Sturm-Liouville Problems with Reducible Boundary Conditions



Paul A. Binding † Department of Mathematics and Statistics University of Calgary Calgary, Alberta, Canada T2N 1N4 Patrick J. Browne † Mathematical Sciences Group Department of Computer Science University of Saskatchewan Saskatoon, Saskatchewan, Canada S7N 5E6 Bruce A. Watson ‡ School of Mathematics University of the Witwatersrand Private Bag 3, P O WITS 2050, South Africa January 31, 2005



Keywords: Sturm-Liouville, eigenparameter dependent boundary conditions, Mathematics subject classification (2000): 34B25, 46D05, 47E05. † 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

Abstract The regular Sturm-Liouville problem τ y := −y 00 + qy = λy

on [0, 1], λ ∈ C,

is studied subject to boundary conditions Pj (λ)y 0 (j) = Qj (λ)y(j),

j = 0, 1,

where q ∈ L1 (0, 1) and Pj , Qj are polynomials with real coefficients. Comparison is made between this problem and the corresponding “reduced” one where all common factors are removed from the boundary conditions. Topics treated include Jordan chain structure, eigenvalue asymptotics and eigenfunction oscillation.

1

Introduction

This is the first of two papers in which we shall study certain factorization questions for the regular Sturm-Liouville problem τ y := −y 00 + qy = λy

on [0, 1], λ ∈ C,

(1.1)

subject to boundary conditions Pj (λ)y 0 (j) = Qj (λ)y(j),

j = 0, 1,

(1.2)

where q ∈ L1 (0, 1) and Pj , Qj are polynomials with real coefficients. Here we shall discuss the problem from the differential equation viewpoint, and factorization enters, at least initially, via an appropriately defined characteristic function - see (2.3). It turns out that some of the ensuing analysis will also be important for our sequel where we consider an operator theoretic setting, and, in particular, operator factorization. Sturm-Liouville problems with λ-dependent boundary conditions have been studied by many authors. The reference lists of [5, 8, 9, 15] will give some idea of the scope, both applied and theoretical, of this activity. While most investigations have concerned polynomials Pj , Qj of degree one, various authors, e.g., in [2, 6, 7, 14] have studied polynomials of higher degree. In general, these conditions have been assumed in reduced form, i.e., so that for each fixed j, P j and Qj have no common factors. In [3, 11], however, mention was made of the case in which P 1 and Q1 have a common root. Moreover, as will be seen in our sequel, even when (1.2) is in reduced form, the operator factorization mentioned above naturally leads to associated nonreduced problems. Indeed, this aspect is our principal motivation for the study of the present topic. An eigenvalue for our problem is a value of the parameter λ for which (1.1)-(1.2) can be solved non-trivially for y. Even though the boundary conditions are separated, it

2

turns out that the eigenvalues may be (geometrically) nonsimple and/or (algebraically) nonsemisimple. For example, if P0 and Q0 have a common root ξ, then λ = ξ will be an eigenvalue since (1.2) is satisfied with j = 0 and λ = ξ, so it is enough for y to satisfy (1.1), and the terminal condition (1.2) with j = 1, at λ = ξ. If ξ is also a root of P 1 and Q1 , then ξ will be a eigenvalue of geometric multiplicity two. Suppose for some λ that it is possible to construct a Jordan chain of length m of associated eigenfunctions; i.e., that there exist functions y [0] , ..., y [m−1] satisfying (τ − λ)y [0] = 0,

(τ − λ)y [r] = y [r−1] ,

and the boundary conditions r i X 1 h (i) 0 (i) 0= Pj (λ)y [r−i] (j) − Qj (λ)y [r−i] (j) , i!

(1.3)

j = 0, 1,

(1.4)

i=0

for r = 1, ..., m−1. If m is the greatest integer with these properties, then λ has algebraic multiplicity m – see [13, Section 2.3] for details. If m = 1 then λ is called semisimple. As we shall see, eigenvalues can be semisimple or not independently of their geometric multiplicity. By definition, the ‘reduced’ problem arising from (1.1)-(1.2) is that in which (1.2) is replaced by ˆ j (λ)y(j), Pˆj (λ)y 0 (j) = Q

j = 0, 1,

(1.5)

ˆ j are obtained from Pj , Qj by the removal of all common factors of the where Pˆj and Q k ˆ j have no common roots. form (λ − ξ) . Thus Pˆj and Q In Section 2 we study the relationship between the Jordan chain structures of (1.1)(1.2) and the reduced problem. It turns out that there are interesting “interference” effects when there are common roots of (1.2) that are also eigenvalues of the reduced problem – see Theorem 2.3. We give asymptotic developments for the eigenvalues of (1.1)-(1.2) in Section 3. The results resemble those for the standard case of λ-independent boundary conditions but with an index shift. The associated oscillation theory for the eigenfunctions is presented in Section 4. Again there is an index shift, but it may be different from the previous one. As far as we know, these results are new, at least in this generality. We conclude with an example in Section 5 illustrating some of the above ideas.

