ) and entire in λ for a fixed x∈ [0, ∞).For a proof of this fact, consult, for instance, [8].

Define the linear space

D := {y ∈ L2(0,∞) : y, y ∈ ACloc(0,∞), Ly ∈ L2(0,∞)}.

For functions y1, y2∈ D we have Green’s identity:

(3.2)

 b

a

(Ly1)(t)y2(t)dt−

 b

a

y1(t)(Ly2)(t)dt = [y1, y2](b)− [y1, y2](a), where

[y1, y2](x) := y1(x)y2(x)− y1(x)y2(x).

Clearly, the limits for a→ 0 and b → ∞ exist in (3.2), so that (3.3) (Ly1, y2)− (y1, Ly2) = [y1, y2](∞) − [y1, y2](0), for all functions y1, y2∈ D.

(16)

The linear spaceD is the largest subspace of H = L2(0,∞) on which the formal operator L is well-defined as an operator inH.Define the linear relation (3.4) Tmax={ {y, Ly} : y ∈ D} ,

so that, in fact, Tmaxis (the graph of) an operator.In general, the right hand side of (3.3) is nonzero, which implies that Tmaxis not a symmetric operator.

Define the subspace

D0:={y ∈ D : y(0) = y(0) = 0, y vanishes outside a bounded interval}.

The following observation will be useful.

Lemma 3.1. Let 0 < b <∞ and let f ∈ L2(0,∞) be zero outside the interval (0, b). Then the equation Ly = f has a solution satisfying the conditions

y(0) = y(0) = y(b) = y(b) = 0

if and only if f is orthogonal to the functions u1(·, 0) and u2(·, 0). In addition, such a solution belongs to D0.

Proof. Let ϕ be the unique solution of

Ly = f, y(b) = y(b) = 0,

in the interval (0, b).Extend ϕ outside the interval (0, b) by setting ϕ equal to zero.Clearly, ϕ∈ D0.By Green’s identity (3.2) it follows that

ϕ(0) =−

 b

0

u1(t, 0)f (t)dt, ϕ(0) =

 b

0

u2(t, 0)f (t)dt.

This completes the proof.

Define the linear relation

(3.5) Tmin:= clos{ {y, Ly} ∈ Tmax : y∈ D0}.

Proposition 3.2. The adjoint of the linear relation Tmin is given by Tmin = Tmax. In particular, the linear relation Tmin is symmetric.

Proof. Define the linear relation

T0:={ {ϕ, Lϕ} ∈ Tmax : ϕ∈ D0},

so that Tmin= clos T0.Clearly, it suffices to show that T0= Tmax. Let{y, Ly} ∈ Tmax, then, by Green’s identity (3.3),

(Ly, ϕ)− (y, Lϕ) = [y, ϕ](∞) − [y, ϕ](0) = 0, for all elements{ϕ, Lϕ} ∈ T0.This proves the inclusion Tmax⊂ T0.

Conversely, let{h, k} ∈ T0.Denote by h0a solution of the equation Ly = k.

Let 0 < b <∞ and let f ∈ L2(0,∞) be such that f is zero outside the interval (0, b) and f is orthogonal to the functions u1(·, 0) and u2(·, 0).By Lemma 3.1 there is a function ϕ∈ D0 such that Lϕ = f and

ϕ(0) = ϕ(0) = ϕ(b) = ϕ(b) = 0.

(17)

Then

(f, h) = (Lϕ, h) = (ϕ, k),

where the last equality follows from the fact that{ϕ, Lϕ} ∈ T0and{h, k} ∈ T0. On the other hand, by Green’s identity (3.3),

(f, h0) = (Lϕ, h0) = (ϕ, Lh0) = (ϕ, k) This implies that

(f, h− h0) = 0.

By choice of the function f it follows that h = h0+ u in the interval (0, b) where u is a solution of Ly = 0.Hence, h, h ∈ ACloc(0, b) and Lh = k almost everywhere in the interval (0, b).Since b is arbitary, it follows that h∈ D and Lh = k almost everywhere.This proves the inclusion T0⊂ Tmaxand completes the proof.

