Sturm-Liouville problems with boundary conditions rationally dependent on the eigenparameter, I ∗ Paul A. Binding † Department of Mathematics and Statistics University of Calgary Calgary, Alberta, Canada T2N 1N4 Patrick J. Browne † Department of Mathematics and Statistics University of Saskatchewan Saskatoon, Saskatchewan, Canada S7N 5E6 Bruce A. Watson ‡ Department of Mathematics University of the Witwatersrand Private Bag 3, P O WITS 2050, South Africa September 29, 2008

Abstract We consider the Sturm-Liouville equation −y ′′ + qy = λy on [0, 1], subject to the boundary conditions y(0) cos α = y ′ (0) sin α, α ∈ [0, π), ∗

Keywords: Sturm-Liouville, eigenparameter dependent boundary conditions, Mathematics subject classification (2000): 34B24, 34L20. † Research supported in part by grants from the NSERC of Canada ‡ Research conducted while visiting University of Calgary and University of Saskatchewan and supported in part by the Centre for Applicable Analysis and Number Theory.

1

and N

X bk y′ (1) = aλ + b − . y λ − ck k=1

Topics treated include existence and asymptotics of eigenvalues, oscillation of eigenfunctions, and transformations between such problems.

1

Introduction

We consider the regular Sturm-Liouville equation −y ′′ + qy = λy on [0, 1],

(1.1)

subject to the boundary conditions y(0) cos α = y ′ (0) sin α, α ∈ [0, π),

(1.2)

y′ (1) = f (λ), y

(1.3)

and

for a class of functions f . The case when f is constant is the “standard” Sturm-Liouville problem on which there is a vast literature. Cases where f is affine or bilinear make up the majority of the literature on so-called “eigenvalue dependent boundary conditions” and we refer to [15, 26] and the many references there for some of this activity. Alternative settings have also been studied, e.g. [14, 16, 17] for singular equations; [9, 12, 22] for higher order (and matrix) equations; and [6, 19, 27] for partial differential equations. There have also been several investigations of (1.1)-(1.3) where f is a more general (usually rational) function, e.g., [3, 21, 23, 25] and their references. Our study is the first of two parts on problems involving a particular class RN of rational f of the form f (λ) = aλ + b −

N X k=1

bk λ − ck

(1.4)

admitting a rather rich spectral theory. In fact these f will also belong to a class usually associated with the names of Herglotz or Nevanlinna. Boundary value problems as above involving functions f of this (and more general) type have been analysed in e.g., [10, 11, 13, 24]. These papers (and those cited earlier) have focussed mainly on operator-theoretic formulations in Hilbert, Pontryagin or Krein spaces, usually leading to expansion theorems (see also the second part of our study, where further attention to

2

this aspect will be given). Here we shall use differential equation techniques to derive properties of the eigenvalues and eigenfunctions generalizing classical Sturm theory. A key tool in our analysis will be a modification of a transformation which was first used to our knowledge by Darboux (see [18, p. 132]), and which has subsequently been explored in [2, 7, 8]. These papers focus on transformation of the differential equation (1.1), but for us the effect on the boundary conditions will be crucial. Moreover if an original boundary condition is Dirichlet, then the cited works produce a transformed problem which is singular, whereas we ensure regularity. This is an important difference since non-Dirichlet conditions transform to Dirichlet, and repeated transformations will 0 be needed. It turns out that each class RN is the union of two subclasses R+ N and RN and our transformation will provide direct links between these subclasses for various values of N . This fact and the connections with Herglotz-Nevanlinna functions are derived in Section 2. Section 3 contains an analysis of existence, oscillation and comparison theory, mainly via Pr¨ ufer methods. In contrast with the usual Sturm theorem which gives one eigenvalue per oscillation count, here N “extra” eigenvalues appear with arbitrary oscillation counts (for example, all could be equal). In Section 4 we show that if the transformation of the previous paragraph is applied to (1.1)-(1.3), then the new spectrum contains the old eigenvalues (except possibly the first one). Using the oscillation theory of Section 3, we show that these are the only eigenvalues of the new problem, so the transformation is isospectral (with the exception above). The new problem is “simpler” than the original one, and after at most 2N + 1 transformations we eventually produce a standard problem, i.e., with constant boundary conditions (Corollary 4.2). In Section 5, we discuss eigenvalue asymptotics, again via repeated transformations. An alternative approach, based on the “asymptotic problem” given in [5] for a special case, is also considered.

2

Preliminaries

We consider the class RN of rational functions f as in (1.4) where all the coefficients are real and a ≥ 0, bk > 0 and c1 < c2 < ... < cN , N ≥ 0. Recall that f : C → C is a Herglotz-Nevanlinna function if f (z) = f (z) and f maps the closed upper half plane into itself. Lemma 2.1 A rational function f with simple real poles is a Herglotz-Nevanlinna function if and only if f ∈ RN for some N . Proof: If f ∈ RN then each summand is evidently Herglotz-Nevanlinna, and hence so is f . Conversely, simplicity of the poles allows a partial fraction expansion of the form

3

P bk f (λ) = p(λ)+aλ+b− N k=1 λ−ck , where p is a polynomial whose terms all have degree at least two, and the remaining summands contain coefficients unrestricted in sign. Since f maps the upper half plane into itself, we can let λ tend to ∞ along appropriate rays to see that p must vanish and a must be positive. Finally letting λ tend to ck we obtain positivity of bk . 0 Let R+ N (resp. RN ) denote the subclass of RN for which a > 0, (resp. a = 0). The following properties are easily established.