2

Comparison of Jordan Structures

Let ϕ and ψ be the solutions of (1.1) satisfying the initial conditions ϕ(0, λ) = 1,

ψ(0, λ) = 0,

0

0

ϕ (0, λ) = 0,

ψ (0, λ) = 1,

3

(2.1) (2.2)

and define the characteristic function D by D(λ) = P1 (λ)[Q0 (λ)ψ 0 (1, λ) + P0 (λ)ϕ0 (1, λ)] − Q1 (λ)[Q0 (λ)ψ(1, λ) + P0 (λ)ϕ(1, λ)].(2.3) Then D(λ) is an entire function of order 1/2 since ϕ, ϕ 0 , ψ and ψ 0 are of order 1/2, [10, Appendix]. The eigenvalues of (1.1)-(1.2) are the solutions of D(λ) = 0 with the order ν(λ) of the zero of D at λ coinciding with the algebraic multiplicity of λ as an eigenvalue of (1.1)-(1.2), see [13, Section 2.3]. For an entire funtion f , let µ(f ; λ) denote the order of λ as a root of f = 0, with µ(f ; λ) = 0 if f (λ) 6= 0. Thus, as above, ν(λ) = µ(D; λ),

(2.4)

ζj (λ) = min{µ(Pj ; λ), µ(Qj ; λ)}.

(2.5)

and we also define

In general, we shall use ν(λ) to denote the algebraic multiplicity of λ for arbitrary λ ∈ C with ν(λ) = 0 indicating that λ is not an eigenvalue of (1.1)-(1.2). The spectrum of (1.1)-(1.2) then consists of all eigenvalues repeated according to algebraic multiplicity. In what follows, it will be convenient to regard chains of length 0 as nonexistent. For example, if the Jordan structure at λ consists of two chains of lengths 0 and m > 0, it is to be understood that λ has only one chain (of length m). Our first result compares the Jordan structure at λ = ξ of two problems, one of which is partially reduced at ξ in the sense that P j and Qj do not have ξ as a common root, while the other has Pj (λ) and Qj (λ) replaced by (λ − ξ)kj Pj (λ) and (λ − ξ)kj Qj (λ). Proposition 2.1 Let ξ ∈ C be such that |P j (ξ)| + |Qj (ξ)| 6= 0, j = 0, 1. If ν(ξ) = m then the boundary value problem with equation (1.1) and boundary conditions (λ − ξ)kj Pj (λ)y 0 (j) = (λ − ξ)kj Qj (λ)y(j),

j = 0, 1,

(2.6)

where kj ∈ N0 := N ∪ {0}, k0 + k1 ≥ 1, has ξ as an eigenvalue of algebraic multiplicity m + k0 + k1 . The Jordan structure at ξ for (1.1),(2.6) consists of two chains of lengths m+max{k0 , k1 } and min{k0 , k1 }. The chain of greater length commences with a solution of τ y = ξy which satisfies (1.1) for j = 0 (respectively j = 1) when k 0 ≤ k1 (respectively k0 ≥ k1 ). Proof: Without loss of generality we assume 0 ≤ k 0 ≤ k1 . Let [0]

y1 (x) = P0 (ξ)ϕ(x, ξ) + Q0 (ξ)ψ(x, ξ)

4

[0]

[0]

so that y1 satisfies (1.2) for λ = ξ, j = 0. Let y0 be another solution of τ y = ξy linearly [0] [0] [0] independent of y1 and put z [0] = Ay1 + By0 . Now choose a particular solution z [r] to τ z [r] = ξz [r] + z [r−1] , r = 1, 2, . . ., (for example using variation of parameters with [0] [0] y0 , y1 as a fundamental pair of solutions to τ y = ξy). We shall show that the z [r] form a Jordan chain for (1.1), (2.6) of length m + k 1 if B = 0, and of length k0 otherwise. To this end we examine the boundary conditions r i X 1 ∂i h kj [r−i] 0 [r−i] 0= , (λ − ξ) [P (λ)z (j) − Q (λ)z (j)] j j i! ∂λi λ=ξ

(2.7)

i=0

for j = 0, 1 and r = 0, 1, .... For r = 0, ..., k 0 − 1, (2.7) is automatically satisfied, while for r = k0 , j = 0, (2.7) becomes 0

P0 (ξ)z [0] (0) − Q0 (ξ)z [0] (0) = 0 which holds only if B = 0. Hence when B 6= 0 the chain z [0] , ..., z [k0 −1] cannot be extended without violating (2.7). [r]