It follows from Proposition 3.2 that Tmin is (the graph of) a densely defined operator.Hence, the operator Tmax, as an extension of Tmin, is also densely defined.In addition, the operator Tmax is closed.

3.2 Defect numbers: limit point and limit circle By definition, the defect numbers of Tmin are given by

n±= dim ker (Tmax− ¯λ), λ ∈ C \ R.

Since the potential q(x) is real valued, it follows that the defect numbers are equal.Indeed, let ϕ be a solution of the equation Ly = λy for λ∈ C \ R.Clearly,

¯

ϕ is a solution of the equation Ly = ¯λy and ϕ belongs to D if and only if ¯ϕ belongs to D.Consequently the defect numbers are equal, i.e.n = n+ = n.

In particular, it follows that for all λ ∈ C \ R there are precisely n linearly independent solutions of the equation Ly = λy that belong to L2(0,∞).

The following theorem gives a description of the defect spaces of Tmin. Theorem 3.3 (Weyl’s alternative). Let λ∈ C \ R. The solution ω(·, λ) = u1(·, λ) m− u2(·, λ) satisfies the boundary condition

(3.6) cos β y(b) + sin β y(b) = 0, 0≤ β < π,

at 0 < b <∞ if and only if m lies on a circle Cb(λ) in the complex plane. Then, independent of λ∈ C \ R, precisely one of the following alternatives holds:

(i) the limit circle case: Cb(λ)→ C(λ), a limit circle with positive radius, as b→ ∞, in which case every solution of Ly = λy belongs to L2(0,∞);

(ii) the limit point case: Cb(λ) → m(λ), a limit point, as b → ∞, in which case only the scalar multiples of the solution

ω(·, λ) = u1(·, λ)m(λ) − u2(·, λ), belong to L2(0,∞).

(18)

Proof. Clearly, the solution ω(·, λ) = u1(·, λ) m− u2(·, λ) satisfies the boundary condition (3.6) if and only if

(3.7) m = cot β u2(b, λ) + u2(b, λ) cot β u1(b, λ) + u1(b, λ).

As λ, b, β vary, m becomes a function of these arguments, i.e. m = m(λ, b, β). If z = cot β and λ, b are held fixed, (3.7) may be written as

(3.8) m = Az + B

Cz + D, where

(3.9) A = u2(b, λ), B = u2(b, λ), C = u1(b, λ), D = u1(b, λ),

are fixed while z varies over the real line if β varies from 0 to π.The image of the real axis under the mapping (3.8) is a circle Cb(λ) in the complex plane.

Therefore, ω satisfies (3.6) if and only ifm lies on the circle C b(λ).

From (3.8) it follows that the image of the real axis, Im z = 0, becomes (3.10) (Dm − B)(C m− A) − (D m− B)(C m− A) = 0,

which is the equation for Cb(λ).A straightforward calculation shows that the center and the radius of Cb(λ) are given by

ξb(λ) = BC− AD

CD− CD, rb(λ) = |AD − BC|

|CD − CD|,

respectively. From (3.9) and (3.10) it follows that the equation for Cb(λ) is

(3.11) [ω, ω](b) = 0,

and that

BC− AD = [u1, u2](b), CD− CD = [u1, u1](b), AD− BC = −1, so that

(3.12) ξb(λ) = [u1, u2](b)

[u1, u1](b), rb(λ) = 1

|[u1, u1](b)|.

Since the coefficient of | m|2 in (3.11) equals [u1u1](b), the interior of Cb(λ) is given by

(3.13) [ω, ω](b)

[u1, u1](b) < 0.

By Green’s identity,

[u1, u1](b) = 2iIm λ

 b

0 |u1(x, λ)|2dx, and

[ω, ω](b) = 2iIm λ

 b

0 |ω(x, λ)|2dx + [ω, ω](0).