Lemma 2.2 Let f ∈ RN . Then (i) f ′ (λ) > 0 for each real λ where f (λ) is finite, (ii) lim f (λ) = ∓∞, λ→ck ±

(iii) if f ∈ R+ N , then

lim f (λ) = ±∞, while if f ∈ R0N , then f (λ) → b from below

λ→±∞

(resp. above) as λ → ∞ (resp. −∞).

The graph of a typical member of RN is shown in Fig.1.

f (λ) 6

h

c1

c2

c3



Figure 1: f (λ) Given a function f ∈ RN , and a constant µ < c1 , we define F (λ) =

µ−λ − f (µ) f (λ) − f (µ)

(2.1)

extending this definition by continuity where possible, so F (ck ) = −f (µ), 1 ≤ k ≤ N , and F (µ) = −f ′ (µ)−1 − f (µ). The principal result of this section shows, in particular,

4

that if f ∈ RN then F ∈ RM , i.e., F (λ) = Aλ + B −

M X k=1

Bk , λ − Ck

(2.2)

where M is N − 1 or N depending on the value of a. Theorem 2.3 In the notation above 0 (i) if f ∈ R+ N then F ∈ RN and µ < c1 < C1 < c2 < ... < cN < CN , (ii) if f ∈ R0N then F ∈ R+ N −1 and µ < c1 < C1 < c2 < ... < CN −1 < cN . Proof: We calculate F (λ) = − =

QN

k=1 (λ

− ck ) − f (µ) r(λ)

p(λ) r(λ)

(2.3) (2.4)

where p(λ) is a polynomial of degree at most N , r(λ) = a

N Y

(λ − ck ) +

N X

ek

(λ − cj )

(2.5)

j=1,j6=k

k=1

k=1

N Y

is a polynomial of degree N or N −1 according as a > 0 or a = 0 and ek = bk /(µ−ck ) < 0. It is easy to check that sgn(r(ck )) = (−1)N −k+1 ,

k = 1, ..., N,

and thus r has roots µk ∈ (ck , ck+1 ), k = 1, ..., N − 1. When a = 0, these must be all the roots of r since its degree is N − 1. When a > 0, we see that r(cN ) < 0 and that r has a > 0 as its leading coefficient. Hence there is an additional root µN > cN , and now the µk , k = 1, ..., N , form all the roots of r. All of these roots are simple so the rational function F has simple poles comprising µ, µk , k = 1, ...N, interlacing the poles of f in the manner described. From this it readily follows that F can be expressed in the form (2.2) where A = 0 and P −1 N M = N when a > 0, and, when a = 0, M = N − 1 and A = − e > 0 by k=1 k (2.5). The poles of F are precisely the roots of f (λ) = f (µ) (except λ = µ) so the basic properties of f show that f (λ) − f (µ) → 0 ±

as λ → Ck ± .

Thus the original definition of F gives lim F (λ) = ∓∞

λ→Ck ±

5

which forces Bk > 0, k = 1, ..., M. Remark 1 By composing these transformations we can map RN into R00 which will enable us to convert eigenvalue problems with boundary conditions in RN into standard problems with boundary conditions independent of λ. Remark 2 Routine algebraic calculations can be used to give A, B etc. in terms of a, b etc.. For example f ∈ R+ N f ∈ R0N

1 =⇒ A = 0, B = − − f (µ), a !−1 N X . ek =⇒ A = −

(2.6) (2.7)

k=1

3

Eigenvalues: existence and oscillation theory

We consider here the existence of eigenvalues and the associated oscillation theory for the problem (1.1)-(1.3) where f ∈ RN . If α = 0, we interpret (1.2) as y(0) = 0. Our approach will be via Pr¨ ufer theory and to this end for a given λ, we consider the solution y(λ, x) of (1.1), (1.2) and define θ(λ, x) via the initial value problem θ ′ = cos2 θ + (λ − q) sin2 θ, which leads to cot θ(λ, x) =

θ(λ, 0) = α

(3.1)

y ′ (λ, x) . y (λ, x)

In particular the eigencondition (1.3) becomes cot θ(λ, 1) = f (λ).