Assume, then, that B = 0 and, without loss of generality, A = 1 so that z [r] = y1 . For r ≥ k0 , the right hand side of (2.7) is   i−p  r i  p X ∂ 1X i [r−i] 0 [r−i] kj ∂ (λ − ξ) [Pj (λ)y1 (j) − Qj (λ)y1 (j)] p i! ∂λi−p ∂λp λ=ξ p=0 i=0   r X 1 i (i−k ) (i−k ) [r−i] 0 [r−i] kj ![Pj j (ξ)y1 (j) − Qj j (ξ)y1 (j)] = i − k i! j i=kj

=

r X

i=kj

1 (i−k ) (i−k ) [r−i] [r−i] 0 (j) − Qj j (ξ)y1 (j)] [Pj j (ξ)y1 (i − kj )!

r−kj

=

X 1 (i) [(r−kj )−i] [(r−kj )−i] 0 (i) (j)]. (j) − Qj (ξ)y1 [Pj (ξ)y1 i!

(2.8)

i=0

[r]

By the construction of the functions y 1 , (2.8) vanishes for j = 0 and r = k0 , .... Recall [m−1] [0] is a Jordan chain at ξ for (1.1)-(1.2) and hence for j = 1, (2.8) that y1 , ..., y1 vanishes for r − k1 = 0, ..., m − 1. Thus, relative to (1.1),(1.5), ξ has a Jordan chain [0] commencing with y1 of length at least m + k1 . If it were possible to extend this chain, [m+k1 ] to satisfy (2.7) for j = 1 and r = m + k1 . However (2.8) then one could find y1 would then show that relative to (1.1)-(1.2), ν(ξ) > m - a contradiction. The next result shows that the two problems (1.1)-(1.2) and (1.1), (2.6) have Jordan structures which differ only at ξ.

5

Proposition 2.2 Suppose that for (1.1)-(1.2), ξ has Jordan chains Cs : ys[0] , ..., ys[ms −1] ,

ms ≥ 0, s = 0, 1.

Let fj be entire functions with fj (ξ) 6= 0, j = 0, 1. Then C0 , C1 are also Jordan chains for (1.1) with boundary conditions fj (λ)Pj (λ)y 0 (j) = fj (λ)Qj (λ)y(j), [0]

[0]

[r]

[r]

[r−1]

Proof: Since τ ys = ξy0 and τ ys = ξys + ys show that for s = 0, 1,

j = 0, 1.

for r = 1, ..., ms − 1, it remains to

r i X 1 ∂i h [r−i] 0 [r−i] f (λ)[P (λ)y (j) − Q (λ)y (j)] j j j s s i! ∂λi λ=ξ

(2.9)

i=0

[ms ]

vanishes for r = 0, ..., ms − 1, j = 0, 1, and that for any solution y s

of

τ ys[ms ] = ξys[ms ] + ys[ms −1] , (2.9) does not vanish for at least one value of j = 0, 1. Now (2.9) can be expressed as r i X 1 ∂i h [r−i] 0 [r−i] f (λ)[P (λ)y (j) − Q (λ)y (j)] j j j s s i! ∂λi λ=ξ i=0   i r X 1X i 0 (k) (i−k) (i−k) fj (ξ)[Pj (ξ)ys[r−i] (j) − Qj (ξ)ys[r−i] (j)] = k i! i=0 k=i " (i−k) # (k) (i−k) r r X fj (ξ) X Pj (ξ) [r−i] 0 Qj (ξ) [r−i] = y (j) − y (j) k! (i − k)! s (i − k)! s k=0 i=k " r−k !# (n) (n) (k) r X Pj (ξ) [(r−k)−n] 0 Qj (ξ) [(r−k)−n] fj (ξ) X = ys (j) − ys (j) . (2.10) k! n! n! n=0

k=0

For r = 0, ..., ms − 1, the internal summation in (2.10) vanishes since C s is a Jordan [m ] chain for (1.1)-(1.2) of length ms . If ys s could be found with the properties described, we see that for r = ms , (2.10) reduces to fj (ξ)

ms X

n=0

(n)

Pj (ξ) n!

0 ys[ms −n] (j)

(n)



Qj (ξ) n!

ys[ms −n] (j)

!

which cannot vanish for both j = 0, 1, since f j (ξ) 6= 0 by hypothesis and Cs has length ms . The combination of the above two propositions leads to the main result of this section, in which we compare the problem (1.1)-(1.2) with its reduced version (1.1), (1.5).