(19)

Since [ω, ω](0) =−2iIm m, it follows from (3.11) and (3.13) that a point m is on or inside Cb(λ) if and only if

(3.14)

 b

0 |ω(x, λ)|2dx≤Imm Im λ.

A point m is on C b(λ) if and only if equality holds in (3.14); otherwisem is in the interior of Cb(λ).The radius of Cb(λ) is determined by

(3.15) rb(λ)−1= 2|Im λ|

 b

0 |u1(x, λ)|2dx.

Now let 0 < a < b <∞.Then if m is inside or on Cb(λ)

 a

0 |ω(x, λ)|2dx <

 b

0 |ω(x, λ)|2dx≤ Imm Im λ,

and thereforem is inside C a(λ).This shows that the circles Cb(λ) are nested:

the interior of Ca(λ) contains Cb(λ) for 0 < a < b <∞.Thus for a given λ such that Im λ= 0, as b → ∞ the circles Cb(λ) converge either to a circle C(λ) or to a point m(λ).

In the case that the Cb(λ) converge to a circle C(λ), let m be a point on C; otherwise letm = m(λ), the limit point.In both cases m is inside C b(λ) for all b > 0.Hence,

 b

0 |u1(x, λ)m − u2(x, λ)|2dx < Im m Im λ,

and letting b → ∞ on sees that u1(·, λ) m− u2(·, λ) ∈ L2(0,∞).Therefore, if Im λ= 0, there always exists a solution of Ly = λy that belongs to L2(0,∞).

If the circles Cb(λ) converge to a circle C(λ), then its radius r(λ) = lim rb(λ) is strictly positive, and from (3.15) it follows that u1(·, λ) ∈ L2(0,∞).

In the case that the Cb(λ) converge to a point m(λ), it follows that lim rb(λ) = 0 so that u1∈ L/ 2(0,∞) in this case.This completes the proof.

Theorem 3.3 implies that the defect numbers of Tmin are either (2, 2), the limit circle case, or (1, 1), the limit point case.In the limit circle case the defect space Nλ = ker (Tmax − λ) is spanned by the functions u1(·, λ) and u2(·, λ).In the limit point case the defect space is spanned by the function ω(·, λ) = u1(·, λ)m(λ) − u2(·, λ).

3.3 The limit point case

In all forthcoming sections it will be assumed that the formal operator L is in the limit point case at infinity.For instance, L is in the limit point case when q(x) satisfies the inequality q(x)≥ −kx2where k is a positive constant, see [8].

In particular, for all nonnegative constant functions q(x) the formal operator L is in the limit point case.

Proposition 3.4. Suppose that the operator L is in the limit point case at infinity. Then [y1, y2](∞) = 0 for all y1, y2∈ D. In particular, Green’s identity reduces to

(3.16) (Ly1, y2)− (y1, Ly2) =−[y1, y2](0), for all y1, y2∈ D.

Referenties

GERELATEERDE DOCUMENTEN

De samenstelling van een eindexamen kan men van verschillende kanten bena- deren. Zo kan men trachten na te gaan in welke mate het examen in een vak representatief is voor

anxiety and stress and the neurobehavioural development of the fetus and child: Links and possible mechanisms. Social support and strain from partner, family, and friends: Costs

The field of bioinformatics is very broad and encompasses a wide range of research topics: sequence analysis, data analysis of vast numbers of experimental data (high

In this paper we have proposed a frequency-domain per-tone equal- izer (PTEQ) for OFDM transmission over doubly selective channels with IQ-imbalance and CFO (viewed as part of

Afbeelding sonar mozaïek van het onderzoeksgebied samengesteld met het Isis programma (Triton Elics) met bewerkingen van het Side-scan sonar mozaïek in Matlab; rood codering van

[r]

In Molenaar et al. [72] it was found, for forced 2D flow on square domain with no-slip boundaries, that a different forcing protocol results in a continuous cycle of

As far as we know, the relation between the spectral radius and diameter has so far been investigated by few others: Guo and Shao [7] determined the trees with largest spectral