(3.2)

Standard properties of the Pr¨ ufer angle θ (for example, that for a given x, θ(λ, x) is continuous and increasing in λ, and that limλ→−∞ θ (λ, 1) = 0 and limλ→+∞ θ (λ, 1) = ∞) can be found in, e.g., [2]. The graph of cot θ(λ, 1) is displayed in Figure 2. Geometrically, the real eigenvalues of (1.1)-(1.3) correspond to the λ-values at which the graphs of cot θ(λ, 1) and f (λ) intersect. (If these graphs share a common vertical asymptote ck say, then λ = ck is an eigenvalue for the terminal condition y(1) = 0). Since cot θ(λ, 1) and f (λ) are respectively decreasing and increasing on each branch, there is a sequence of simple intersections of these two graphs. More precisely, we define y(x, λ) to be a nonzero solution of (1.1)-(1.2), analytic in λ ∈ C, and we write ω(λ) = y ′ (1, λ) − f (λ)y(1, λ). By definition, λ is an eigenvalue of (1.1)(1.3) if ω(λ) = 0 and we call λ a simple eigenvalue if in addition ωλ (λ) 6= 0, where suffix

6

y′ (1) y

6

λD 0

λD 1

λD 2

h



Figure 2: cot θ(λ, 1) denotes differentiation. To discuss the poles of f we use Ω(λ) = y(1, λ) − y ′ (1, λ)/f (λ) instead of ω(λ). Specifically, λ = ck is an eigenvalue if y(1, λ) = 0, and is a simple eigenvalue if, in addition, yλ (1, λ) 6= 0.

Theorem 3.1 (i) The eigenvalues of (1.1)-(1.3) are real, simple and form a sequence λ0 < λ1 < ... accumulating only at ∞ and with λ0 < c1 . (ii) If b is decreased and ck , q are increased then each λj is increased. (iii) If a > 0 is decreased and bk is increased then each positive λj > ck is increased.

Proof: (i) Suppose λ is a nonreal eigenvalue. Then (1.1)-(1.3) hold with y 6= 0, and also −y ′′ + qy = λy,

(3.3)

(y ′ /y)(0) = cot α and (y ′ /y)(1) = f (λ) = f (λ) since f is Herglotz-Nevanlinna. Thus λ is also an eigenvalue and without loss we shall assume Imλ > 0. Now by the standard procedure of multiplying (1.1) by y, (3.3) by y, integrating and subtracting, we obtain Z 1 ′ ′ (−y y + yy )(1) = (λ − λ) |y|2 0

7

which we shall call Lagrange’s formula for (1.1) and (3.3). The right side is positive imaginary but the left side is (f (λ) − f (λ))|y(1)|2 which is negative imaginary since f is Herglotz-Nevanlinna. This contradiction establishes reality of the eigenvalues. To prove simplicity, suppose ω(λ) = ωλ (λ) = 0 for some λ, which by the above reasoning we take to be real, and hence without loss we also assume y to be real. From (1.1)-(1.3) we obtain −yλ′′ + qyλ = λyλ + y,

(3.4)

(yλ′ /yλ )(0) = cot α and yλ′ yλ (1) = f ′ (λ) + f (λ) (1). y y By Lagrange’s formula for (1.1) and (3.4) we have Z ′ ′ (−y yλ + yyλ )(1) = −

(3.5)

1

y2.

(3.6)

0

The right side is negative, whereas by (3.5) the left side is   yλ yλ 2 ′ y(1) −f (λ) (1) + f (λ) + f (λ) (1) = y(1)2 f ′ (λ) y y which is positive by Lemma 2.2(i), contradiction. If λ = ck is a non-simple eigenvalue, then y(ck ) = yλ (1, ck ) = 0 so the left side of (3.6) vanishes but the right side is negative, and again we have a contradiction. The final contention follows from the geometry of the cot θ and f graphs. (ii) We note that a decrease in b or an increase in any of ck , 1 ≤ k ≤ N, causes f (λ) to decrease, while an increase in q causes θ(λ, 1) to increase (see (3.2)). The net effect is an increase in each λj . The proof of (iii) is similar. Closer examination of the graphs reveals some interesting interlacing relationships which we shall now explore. We define λD i , i = 0, 1, ..., to be the eigenvalues for the standard Sturm-Liouville problem consisting of (1.1)-(1.2) and the Dirichlet condition y(1) = 0. We also use the notation λcD to denote the sequence consisting of all cj and λD i k in for some j and k). Let k nondecreasing order (counted by multiplicity if cj = λD j be k the number of points ci ≤ λj . Analysis of the f and cot θ graphs now yields Theorem 3.2 (i) The λi interlace the λcD j in the sense that cD cD λ0 < λ0 ≤ λ1 ≤ λ1 ≤ ... .

8

(ii) Where defined, ckj ≤ λj < ckj +1 . D D (iii) Setting λD −1 = −∞ we have λj−1−kj < λj ≤ λj−kj , for all j ≥ 0.

f (λ)

y′ (1) y

cot θ(λ, 1)

6

h

c1

c2 λD 0

c3 λD 1

λD 2



Figure 3: f (λ) and cot θ(λ, 1) We can now deduce the oscillation properties of the eigenfunctions. Note that if λ ∈ D (λD i−1 , λi ], then a solution of (1.1)-(1.2) has i zeros in (0, 1). Thus the oscillation number (i.e. the number of zeros in (0, 1) of an eigenfunction) associated with such an eigenvalue D can be found by determining the interval (λD i−1 , λi ] into which it falls. Corollary 3.3 Let ωj be the oscillation number associated with λj . Then ωj = j − kj and in particular, ω0 = 0 and ωj = j − N when λj > cN .