6

Theorem 2.3 For the reduced problem (1.1), (1.5), let νˆ(λ) denote the algebraic multiplicity of λ. Then, in the notation of (2.4) and (2.5), ν(λ) = νˆ(λ) + ζ0 (λ) + ζ1 (λ). All eigenvalues of (1.1), (1.5) are geometrically simple and relative to (1.1)-(1.2), the Jordan structure at λ consists of two chains with lengths min{ζ 0 (λ), ζ1 (λ)} and νˆ(λ) + max{ζ0 (λ), ζ1 (λ)}. The chain of greater length must have as its zeroth element a solution of (1.1) which satisfies (1.5) for j = 0 (resp. j = 1) if ζ 0 (λ) ≤ ζ1 (λ), (resp. ζ0 (λ) ≥ ζ1 (λ)). ˆ Proof: For (1.1),(1.5) let D(λ) denote the expression corresponding to D(λ) for (1.1)(1.2). Then clearly ˆ λ) + ζ0 (λ) + ζ1 (λ). µ(D; λ) = µ(D; ˆ λ), the first claim is immediate. The remainder Since ν(λ) = µ(D; λ) and νˆ(λ) = µ(D; of the theorem follows from Propositions 2.1, 2.2.

3

Spectral Asymptotics

In this section we shall produce asymptotic expressions for the eigenvalues, λ n , of (1.1)(1.2). This extends the results of [4] where reduced problems were considered, moreover with only one λ-independent boundary condition. The eigenvalues of (1.1)-(1.2) will be labelled λn , n ≥ 0, repeated according to algebraic multiplicity and listed by increasing real parts. We write Pj (λ) = aj0 λdj + ... + ajdj ,

(3.1)

+ bjdj ,

(3.2)

Qj (λ) =

bj0 λdj

+ ...

where dj = max{deg(Pj ), deg(Qj )}, so that |aj0 | + |bj0 | 6= 0, j = 0, 1. Theorem 3.1 For sufficiently large n, λ n is real and is given by   R1 b0 b1  a00 6= 0 6= a10 , n ˆ 2 π 2 − 2 a01 + 2 a00 + 0 q + O n1 ,   0 0   2 R1   a1 b0   n ˆ + 21 π 2 + 2 b11 + 2 a00 + 0 q + O n1 , a00 6= 0 = a10 , 0 0   R1 b10 a01 1 2 2 1   , a00 = 0 6= a10 , n ˆ + π − 2 − 2 + q + O 1 0  2 n 0 a0 b0     R1  a1 a0  (ˆ n + 1)2 π 2 + 2 11 − 2 01 + q + O 1 , a0 = 0 = a1 . b0

b0

0

where n ˆ = n − d 0 − d1 .

7

n

0

0

Proof: We rely on the following asymptotic expressions as |λ| → ∞ for the functions ϕ(1, λ), ϕ0 (1, λ), ψ(1, λ) and ψ 0 (1, λ) - see [10] for details. √ ! √ Z √ sin λ 1 e|= λ| ϕ(1, λ) = cos λ + √ , q+O λ 2 λ 0 √ ! √ Z 1 |= λ| √ √ λ e cos √ ϕ0 (1, λ) = − λ sin λ + , q+O 2 λ 0 √ ! √ Z √ e|= λ| sin λ cos λ 1 √ q+O , − ψ(1, λ) = 2λ λ3/2 λ 0 √ ! √ Z 1 |= λ| √ sin λ e ψ 0 (1, λ) = cos λ + √ . q+O λ 2 λ 0 The expressions for ψ(1, λ), ψ 0 (1, λ) given in [10] involve only the leading terms above, but the more accurate asymptotics can readily be obtained using the methods of [10] by which the quoted expressions for ϕ(1, λ) and ϕ 0 (1, λ) were generated - see also [4, Appendix]. We may write Pj (λ) = λ

! 1 , λ2  ! bj1 1 j b0 + +O . λ λ2

aj0

dj

Qj (λ) = λdj

aj + 1 +O λ



Consider the cases for which a00 6= 0. Then (2.3) and the above expressions lead to  1  |=√λ| i h√  √ √ R  b0 b00 1 1 e√ 1  − − q cos λ sin λ + λ + O , a10 6= 0, a 1 0 0 2 0 −D(λ) a0 a0 λ h   √ i  √ = (3.3)  b1 cos √λ + b00 + a11 + 1 R 1 q sin√ λ + O e|= λ| , a00 λd0 +d1 a1 = 0. 0

a00

When a10 6= 0, we put

2

b10

0

b1 b0 1 c = 01 − 00 − a0 a0 2

Z

0

λ

λ

1

q 0

and write (3.3) as " √ √ √ D(λ) λ sin λ + c cos λ + O = −a10 a00 d +d 0 1 λ



e|= λ| √ λ

!#

.

For large n ∈ N, D



1 n+ 2

2

π

2

!

