Oscillation Theory for a Quadratic Eigenvalue Problem Patrick J. Browne∗ Department of Mathematics and Statistics University of Saskatchewan, Saskatoon, Saskatchewan, Canada S7N 5E6 [email protected] Bruce A. Watson† School of Mathematics University of the Witwatersrand Private Bag 3, P O WITS 2050, South Africa [email protected] September 15, 2008

Abstract We study the Sturm-Liouville problem with quadratic dependence on the spectral parameter, −(p(x)y ′ (x))′ + q(x)y(x) − λs(x)y(x) = λ2 y(x),

0 ≤ x ≤ 1,

subject to boundary conditions y ′ (0) sin α = y(0) cos α, α ∈ [0, π), and y ′ (1) sin β = y(1) cos β, β ∈ (0, π]. Here, p, q, s are real valued functions satisfying p > 0, q, 1/p ∈ L1 (0, 1) and s ∈ L∞ (0, 1). The eigenvalues are not necessarily real, but, as we show, all but a finite number of the eigenvalues are real and algebraically simple. We are mainly concerned with determining the number of zeros in (0, 1) for the eigenfunctions associated with real eigenvalues using two parameter eigencurve theory. En route we develop asymptotic estimates for the eigenvalues. Keywords: Quadratic pencil, Sturm-Liouville, oscillation count. Mathematics subject classification (2000): 34B25, 47E05. ∗

Research supported by the NSERC of Canada and conducted while visiting The University of The Witwatersrand. † Research supported in part by the South African National Research Foundation grant FA2007041200006 and by the John Knopfacher Centre for Applicable Analysis and Number Theory.

1

1

Introduction

We shall study the Sturm-Liouville problem with quadratic dependence on the spectral parameter −(p(x)y ′ (x))′ + q(x)y(x) − λs(x)y(x) = λ2 y(x), 0 ≤ x ≤ 1, (1.1) subject to boundary conditions

y′ (0) = cot α, y y′ (1) = cot β, y

α ∈ [0, π),

(1.2)

β ∈ (0, π],

(1.3)

which are to be interpreted as the Dirichlet conditions y(0) = 0 (respectively, y(1) = 0) if α = 0 (respectively, β = π). Here, p, q, s are real valued functions satisfying p > 0, q, 1/p ∈ L1 (0, 1) and s ∈ L∞ (0, 1) with −∞ < s ≤ s(x) ≤ s¯ < ∞ a.e. in [0, 1]. There is considerable literature on this problem but frequently with p = 1 and with stronger continuity or differentiability restrictions on the coefficients q and s. As an example we cite the work of Roach and Sleeman [19, 20] where an abstract version of the problem is studied and then applied to (1.1 - 1.3). Binding, in [2], also discusses an abstract formulation but does not directly apply his results to (1.1 - 1.3). We also direct the reader to [1, 16, 18, 21] where problems of this or similar type are discussed. Variational theories for (1.1 - 1.3) have been presented in [11, 22, 23] and studies of the inverse spectral problem associated with (1.1 - 1.3) have been undertaken by, for example, [12, 13, 15] - see also the references cited therein. In the forward problem, while asymptotics for eigenvalues have been stated in [12, 13, 15] under stronger assumptions on p, q, s, attention has mostly been directed to completeness and eigenfunction expansion questions. For example, Roach and Sleeman [19, 20] recast (1.1 - 1.3) as a linked two parameter system in L2 (0, 1)⊗C2 and set their completeness results in this space. Binding [2] establishes the equivalence of L2 (0, 1)⊗C2 with L2 (0, 1)⊕L2 (0, 1) and gives abstract eigenvector completeness results in this latter space. These works and other related references require positive definiteness of the Sturm-Liouville operator generated by p, q as in (1.1) and with boundary conditions (1.2 - 1.3). We shall develop this aspect further subsequently, but it suffices here to say that such a demand ensures that the eigenvalues λ for (1.1 - 1.3) are real and algebraically simple and, via Binding’s abstract theory, the associated eigenfunctions form a frame for L2 (0, 1). Without this positivity assumption, (1.1 - 1.3) may have non-real eigenvalues (indeed, eigenvalues for which both λ and λ2 are non-real). It is not difficult to produce simple examples with these properties. None the less, all but a finite number of the eigenvalues are real and algebraically simple, as we shall establish in later sections. We are mainly concerned with oscillation theory for (1.1 - 1.3) in the sense that, for real eigenvalues we shall be interested in determining the number of zeros in (0, 1) of 2