4

Transformations between problems

The aim of this section is to describe a transformation, with certain eigenvalue preserving properties, from a problem of the type (1.1)-(1.3) to a “simpler” one with a new potential qˆ in place of q and with f (λ) replaced by F (λ) constructed in Section 2. In this context “simpler” means that F has either fewer terms or fewer poles than f . By iteration of the procedure one can transform (1.1)-(1.3) to a standard Sturm-Liouville problem whose eigenvalues are those of the original problem except for finitely many. For brevity we denote the oscillation number of a function y on (0, 1) by osc(y).

9

Theorem 4.1 For the problem (1.1)-(1.3) suppose the eigenvalues are λ0 < λ1 < ... with corresponding eigenfunctions y0 , y1 , ... . Let µ = λ0 if α > 0 µ < λ0 if α = 0, and let w be the solution of (1.1) with λ = µ, w(1) = 1 and w′ (1) = f (µ). Define z=

w′ and qˆ = q − 2z ′ . w

Let F (λ) be as in (2.1) and define γ ∈ [0, π) by cot γ = −z(0) if α = 0 and γ = 0 if α > 0. Then the eigenvalues of the problem −y ′′ + qˆy = λy, y′ (0) = cot γ y  ′ y (1) = F (λ) y

(4.1) (4.2) (4.3)

are λj , with corresponding eigenfunctions uj = yj ′ − zyj , where j ≥ 1 for α > 0 and j ≥ 0 for α = 0.

Proof: We begin by showing that w has no zeros in [0, 1]. This is clear when α = 0, for ′ ′ then w = y0 which has no zeros in (0, 1), and moreover ww (0) = cot α and ww (1) = f (λ0 ), both of which are finite. When α = 0, we note that y0 (0) = 0 but y0 has no zeros in (0, 1) and y0 (1) 6= 0 ufer differential equation (3.1) with λ = µ < λ0 since λ0 < λD 0 . Now consider the Pr¨ and subject to the terminal condition θ(µ, 1) = cot−1 f (λ0 ) ∈ (0, π). Since µ < λ0 , we obtain a positive initial value θ(µ, 0) ∈ (0, π). (Recall that θ(µ, x) cannot decrease through multiples of π as x decreases.) If we now replace the terminal condition by θ(µ, 1) = cot−1 f (µ) > cot−1 f (λ0 ) we see that θ(µ, 0) increases again but must remain within (0, π) provided we choose the terminal angle to be less than π. Thus the solution ψ(x) of the Pr¨ ufer equation with λ = µ and terminal condition ψ(1) = cot−1 f (µ) takes values only in (0, π) for all x ∈ [0, 1]. This establishes that w has no zeros in [0, 1], and consequently z and γ as given in the statement are well defined. For the case α > 0 we shall define ψ(x) = θ(λ0 , z). Thus  α α>0 ψ(0) = π − γ α = 0.

10

Throughout the proof when we refer to j, we mean j ≥ 1 if α > 0 and j ≥ 0 for α = 0. As in [4], direct calculation shows that the functions uj satisfy (4.1) and (4.2) and further that u′j µ − λj = y′ − z on [0, 1], uj j − z yj

(4.4)

so uj also satisfies (4.3). Thus the λj are indeed eigenvalues for (4.1)-(4.3) and it remains to show that they constitute all the eigenvalues. Let φ(λ, x) be the Pr¨ ufer angle generated by (4.1) with φ(λ, 0) = γ as in (4.2). In a similar manner let θ(λ, x) be generated by (1.1) with θ(λ, 0) = α as in (1.2). Note that u′j uj (x)

y′

= cot φ(λj , x), yjj (x) = cot θ(λj , x) and z(x) = cot ψ(x). Thus we see from (4.4)

that cot φ(λj , x) =

µ − λj − cot ψ(x). cot θ(λj , x) − cot ψ(x)

(4.5)

Recall that ψ(x) ∈ (0, π) for all x ∈ [0, 1]. Since λj > µ when j ≥ 1, we have θ(λj , x) > ψ(x) on (0, 1] if α > 0, while 0 = θ(λj , 0) < ψ(0) for α = 0. Further, we see from the Pr¨ ufer equation for θ that if cot θ(λj , x) = cot ψ(x) for some d (θ(λj , x) − ψ(x)) > 0 from (3.1) since sin2 ψ(x) > 0. Thus θ(λj , x) − ψ(x) x, then dx increases through multiples of π. From (4.4) it follows that uj has a zero at x ∈ (0, 1) (or equivalently, cot φ(λj , x) is undefined) if and only if θ(λj , x) = ψ(x) + mπ for some positive (resp. nonnegative) integer m, for α > 0 (resp. α = 0). Thus, since φ(λj , x) increases through multiples of π, osc(uj )= n if and only if nπ < θ(λj , 1) − ψ(1) ≤ (n + 1)π if α > 0, (n − 1)π < θ(λj , 1) − ψ(1) ≤ nπ if α = 0. D By Theorem 3.2 (iii), λj ∈ (λD i−1 , λi ] where i = j − kj and so iπ < θ(λj , 1) ≤ (i + 1)π from which we obtain (i − 1)π < θ(λj , 1) − ψ(1) < (i + 1)π. Thus osc(uj ) is i − 1 or i, for α > 0, and i or i + 1, for α = 0, according to whether θ(λj , 1) − ψ(1) ≤ iπ or θ(λj , 1) − ψ(1) > iπ. In terms of f and cot θ(λ, 1), (recall that f (µ) = z(1) = cot ψ(1)), we see that osc(uj ) is i − 1 or i, for α > 0, and i or i + 1, for α = 0, according to whether f (µ) ≤ cot θ(λj , 1) or f (µ) > cot θ(λj , 1).