=

−a10 a00



   π 2(d0 +d1 )+1 1 n nπ + (−1) + O 2 n

8

(3.4)

which oscillates in sign with n. Thus there is µ n = nπ + δn , |δn | < π/2 with D(µ2n ) = 0. From (3.4) we obtain   1 n n (nπ + δn )(−1) sin δn + c(−1) cos δn = O (3.5) n  showing that sin δn = O n1 and hence δn = γn /n where γn = O(1). Now (3.5) yields   1 c γn = − + O , π n and hence µn = nπ −

c +O nπ



1 n2



which gives the existence of eigenvalues of the desired asymptotic form. It remains to show that all eigenvalues of large modulus are of the form µ 2n and to determine the index of µ2n in the listing λ0 , λ1 , .... To this end we use Rouch´e’s theorem. Let R be the entire function √ √ R(λ) = −a10 a00 λd0 +d1 λ sin λ and Γn the path in the complex plane Γn = {λ = ξ 2 |ξ ∈ τn } where τn is the path in the ξ-plane connecting −iζ n to iζn by three line segments, as shown in Figure 1. Here   1 ζn = n − π, n = 1, 2, 3, ... . 2 6

iζn 

6

ζn -

h

-

−iζn

Figure 1: τn in the ξ-plane 9

There is a positive constant κ (independent of n) so that for λ ∈ Γ n √ √ |R(λ)| ≥ κ| λ|2d0 +2d1 +1 e|= λ| ,

and as

 √  D(λ) − R(λ) = O λd0 +d1 e|= λ|

we have that on Γn , for n large,

|D(λ) − R(λ)| < |R(λ)|. Rouch´e’s theorem shows that D and R have the same number of zeros enclosed by Γ n , namely n + d0 + d1 . Moreover, the region between Γn and Γn+1 contains precisely one zero of D, namely µ2n . The upshot is that the µ2n form all the zeros with large modulus of D, and that λd0 +d1 +n = µ2n . This completes the proof for the case a 00 a10 6= 0. For the case a00 6= 0 = a10 , (so that b10 6= 0), we observe from (3.3) that D(n2 π 2 ) oscillates in sign with n, so that D(µ2n ) = 0 where   1 π + δn , µn = n + 2 with |δn | < π/2. Setting c=

b00 a11 1 + + 2 a00 b10

we come to

γn  , n + 21 π

δn =

1

q,

0

c(−1)n cos δn  =O n + 12 π + δn

(−1)n+1 sin δn + so that

Z

where



1 n2



,

γn = O(1).

Then it follows that γn = c + O(1/n) and µ2n

=



1 n+ 2

2

a1 b0 π + 2 11 + 2 00 + b0 a0 2

Z

0

1

  1 . q+O n

The Rouch´e argument now uses the entire function R(λ) = −b10 a00 λd0 +d1 cos

√ λ

and the path Γn with ζn = nπ. We leave the remaining details to the reader.

10

When a00 = 0 6= a10 (so that b00 6= 0) we consider the problem −Y 00 (x) + q(1 − x)Y (x) = λY (x),

(3.6)

0

(3.7)

0

(3.8)

−P1 (λ)Y (0) = Q1 (λ)Y (0),

−P0 (λ)Y (1) = Q0 (λ)Y (1).

Of course, this problem is equivalent to (1.1)-(1.2) via Y (x) = y(1 − x). However, the analysis of the previous case applies and we readily obtain the claim of the theorem. For the case a00 = 0 = a10 we have √ √ √ D(λ) = λ sin λ + cos λ 0 1 b0 b0 λd0 +d1 −1



a01 a11 1 − 1 + 2 b00 b0

Z

1





q +O 0

e|= λ| √ λ

!

.

The argument now follows lines similar to that for the case a 00 a10 6= 0 and we leave the details to the reader.

4

Oscillation Theory

To study the oscillation theory associated with (1.1)-(1.2), we assume in this section that n is large enough for λn to be a simple real eigenvalue. That this is possible follows from Theorem 3.1. We aim to determine the number of zeros in (0, 1) of the eigenfunction for λn . We shall rely heavily on the Pr¨ ufer angle θ, for which [1] gives the essential theory. This angle satisfies the first order equation θ 0 = cos2 θ + (λ − q) sin2 θ

(4.1)

and is related to (1.1) via cot θ = y 0 /y. We impose an initial condition θ(0) = α, 0 ≤ α < π on (4.1) and we write θ(x, λ, α) to display the dependence of θ on λ and α. Evidently, θ is increasing in both λ and α. The oscillation number associated with λ n is the unique integer kn ≥ 0 for which θ(1, λn , αn ) ∈ (kn π, (kn + 1)π]. Here, and subsequently, we shall use   −1 Q0 (λn ) , αn = cot P0