the associated eigenfunctions. It will be necessary here to assume that p is absolutely continuous. To the best of our knowledge this aspect of the quadratic problem has not been developed to any significant extent hitherto. Our approach will rely heavily on the theory of eigencurves for two parameter problems developed by Binding and Browne in [3, 4, 5, 6, 7, 8, 10]. In essence, we shall replace λ2 in (1.1) by, a second parameter, −µ and study carefully the intersections of the resulting eigencurves µk (λ), k ≥ 0, λ ∈ R, with the parabola µ = −λ2 . At a point λ where µk (λ) intersects this parabola, (1.1 - 1.3) has an eigenvalue whose eigenfunction has k zeros in (0, 1). Further the algebraic multiplicity of the eigenvalue λ for (1.1 - 1.3) coincides with the order of contact between the curves µk (λ) and µ = −λ2 . In this manner we shall be able to give a “Sturm oscillation theorem” for (1.1 - 1.3). In Section 2 we formulate the two parameter eigenvalue problem needed in later work and for real λ we relate the algebraic multiplicity of λ as an eigenvalue of (1.1 - 1.3) to the order of contact between µk (λ) and µ = −λ2 . Section 3 contains our main results on the reality and algebraic simplicity of the eigenvalues as well as our oscillation theory. En route we develop asymptotic estimates for the eigenvalues. While these estimates as not as precise as those in [12, 13, 15], our basic assumptions on p, q and s are weaker than those made by the above cited authors. Moreover, our estimates are sufficient to establish our principal oscillation results.

2

Eigencurves

We start by defining the Sturm-Liouville operator, T , generated by p, q, α, β, as T y = −(py ′ )′ + qy,   y, py ′ ∈ AC[0, 1], −(py ′)′ + qy ∈ L2 [0, 1], 2 . D(T ) = y ∈ L [0, 1] y satisfies (1.2 - 1.3)

Then T is self-adjoint in L2 [0, 1], has compact resolvent and has spectrum consisting of isolated real eigenvalues τ0 < τ1 < τ2 < . . . , with τn → ∞ as n → ∞. The eigenfunctions of T associated with τn have n zeros in (0, 1). Further τn has the asymptotic estimate  2 nπ τn =  R 1 1  + o(n2 ), as n → ∞. (2.1) 0



p

Sharper estimates are available if p has additional smoothness. In particular, if p is absolutely continuous then √

τn =

(n + ν) π R 1 1 + o(1) as n → ∞. 0



p

3

(2.2)

Here ν is half the number of Dirichlet conditions specified by (1.2 - 1.3). Further, if p′ is absolutely continuous then o(1) may be replaced by O(1/n), while if p is constant - say p = p0 - then R1 cot∗ α − cot∗ β + 12 0 q √ √ τn = p0 (n + ν)π + + o(1/n), (2.3) √ p0 (n + ν)π where the terms cot∗ α and cot∗ β respectively are to be replaced by 0 if α = 0 and β = π, respectively, see [9, Theorem A3]. We can now describe the two-parameter problem to be studied in connection with (1.1 - 1.3). Consider the equation (T − λS + µ) y = 0,

0 6= y ∈ D(T ),

(2.4)

where S : L2 [0, 1] → L2 [0, 1] is the bounded self-adjoint multiplication operator given by (Sy)(x) = s(x)y(x). For λ ∈ R, (2.4) generates a sequence of eigencurves µ0 (λ) > µ1 (λ) > . . . such that µk (0) = −τk and µk is an analytic function of λ. Here µ0 (λ) is convex in λ ∈ R, µk (λ) → −∞ as k → ∞ for each fixed λ ∈ R and the non-trivial solutions of (1.1 - 1.3) with µ = µk (λ) have exactly k zeros in (0, 1). These and other properties of the curves µk are described in detail in [3, 4, 5, 6, 7]. We recall, below, the algebraic multiplicity of an eigenvalue of (1.1 - 1.3), [17, pages 16 -20]. For brevity we write W (λ) = T − λS − λ2 I. Definition 2.1 (i) An eigenvalue of (1.1 - 1.3) is a complex number λ for which (1.1 1.3) has a non trivial solution. (ii) An eigenvalue λ0 of (1.1 - 1.3) is said to have algebraic multiplicity m ≥ 1 if functions ϕ0 , ϕ1 , . . . , ϕm−1 ∈ D(T ), with ϕ0 6= 0, can be found for which   i X 1 ∂j ϕi−j = 0, W (λ) j j! ∂λ λ=λ0 j=0