We recall the expressions for f (λ) and F (λ) from Section 2: f (λ) = aλ + b −

N X k=1

bk , λ − ck

F (λ) = Aλ + B −

M X k=1

Bk λ − Ck

where A = 0, M = N if a > 0 and A > 0, M = N − 1 if a = 0. Recall also that the Ck are the solutions other than λ = µ of f (λ) = f (µ).

11

Now when a > 0, we have for j large enough to ensure λj > CN > cN ,  ′ yj (1) = cot θ(λj , 1) = f (λj ) > f (CN ) = f (µ) yj so osc(uj ) = osc(yj ) − 1, which equals j − N − 1 by Corollary 3.3, for α > 0. Similarly for α = 0, osc(uj ) = osc(yj ) = j − N . On the other hand if we regard λj as the kth eigenvalue (k ≥ 0) of (4.1)-(4.3) listed in increasing order, Corollary 3.3 shows that the oscillation number associated with λj is k − N , so k = j − 1 for α > 0, and k = j for α = 0. Hence the λj , j ≥ 1 for α > 0 and j ≥ 0 for α = 0, constitute all the eigenvalues of (4.1)-(4.3). When a = 0, we have for j large enough to ensure λj > cN ,  ′ yj (1) = cot θ(λj , 1) = f (λj ) < b < f (µ) yj showing that osc(uj ) = osc(yj ) = j − N , for α > 0, while for α = 0, osc(uj ) = osc(yj ) + 1 = j − N + 1. The argument concludes as before since in this case M = N − 1. By iterating these constructions, we arrive at Corollary 4.2 Given (1.1)-(1.3) with f ∈ RN , there exists a Sturm-Liouville problem ′ −y ′′ + q0 y = µy subject to yy (j) = cot αj , j = 0, 1 with eigenvalues µ0 < µ1 < ..., where (i) α0 = 0 and µj = λj+N +1 if α > 0, f ∈ R+ N (ii) α0 = 0 and µj = λj+N if α = 0, f ∈ R0N (iii) α0 > 0 and µj = λj+N otherwise.

5

Asymptotics of eigenvalues

Our aim in this section is to show how eigenvalue asymptotics for the problem (1.1)(1.3) can be derived from those for “standard” Sturm-Liouville problems (with constant boundary conditions). It is well known that asymptotics for the latter are available to any order, depending on the smoothness of q, cf. [20, p. 23]. For q ∈ L1 , we have the following well known estimates (see [5, Theorem A3], for example). Theorem 5.1 If λ′j denote the eigenvalues of the problem (1.1) with ρ = y ′ /y constant at 0 and 1, then λ′j = j 2 π 2 + o(j 2 ) (5.1) R1 as j → ∞. In fact λ′j = (j + D)2 π 2 + 0 q − 2[ρ∗ ]10 + o(1/j) where D is half the number of Dirichlet conditions specified, ρ∗ = ρ if ρ is finite, and ρ∗ = 0 otherwise.

12

Successive applications of the transformations of Section 4, as in Corollary 4.2, lead to Corollary 5.2 For the problem (1.1)-(1.3), (5.1) holds with λ′j replaced by λj . Higher order asymptotics may be obtained via the same technique, although since one has to keep track of the boundary terms at each stage, an inductive argument is appropriate as below. Theorem 5.3 (i) If f ∈ R0N , then, as j → ∞, ( R1 (j − N )2 π 2 + 0 q − 2b + 2 cot α + o(1/j), α 6= 0 R1 λj = (j + 1/2 − N )2 π 2 + 0 q − 2b + o(1/j), α = 0.

(ii) If f ∈ R+ N , then as j → ∞, ( R1 (j − 1/2 − N )2 π 2 + 0 q + 2 cot α + 2/a + o(1/j), α 6= 0 R1 λj = (j − N )2 π 2 + 0 q + 2/a + o(1/j), α = 0.