βn = cot

−1



Q1 (λn ) P1



where we take the principal value of cot −1 so that αn , βn ∈ (0, π). When (1.1) is subject to boundary conditions y0 (0) = cot α, y

y0 (1) = cot β, y

11

0 ≤ α < π, 0 < β ≤ π,

the resulting eigenvalues λα,β n , n ≥ 0, have asymptotic developments   Z 1 1 α,β 2 2 λn = (n + ν) π − 2 cot β + 2 cot α + q+O , n 0 where ν = 0 if α 6= 0, β 6= π; ν = 1/2 if α = 0, β 6= π or α 6= 0, β = π; and ν = 1 if α = 0, β = π. Moreover, θ(1, λα,β n , α) = nπ + β. With this background, we present our first result. For convenience we shall continue to use n ˆ = n − d 0 − d1 . Q

The value of kn depends on whether the limits as λ → ∞ of Pjj (λ), j = 0, 1, are finite, +∞ or −∞. This apparently gives nine cases, although there are but three outcomes. Theorem 4.1 For large n, the oscillation number associated with λ n is kn = n ˆ + η 0 + η1 Q0 Q1 where η0 = 1 if lim (λ) = −∞, η0 = 0 otherwise, and η1 = 1 if lim (λ) = ∞, λ→∞ P0 λ→∞ P1 η1 = 0 otherwise. Proof: We shall rely heavily on the relations Q1 (λn ), P1 θ(1, λn , αn ) = kn π + βn .

cot θ(1, λn , αn ) =

(i) Suppose

Qj (λ) Pj

→ +∞, j = 0, 1 as λ → ∞, so that αn ↓ 0, βn ↓ 0. By Theorem 3.1

λn = (ˆ n + 1)2 π 2 + c + q¯ + O(1/n), R1 where c is a constant and we write q¯ = 0 q. Then for large n 0,π/2

λnˆ

(4.2)

π/2,π

< λn < λnˆ +1 ,

and so 

1 n ˆ+ 2



    0,π/2 π/2,π π π = θ 1, λnˆ , 0 < θ (1, λn , αn ) < θ 1, λnˆ +1 , = (ˆ n + 2) π. 2

Since βn > 0, we have kn = n ˆ + 1. 0

1

(λ) → constant and PQ1 (λ) → ∞ as λ → ∞, so that αn → α 6= 0, π, βn ↓ (ii) Suppose Q P0 0. Then by Theorem 3.1   1 2 2 π + c + q¯ + O(1/n) (4.3) λn = n ˆ+ 2 > n ˆ 2 π 2 + 2 cot(α − ) + q¯ + O(1/n), α−,π/2

= λnˆ

+ O(1/n).

12

for small  > 0

Hence for large n α−,π/2

λnˆ

< λn < λn0,π ˆ ,

and so       1 α−,π/2 n ˆ+ π = θ 1, λnˆ , α −  < θ (1, λn , αn ) < θ 1, λn0,π , 0 + π = (ˆ n + 2) π. ˆ 2 Since θ (1, λn , αn ) = kn π + βn , where βn ↓ 0, we have kn = n ˆ + 1. 1

0

(iii) Suppose PQ0 (λ) → constant and Q P 1 (λ) → −∞ as λ → ∞, so that αn → α 6= 0, π, βn ↑ π. Then (4.3) holds and for small  > 0 and large n, λn < (ˆ n + 1)2 π 2 + 2 cot(α + ) + (c − 2 cot(α + ) + 2) − 1 + O(1/n) = λnα+,β ˆ +1 − 1 + O(1/n).

where β is defined in (0, π) by −2 cot β = c − 2 cot(α + ) + 2. Hence α+,β λn0,π ˆ −1 < λn < λn ˆ +1 ,

and so

Since

    α+,β n ˆ π = θ 1, λn0,π , 0 < θ (1, λ , α ) < θ 1, λ , α +  = (ˆ n + 1) π + β. n n ˆ −1 n ˆ +1 θ (1, λn , αn ) = kn π + βn ,

where βn ↑ π, we see that kn = n ˆ. (iv) Suppose Theorem 3.1

Q0 (λ) P0

→ ∞ and

Q1 (λ) P1

→ −∞ as λ → ∞, so that αn ↓ 0, βn ↑ π. By

λn < (ˆ n + 1)2 π 2 + c + 2 + q¯ − 1 + O(1/n) = λnα,β ˆ +1 − 1 + O(1/n) where α and β are defined in (0, π) by 2 cot α = 21 c + 1 and −2 cot β = 21 c + 1. Thus for large n 0,π/2 λnˆ < λn < λnα,β ˆ +1 , and so 

1 n ˆ+ 2



    0,π/2 π = θ 1, λnˆ , 0 < θ (1, λn , αn ) < θ 1, λnα,β , α = (ˆ n + 1) π + β. ˆ +1

Since θ (1, λn , αn ) = kn π + βn , where βn ↑ π, we have kn = n ˆ.