i = 0, . . . , m − 1,

(2.5)

but no function ϕm can be found to satisfy (2.5) with i = m. Alternatively, if we solve (1.1, 1.2) with a function y(λ, x) satisfying y ′(λ, 0) = cos α,

y(λ, 0) = sin α, 4

(2.6)

then y(λ, x) is an entire function of λ for each x, and the zeros of the characteristic determinant, ∆(λ) = y(λ, 1) cos β − y ′(λ, 1) sin β, (2.7)

are precisely the eigenvalues of (1.1 - 1.3). Further ∆ is entire in λ and the order of a zero, λ, of ∆ is equal to the algebraic multiplicity of λ as an eigenvalue of (1.1 - 1.3), cf. [17, pages 16 - 20]. Definition 2.2 Two functions f and g, analytic in a neighbourhood of λ0 ∈ C, are said to have contact of order k at λ0 if f (i) (λ0 ) = g (i) (λ0 ), i = 0, . . . , k, f (k+1) (λ0 ) 6= g (k+1) (λ0 ). We are now ready to characterize the algebraic multiplicity of real eigenvalues by the order of contact between the eigencurves µ = µk (λ) and the parabola µ = −λ2 . Theorem 2.3 A point λ0 ∈ R is an eigenvalue of (1.1 - 1.3) of algebraic multiplicity m if, and only if, there is an integer k ≥ 0 for which µ = µk (λ) and µ = −λ2 have contact of order m − 1 at λ0 . In addition (1.1 - 1.3) has countably many eigenvalues, they have no finite accumulation point and each eigenvalue has finite algebraic multiplicity. Proof: Firstly, it is clear that λ0 is an eigenvalue of (1.1 - 1.3) if, and only if, there is k ≥ 0 for which µk (λ0 ) = −λ20 . Now the algebraic multiplicity of λ0 is finite, since if not, the charactersitic determinant ∆, (2.7), would vanish identically and every λ ∈ C would be an eigenvalue. In which case, for each λ = iθ, θ ∈ R, we would be able to solve T y − iθSy = −θ2 y non-trivially, with solution, say, yθ . This gives (T yθ + θ2 yθ , yθ ) = iθ(Syθ , yθ ) where, for large enough θ, the left hand side is positive while the right hand side is pure imaginary or zero, giving that indeed the algebraic multiplicity of λ0 is finite. In addition, the fact that ∆ is not identically zero shows that (1.1 - 1.3) has at most countably many eigenvalues and they have no finite accumulation point. Thus for λ0 we have a finite chain of equations (2.5) as described in Definition 2.1 (ii). Suppose λ0 to have algebraic multiplicity m ≥ 1. From the solution y(λ, x) of (1.1) obeying the initial conditions (2.6), an eigenfunction and chain of associated functions, as specified in Definition 2.1 (ii), may be constructed by taking ϕ0 (x) = y(λ0, x),   1 ∂ i y(λ, x) i ϕ (x) = , i! ∂λi λ=λ0 5

i = 1, . . . , m − 1,

see [17, page 17]. Here is should be noted that

h

∂ i y(λ,x) ∂λi

i

λ=λ0

is in D(T ) for i = 0, . . . , m − 1

but does not obeyhthe boundary condition (1.3) for i = m, since the order of the zero of ∆ i ∂ m y(λ,x) 6∈ D(T ). A direct computation now gives at λ0 is m. Thus ∂λm λ=λ0

(T − λ0 S)y (n) (λ0 , x) − nSy (n−1) (λ0 , x) − [λ2 y](n) (λ0 , x) = 0,

n = 0, 1, . . . , m − 1, (2.8)

n

∂ y (n−1) := 0. where y (n) (λ, x) := ∂λ n (λ, x) and in the case of n = 0, nSy On the other hand, µk (λ) is analytic and the solutions z(λ, x) to

Q(λ)z := (T − λS + µk (λ))z = 0

(2.9)

can also be taken to be analytic in a neighbourhood of λ0 , see [14, Theorem V.ii.3.9, page 392]. Since the the geometric multiplicity of the eigenvalues of (1.1 - 1.3) is 1, we may, without loss of generality, take z(λ0 , x) = y(λ0, x). Further, we may differentiate both sides of (2.9) with respect to λ arbitrarily often, to obtain   i X 1 ∂j 1 ∂ i−j Q(λ) z(λ, x) = 0, j! ∂λj (i − j)! ∂λi−j λ=λ0 j=0

i = 0, 1, 2, . . . .