(5.2)

(5.3)

Proof: When N = 0, (i) and (ii) follow from Theorem 5.1 and [5], so we can assume that (i) and (ii) hold for 0, 1, ..., N . (i) Suppose α > 0 and f ∈ R0N +1 . Then F ∈ R+ N and denoting the other transformed quantities by carets, we have qˆ = q − 2z ′ and α ˆ = 0. By (5.3) and the inductive hypothesis, Z 1

λj+1 = (j − N )2 π 2 +

qˆ + 2/ˆ a + o(1/j)

0

so by (2.7)

λj

2 2

= (j − (N + 1)) π +

Z

1

q − 2[f (λ0 ) − cot α + Σek ] + o(1/j)

0

2 2

= (j − (N + 1)) π +

Z

1

q − 2b + 2 cot α + o(1/j)

0

as µ = λ0 . This establishes (5.2) for N + 1 and α > 0. Now suppose α = 0 instead. Then α ˆ > 0 and we use (5.3) and (2.7) to obtain  2 Z 1 1 2 λj = j− −N π + qˆ + 2 cot α ˆ + 2/ˆ a + o(1/j) 2 0 2  Z 1 1 2 q − 2[f (µ) − z(0) + z(0) + Σek ] + o(1/j) = j + − (N + 1) π + 2 0  2 Z 1 1 2 = j + − (N + 1) π + q − 2b + o(1/j) 2 0

13

which establishes (5.2) for N + 1 and α = 0. 0 (ii) Consider the case when f ∈ R+ N +1 . Then F ∈ RN +1 and we may apply the results just established in (i). When α > 0 we have α ˆ = 0 and we use (5.2) with N + 1 in place of N to give  2 Z 1 1 2 λj+1 = j + − (N + 1) π + qˆ − 2ˆb + o(1/j) 2 0

so by (2.6)  2 Z 1 1 λj = j − − (N + 1) π 2 + q − 2[f (λ0 ) − cot α − (a−1 + f (λ0 ))] + o(1/j) 2 0 2  Z 1 1 q + 2 cot α + 2/a + o(1/j) = j − − (N + 1) π 2 + 2 0

thus establishing (5.3) for N + 1 and α > 0. Finally, when α = 0 we use (5.2) and (2.6) to see that Z 1 2 2 qˆ − 2ˆb + 2 cot α ˆ + o(1/j) λj = (j − (N + 1)) π + 0 Z 1 2 2 q − 2[f (µ) − z(0) − (a−1 + f (µ)) + z(0)] + o(1/j) = (j − (N + 1)) π + 0 Z 1 = (j − (N + 1))2 π 2 + q + 2/a + o(1/j) 0

as in (5.3). An alternative approach, at least to this order, can be given via a simpler “asymptotic” problem whose eigenvalues are ultimately close (but not equal) to those of (1.1)-(1.3). We define this asymptotic problem by (1.1), (1.2) and the “asymptotic boundary condition”  ′ y (1) = b if f ∈ R0N y y(1) = 0 if f ∈ R+ N. A We label the eigenvalues of this problem by oscillation count as λA 0 < λ1 < .... The first A D result concerns the interlacing of the sequences λj , λj and λj .

Theorem 5.4 For the problem (1.1)-(1.3), let j be large enough to ensure λj > cN . D A (i) If f ∈ R0N then λA j−N < λj ≤ λj−N < λj−N +1 . A A (ii) If f ∈ R+ N , then λj−N −1 < λj < λj−N . Proof: (i) is a simple consequence of the superpositioning of the graphs of f (λ) and D cot θ(λ, 1). (ii) follows from Theorem 3.2 (iii) and λA j = λj . D Sharper estimates involving λA j and λj can be given as follows.

14

Theorem 5.5 For the problem (1.1)-(1.3), 2 (i) if f ∈ R0N , then λj − λA j−N = O(1/j ) D 2 (ii) if f ∈ R+ N , then λj − λj−N −1 = 2/a + O(1/j ).  D Proof: Let g(λ) = aλ + b and denote by µn the solution in λD n−1 , λn of cot θ (λ, 1) = g (λ) . Note that f (λ) < g(λ) for λ >> 1 so λn+N > µn for n >> 1. We claim that λn+N − µn = O(1/n2 ) as n → ∞. To this end, consider cot θ(λn+N , 1) − cot θ(µn , 1) = f (λn+N ) − g(µn ) = a(λn+N − µn ) + O(1/n2 ) using Theorem 3.2(iii). From this it follows that, for some ξn ∈ (µn , λn+N ), d cot(θ(ξn , 1)(λn+N − µn ) = a(λn+N − µn ) + O(1/n2 ), dλ O(1/n2 ) . λn+N − µn = d a − dλ cot(θ(ξn , 1) d cot(θ(ξn , 1) ≤ −ǫ < 0 is proven in [5, p. 63, eqn (3.8) et seq.] and this establishes Now dλ our claim.

From this (i) follows directly, and for (ii) we refer to [5, Theorem 5.3] which gives 2 −2 µ n − λA n−1 = a + O(n ) for the case at hand. Theorem 5.3 is now a simple consequence of this result and Theorem 5.1. Higher order asymptotics could in principle be derived by this technique also, but one would need more accurate Taylor expansions of f (λ) and cot θ(λ, 1).