13

(v) Suppose Theorem 3.1

Q0 (λ) P0

→ −∞ and

Q1 (λ) P1

→ ∞ and as λ → ∞, so that αn ↑ π, βn ↓ 0. By

λn > (ˆ n + 1)2 π 2 + c − 2 + q¯ + 1 + O(1/n) = λnα,β ˆ +1 + 1 + O(1/n) where α and β are defined in (0, π) by 2 cot α = 21 c − 1 and −2 cot β = 21 c − 1. Thus for large n 0,π/2 λnˆ +1 > λn > λnα,β ˆ +1 , and so       5 0,π/2 n ˆ+ π = θ 1, λnˆ +1 , 0 + π > θ (1, λn , αn ) > θ 1, λnα,β , α = (ˆ n + 1) π + β. ˆ +1 2

Since

θ (1, λn , αn ) = kn π + βn , where βn ↓ 0, we have kn = n ˆ + 2. (vi) Suppose 0, π. Then

Qj (λ) Pj

→ constants, j = 0, 1, as λ → ∞, so that α n → α 6= 0, π, βn → β 6= λn = n ˆ 2 π 2 + c + q¯ + O(1/n),

and for small  > 0 and large n,     1 2 1 2 + 2 cot(α − ) + q¯ + O(1/n) < λn < n ˆ+ + 2 cot(α + ) + q¯ + O(1/n) < λn n ˆ− 2 2 and thus and so

α+,π λnα−,π , ˆ −1 < λn < λn ˆ

    α+,π n ˆ π = θ 1, λnα−,π , α −  < θ (1, λ , α ) < θ 1, λ , α +  = (ˆ n + 1) π. n n ˆ −1 n ˆ

Hence kn = n ˆ.

The remaining cases can be obtained from cases (i)-(iii) by using a reflection argument and considering the problem (3.6)- (3.8) as in the proof of Theorem 3.1.

5

Example

We consider the problem −y 00 = λy 0

(5.1) 2

λy (0) = −λ y(0) 2 0

3

λ y (1) = λ y(1).

14

(5.2) (5.3)

5.1 The corresponding reduced problem has (5.2)-(5.3) replaced by y 0 (0) = −λy(0),

y 0 (1) = λy(1).

(5.4)

It is a straightforward exercise to show that (5.1), (5.4) has a sequence of eigenvalues µ 2n with corresponding eigenfunctions y n , where µ0 = 0, 2µn tan µn = 2 , µn − 1

y0 (x) = 1; yn (x) = sin(µn x) − µ−1 n cos(µn x),

n ≥ 1.

(5.5)

We note (taking µn ≥ 0 for convenience) that µn = (n − 1)π + n where 0 < n = O(1/n). By Theorem 3.1 applied to the reduced problem, we see that these are the only eigenvalues of (5.1), (5.4), and moreover that each eigenvalue is simple. We remark that this does require proof, since (5.4) does not satisfy the “right definiteness” condition of [3] (in fact a self-adjoint operator formulation, as in say [14], would require a Pontryagin space of index 2). 5.2 For the non-reduced problem, Proposition 2.2 shows that the only difference in Jordan structure will occur at λ = 0, and then (5.1) has general solution y = Cx + D. Thus we can start a Jordan chain for (5.1)-(5.3) with y [0] (x) = Cx + D. The equation for y [1] is −y 00 = Cx + D, which has solution 1 1 y(x) = − Cx3 − Dx2 + Ex + F. 3! 2! 0

The boundary conditions to be satisfied are y [0] (0) = 0, which requires C = 0, and a null condition at x = 1. Thus unless C = 0, the chain terminates with y [0] and so has length one. When C = 0, we can without loss of generality take D = 1, and then y [1] (x) = − The equation for y [2] is

1 2 x + Ex + F. 2!

1 −y 00 = − x2 + Ex + F 2

which has solution y(x) =

1 4 E 3 F 2 x − x − x + Gx + H. 4! 3! 2! 0

0

The boundary conditions are 0 = y [1] (0) + y [0] (0) which forces E = −1, and y [0] (1) = 0 which is automatically satisfied. Thus we now have y [2] (x) =

1 1 1 4 x + x3 − F x2 + Gx + H. 4! 3! 2!