(2.10)

Here z (n) (λ0 , x) ∈ D(T ) for all n = 0, 1, . . . , and a direct computation yields (T − λ0 S)z (n) (λ0 , x) − nSz (n−1) (λ0 , x) + [µk z](n) (λ0 , x) = 0,

n = 0, 1, . . . .

(2.11)

Suppose that the order of contact between µ = −λ2 and µ = µk (λ) at λ0 is p ≥ 0. Setting ψ 0 (x) = z(λ0 , x),   1 ∂ i z(λ, x) i , ψ (x) = i! ∂λi λ=λ0

i = 0, 1, . . . ,

(j)

and replacing µk (λ0 ) by −[λ2 ](j) |λ=λ0 for j = 0, 1, . . . , p one obtains from (2.10) and the Jordan chain condition (2.5) that ψ 0 , ψ 1 , . . . , ψ p is a Jordan chain of (1.1 - 1.3) at λ0 of length p + 1. Since the algebraic multiplicity of λ0 as an eigenvalue of (1.1 - 1.3) is m, it follows that p + 1 ≤ m. Hence the order of contact between µ = −λ2 and µ = µk (λ) at λ0 is at most m − 1. Taking the inner product of (2.8) for n = 0 with z(λ0 , x) and the inner product of y(λ0, x) with (2.11) combined with the self-adjointness of T − λ0 S we obtain λ20 ky(λ0 , x)k2 = = = =

λ20 (y, z)|λ=λ0 ((T − λ0 S)y, z)|λ=λ0 (y, (T − λ0 S)z)|λ=λ0 −µk (λ0 )(y, z)|λ=λ0 = −µk (λ0 )ky(λ0 , x)k2 . 6

Hence µ = −λ2 and µ = µk (λ) have order of contact at least 0 at λ0 . Suppose that µ = −λ2 and µ = µk (λ) have order of contact at least p − 1 ≤ m − 2 at λ0 . Taking the inner product of (2.8) for n = p − j with z (j) (λ0 , x) and the inner product of y (p−j)(λ0 , x) with (2.11) for n = j, j = 0, 1, . . . , p, combined with the self-adjointness of T − λ0 S we obtain

((p − j)Sy (p−j−1) + [λ2 y](p−j), z (j) )|λ=λ0 = ((T − λ0 S)y (p−j), z (j) )|λ=λ0 = (y (p−j), (T − λ0 S)z (j) )|λ=λ0 = (y (p−j), jSz (j−1) − [µk z](j) )|λ=λ0 .   p Multiplying the above equation by and summing over j = 0, 1, . . . , p, we come to j   p  p  X X p p 2 (p−j) (j) (y (p−j), [µk z](j) )|λ=λ0 , (2.12) ([λ y] , z )|λ=λ0 = − j j j=0

j=0

where we have used the elementary computation   p−1 p X X p! p (p−j−1) (j) (p − j) (p − j) (Sy ,z ) = (Sy (p−j−1), z (j) ) j j!(p − j)! j=0

j=0

p X

p! (Sy (p−k), z (k−1) ) k!(p − k)! k=1   p X p (Sy (p−k), z (k−1) ). k = k

=

k

k=0

Expanding the derivatives in (2.12) at λ = λ0 , we see that     j  p X p−j  p X X X p j p p−j (j−r) 2 (r) (p−j−r) (j) µk (y (p−j), z (r) ). [λ ] (y ,z ) = − j r j r j=0 r=0

j=0 r=0

After re-arrangement this can be written as p p−j  X p − j  p  X (r) ([λ2 ](r) + µk )(y (p−j−r), z (j) )|λ=λ0 . 0= r j j=0 r=0

Since the order of contact between µ = −λ2 and µ = µk (λ) at λ = λ0 is at least p − 1 the above summation reduces to (p)

0 = ([λ2 ](p) + µk )(y, z)|λ=λ0 . Thus

(p)