References ` Adler, A modification of Crum’s method, Theoret. Math. Phys., 101 (1994), [1] V. E. 1381-1386. [2] F. V. Atkinson,Discrete and Continuous Boundary Problems, (Academic Press, 1964). [3] A. I. Benedek, R. Panzone, On Sturm-Liouville problems with the square root of the eigenvalue parameter contained in the boundary conditions, Notas de Algebra y Anal., 10 (1981), 1-62. [4] P. A. Binding, P. J. Browne, B. A. Watson, Transformations between SturmLiouville problems with eigenvalue dependent and independent boundary conditions, Bull. London Math. Soc., to appear.

15

[5] P. A. Binding, P. J. Browne, K. Seddighi, Sturm-Liouville problems with eigenparameter dependent boundary conditions, Proc. Edinburgh Math. Soc., 37 (1993), 57-72. [6] P. A. Binding, R. Hryniv, H. Langer, B. Najman, Elliptic eigenvalue problems with eigenparameter dependent boundary conditions, J. Diff. Eq., 174 (2001), 30-54. [7] M. M. Crum, Associated Sturm-Liouville systems, Quart. J. Math. Oxford, 6 (1955), 121-127. [8] P. Deift, Applications of a commutation formula, Duke Math. J., 45 (1978), 267310. [9] A. Dijksma, Eigenfunction expansions for a class of J-selfadjoint ordinary differential operators with boundary conditions containing the eigenvalue parameter, Proc. Roy. Soc. Edinburgh, 86A (1980), 1-27. [10] A. Dijksma, H. Langer, Operator theory and ordinary differential operators, Fields Inst. Monographs 3, (1996), 75-139. [11] A. Dijksma, H. Langer, H. de Snoo, Symmetric Sturm-Liouville operators with eigenvalue depending boundary conditions, Canadian Math. Soc. Conf. Proc. 8 (1987), 87-116. [12] W. Eberhard, G. Freiling, A. Schneider, Note on a paper of E. M. E. Zayed and S. F. M. Ibrahim: “Eigenfunction expansion for a regular fourth order eigenvalue problem with eigenvalue parameter in the boundary conditions” [Internat. J. Math. Math. Sci. 12 (1989), no. 2, 341–348; MR 90h:34043]., Internat. J. Math. Math. Sci., 15 (1992), 809–811. ` [13] A. E. Etkin, Some boundary value problems with a spectral parameter in the boundary conditions, Amer. Math. Soc. Transl. Series 2, 136 (1987), 35-41. [14] C. T. Fulton, Singular eigenvalue problems with eigenvalue-parameter contained in the boundary conditions, Proc. Edinburgh Math. Soc., 87A (1980), 1-34. [15] C. T. Fulton, Two-point boundary value problems with eigenvalue parameter contained in the boundary conditions, Proc. Roy. Soc. Edinburgh, 77A (1977), 293-308. [16] D. B. Hinton, Eigenfunction expansions for a singular eigenvalue problem with eigenparameter in the boundary conditions, SIAM J. Math. Anal., 12 (1981), 572584. [17] D. B. Hinton, J. K. Shaw, Differential operators with spectral parameter incompletely in the boundary conditions, Funkcialaj Ekvacioj, 33 (1990), 363-385. [18] E. L. Ince, Ordinary Differential Equations (Dover, New York, 1956). [19] A. Kozhevnikov, S. Yakubov, On operators generated by elliptic boundary problems with a spectral parameter in boundary conditions, Integral Eq. Oper. Theory, 23 (1995), 205-231. [20] B. M. Levitan, M. G. Gasymov, Determination of a differential equation from two of its spectra, Russian Math. Surveys, 19 (1964), 1-63.

16

[21] E. M. Russakovskii, Operator treatment of boundary problems with spectral parameters entering via polynomials in the boundary conditions, Functional Anal. Appl., 9 (1975), 358-359. [22] E. M. Russakovskii, The matrix Sturm-Liouville problem with spectral parameter in the boundary conditions. Algebraic and operator aspects, Trans. Moscow Math. Soc., 57 (1996), 159-184. [23] A. A. Shkalikov, Boundary problems for ordinary differential equations with parameter in the boundary conditions, J. Sov. Math., 33 (1986), 1311-1342. ˇ [24] A. V. Straus, On spectral functions of differential operators, Izvest. Akad. Nauk SSSR Ser. Mat., 19 (1955), 201-220. [25] C. Tretter, On λ-nonlinear boundary eigenvalue problems, Mathematics Research 71 (1993), Akademie Verlag. [26] J. Walter, Regular eigenvalue problems with eigenparameter in the boundary conditions, Math. Z., 133 (1973), 301-312. [27] E. M. E. Zayed, S. F. M. Ibrahim, An expansion theorem for an eigenvalue problem on an arbitrary multiply connected domain with an eigenparameter in a general type of boundary conditions, Acta. Math. Sinica (N.S.), 11 (1995), 399-407.

17

Sturm-Liouville problems with boundary conditions ...

Sep 29, 2008 - More precisely, we define y(x, λ) to be a nonzero solution of (1.1)-(1.2), analytic in λ ∈ C, and we write ω(λ) = y′(1,λ) − f(λ)y(1,λ). By definition ...