15

0

The equation for y [3] is −y 00 = y [2] with boundary conditions 0 = y [2] (0) + y [1] (0) which 0 forces G = −F , and y [1] (1) = y [0] (1) which requires −2 = 1, an impossibility. Hence this chain terminates with y [2] and so has length three. For (5.1)-(5.3), then, the eigenvalue λ = 0 generates two chains of lengths three and one, and thus has algebraic multiplicity four. This is in accord with the results of Section 2, since λ = 0 is simple for (5.1), and (5.4) ζ j (0) = 1, ζj (1) = 2 in the notation of (2.5). 5.3 Now we can turn to the results of Sections 3 and 4. Counting λ 0 four times (as above) we have λn = µ2n+4 for n ≥ 1 in the notation of Theorem 3.1 (P, Q being unreduced) so there is an index shift of 4 in the asymptotics. To illustrate Theorem 4.1, we consider the oscillation of yn for large n. Note that sin(µn x) has n − 1 zeros in (0, 1) and the −µ−1 n cos(µn x) term adds an extra zero of y n near x = 0. Finally   µ2n + 1 1 = yn (1) = cos(µn ) tan µn − µn µn (µ2n − 1) from (5.5), so no extra zero is added near x = 1. Thus y n has n zeros in (0, 1), and the index shift for oscillation is 5. This exceeds the shift for asymptotics, a phenomenon which cannot happen for right definite, [3], or λ-independent boundary condition, [12], problems.

References [1] F. V. Atkinson, Discrete and continuous boundary value problems, Academic Press, New York, 1964. [2] A. I. Benedek, R. Panzone, On inverse problems for second-order differential operators with boundary dependence on the eigenvalue parameter, Rev. Un. Mat. Argentina, 30 (1982/83), 167–173. [3] P. A. Binding, P. J. Browne, K. Seddighi, Sturm-Liouville problems with eigenparameter dependent boundary conditions, Proc. Edinburgh Math. Soc., 37 (1993), 57-72. [4] P. A. Binding, P. J. Browne, B. A. Watson, Equivalence of inverse SturmLiouville problems with boundary conditions rationally dependent on the eigenparameter, J. Math. Anal. Appl., 291 (2004), 246-261. [5] L. Collatz, Eigenwertaufgaben mit technischen Anwendungen, Akademische Verlag, Leipzig, 1963. [6] A. Dijksma, Eigenfunction expansions for a class of J-selfadjoint ordinary differential operators with boundary conditions containing the eigenvalue parameter, Proc. Roy. Soc. Edinburgh, Series A, 87 (1980), 1-27. ˇ [7] 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.

16

[8] C. T. Fulton, Two-point boundary value problems with eigenparameter contained in the the boundary conditions, Proc. Roy. Soc. Edinburgh, 77A (1977), 293-308. [9] C. T. Fulton, Singular eigenvalue problems with eigenvalue-parameter contained in the the boundary conditions, Proc. Roy. Soc. Edinburgh, 87A (1980), 1-34. [10] H. Hochstadt, On inverse problems associated with Sturm-Liouville operators, J. Diff. Equations, 17 (1975), 220-235. [11] H. Langer, A. Schneider, On spectral properties of regular quasidefinite pencils F − λG, Results Math., 19 (1991), 89–109.

[12] E. L. Ince, Ordinary differential equations, Dover, New York, 1956.

[13] M. A. Naimark, Linear Differential Operators I, Ungar, New York, 1967. [14] 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. [15] J. Walter, Regular Eigenvalue Problems with Eigenparameter in the Boundary conditions, Math. Z., 133 (1973), 301-312.

17

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.

160KB Sizes 0 Downloads 154 Views

Recommend Documents

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 ...

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.

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 ...

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 ...

Saliency Detection based on Extended Boundary Prior with Foci of ...
K) and its mean position and mean color in. LAB color space are denoted as pi and ci respectively (both. normalized to the range [0,1]). 2.1. Graph Construction.

Cryptarithmetic Multiplication Problems with Solutions - Download PDF
Cryptarithmetic Multiplication Problems with Solutions - Download PDF - Papersadda.pdf. Cryptarithmetic Multiplication Problems with Solutions - Download ...

Estimating Bayesian Decision Problems with ...
Nov 11, 2014 - data is to estimate the decision-making parameters and understand, .... Therefore one can recover (θit,πit) if ρt, γit,0, and γit,1 are known.

ou problems with solutions.pdf
In Newton's rings experiment, the diameter of the 10th ring changes from 1.40 to 1.27cm, when. a liquid is introduced between the lens and the glass plate.

Problems with Online Focus groups
market researchers. keywords: online qualitative research; focus groups; self-disclosure; communication .... simulated a normal situation in a collective PC-room.

boundary layer
RH and OH not only depends on the strength and spatial distribution of the. RH emissions ..... Any present-day numerical model is only capable of sim- ulating 2 ...

Boundary Final.pdf
Effective. 29 May, 2016. Page 1 of 1. Boundary Final.pdf. Boundary Final.pdf. Open. Extract. Open with. Sign In. Main menu. Displaying Boundary Final.pdf.

Growing Garlic - Boundary Garlic
on time and curing properly are all important for producing bulbs with good ... Grain straw is not recommended because it can host wheat curl mite which will ...

School District Boundary Task Force
A request was made for the department to gather information about how other ... Disclaimer: Staffing of the task force by the Department of Education and ...