186KB Sizes 1 Downloads 152 Views

Recommend Documents

Sturm-Liouville Problems with Reducible Boundary ...
Jan 31, 2005 - University of Calgary. Calgary, Alberta, Canada T2N 1N4. Patrick J. Browne †. Mathematical Sciences Group. Department of Computer Science.

Boundary conditions at the limiter surface obtained in ...
M2P2 UMR 6181, CNRS - Aix-Marseille Université, * CEA IRFM, 13108 Saint Paul-Lez-Durance, France. Overview Summary. Conclusions. References.

Weakly Enforced Essential Boundary Conditions for ...
The boundary of the extended cell domain ∂Ωα is assumed traction free. Traction ..... WEAK BOUNDARY CONDITIONS FOR THE FCM. 13. AΩ/AΩext. C. S. 100.

Variational boundary conditions based on the Nitsche ...
stitute an e ective alternative to boundary- ed isogeometric parametrizations for constructing. C .... For a detailed discussion of the theory that forms the basis for ..... F 2: e numerical solution of the concentration c and the free energy ψ at .

Boundary conditions for plasma fluid models at the ...
The problem of fluid codes. ▻ Quasi-neutrality ne ≃ ni ... Need B.C. that supply the sheath physics to fluid codes. ▻ To describe the main ... + cs (±1 + θn ± θTe /2) ∂2 s v||i i. Effect of radial gradients ∂x n, ∂x φ, ∂x Te : θA

Indefinite boundary value problems on graphs
Apr 12, 2011 - 31-61. [10] S. Currie; Spectral theory of differential operators on graphs, PhD Thesis, University of the Witwatersrand, Johannesburg, 2006. [11] S. Currie, B.A. Watson; Dirichlet-Neumann bracketing for boundary-value problems on graph

ON INITIAL-BOUNDARY VALUE PROBLEMS FOR A ...
Abstract. We consider a Boussinesq system of BBM-BBM type in two space dimensions. This system approximates the three-dimensional Euler equations.

initial-boundary-value problems for the bona-smith ...
Sep 21, 2008 - the solutions of the resulting system with suitable initial data will ..... mapping theorem applies to Γ consider as a mapping of BR into itself.

System, Subsystem, Hive: Boundary Problems in ...
Jul 27, 2016 - Keywords: consciousness, open dynamical system, brain dynamics, ..... 4Note that both “inputs” in the classical sense, that is, energy transduced by ..... of bees with an alternative system of pheromones in our example.

Art Therapy with Neurological Conditions - icaat
REF: Weston 2015. Follow us on Twitter: http://twitter.com/#!/JKPBooks. Find us on Facebook: http://www.jkp.com/jkp/socialnetworks.php.

Art Therapy with Neurological Conditions - icaat
REF: Weston 2015. Follow us on Twitter: http://twitter.com/#!/JKPBooks. Find us on Facebook: http://www.jkp.com/jkp/socialnetworks.php.

Phase-field boundary conditions for the voxel finite cell ...
Minnesota, 500 Pillsbury Drive S.E., Minneapolis, MN 55455, USA; Phone: +1 ..... based on CT scans in Section 5, the distinction between a hard tissue ..... solvability of the system, we assign a very small stiffness (in our case c ... quantitative C

A Study of Some Lacunary Boundary Value Problems by B-Spline.pdf ...
In partial fulfillment of the requirements for the degree of. Master of Science in Mathematics. (Numerical Analysis). By. Bzhar Jamal Aziz. B.Sc. in Mathematics ...

External guidance with the general conditions for recruitment and ...
May 15, 2018 - AST: from the time when a post-secondary education diploma was ... the time when, on completion of a minimum of three years of study, the ...

Boundary estimates for solutions of non-homogeneous boundary ...
values of solutions to the non-homogeneous boundary value problem in terms of the norm of the non-homogeneity. In addition the eigenparameter dependence ...

Boundary Final.pdf
BOYD LAKE. WELD COUNTY. ROAD 62 ... POWELL. SANCTUARY. SEA GULL. GOLDCO GRUMMAN. DRAFT. HORSE ... Page 1 of 1. Boundary Final.pdf.

Saliency Detection based on Extended Boundary Prior with Foci of ...
Page 1 of 5. SALIENCY DETECTION BASED ON EXTENDED BOUNDARY. PRIOR WITH FOCI OF ATTENTION. Yijun Li1. , Keren Fu1. , Lei Zhou1. , Yu Qiao1. , Jie Yang1∗. , and Bai Li2. 1. Institute of Image Processing and Pattern Recognition, Shanghai Jiao Tong Uni

Comparison of Planetary Boundary Layer Model Winds with ...
with the unique characteristic of the PBL model being able to account for the nonlinear effects of organized large eddies. .... event off France in 1999. This ability to predict .... sure analyses are corrected with the surface pressure observations 

A Direct-Forcing Immersed Boundary Method with Dynamic Velocity ...
A Direct-Forcing Immersed Boundary. Method with Dynamic Velocity Interpolation ... GIS terrain data (10 m resolution) tunnel fires http://www.protectowire.com ...