0 = ([λ2 ](p) + µk )|λ=λ0 and the order of contact between µ = −λ2 and µ = µk (λ) at λ = λ0 is at least p. This inductive process can be applied repeatedly until the order of contact p = m − 1. 7

3

Main Results

We commence by showing that the number of non-real eigenvalues (counted according to algebraic multiplicity) is finite. √ Theorem 3.1 If λ is a non-real eigenvalue for (1.1 - 1.3) then |λ| ≤ −τ0 . Proof: Suppose λ = a + ib to be a non-real eigenvalue (so b 6= 0) for (1.1 - 1.3) with eigenfunction y normalised so that (y, y) = 1. Then from (1.1) multiplied throughout by y¯, we have (T y, y) − λ(Sy, y) = λ2 . (3.1) From the imaginary parts of (3.1) we see that

−b(Sy, y) = 2ab, and since b 6= 0, (Sy, y) = −2a. Now the real parts of (3.1) yield (T y, y) − a(Sy, y) = a2 − b2 , and thus (T y, y) + a2 + b2 = 0. Consequently which establishes the result.

|λ|2 = −(T y, y) ≤ −τ0 ,

Corollary 3.2 (i) The number of non-real eigenvalues for (1.1 - 1.3), counted according to algebraic multiplicity, is finite. (ii) If T ≥ 0 (i.e. τ0 ≥ 0) then there are no non-real eigenvalues for (1.1 - 1.3). (iii) If T ≥ 0 and λ = 0 is an eigenvalue of (1.1 - 1.3), then λ = 0 has algebraic multiplicity at most 2. Proof: From Theorem 2.3, the eigenvalues when counted according to algebraic multiplicity have no finite point of accumulation. This along with Theorem 3.1 is enough to establish (i) while (ii) is an immediate consequence of Theorem 3.1. For (iii), note that the hypotheses imply that µ0 (0) = 0 = τ0 . Further, µ0 being con2 vex, we have ddλµ20 (0) ≥ 0, while the parabola µ = −λ2 has −2 as its second derivative. Theorem 2.3 completes the argument. We shall say that a real eigenvalue λ has oscillation number n if the eigenfunction associated with λ has n zeros in (0, 1). Our first oscillation result shows that ultimately all eigenvalues are real and algebraically simple and further, we find the “Richardson index”, beyond which we can determine the oscillation number associated with each eigenvalue. 8

Theorem 3.3 (i) Let k + be the first integer k for which   2 ss s − 2¯ . τk > max 0, 4

(3.2)

Then for each n ≥ k + there is exactly one positive eigenvalue, λ+ n > 0, with oscillation number n, in addition it is algebraically simple. (ii) Let k − be the first integer k for which s¯2 − 2¯ ss τk > max 0, 4 



.

Then for each n ≥ k − there is exactly one negative eigenvalue, λ− n < 0, with oscillation number n, in addition it is algebraically simple. (iii) Let K be the first integer for which   2 ss s¯2 − 2¯ ss s − 2¯ . , τk > max 0, 4 4 − Then for each integer n ≥ K there are exactly two eigenvalues, λ+ n > 0 and λn < 0, each of which has oscillation number n, in addition, both are algebraically simple and − + − λ+ n+1 > λn and λn+1 < λn .

Proof: (i) First recall that eigencurve theory, [6, 7, 9], shows that for any k, and all λ ∈ R, dµk ≤ s¯ s≤ dλ so that if k is large enough to ensure τk > 0, it follows from the fact that µk (0) = −τk that −τk + sλ ≤ µk (λ) ≤ −τk + s¯λ,

for all λ ≥ 0.

Thus for λ > 0 there will be intersections of the eigencurve µk (λ) with the parabola µ = −λ2 all of which will fall within the real interval     1 p p 1 (3.3) −¯ s + s¯2 + 4τk , −s + s2 + 4τk . 2 2 Further, if we demand that k be so large that p s¯ − s¯2 + 4τk < s,

(3.4)

then, for all λ in the interval (3.3), the slope of the parabola is less than that of the eigencurve and hence there is only one intersection between the parabola and the eigencurve. At this intersection, the order of contact between the parabola and the eigencurve is 0 9

making, by Theorem 2.3, the eigenvalue algebraically simple. It is easy to see that (3.4) is equivalent to (3.2). The argument for (ii) is similar, while (iii) is a combination of (i) and (ii). From the analysis above we can produce asymptotic estimates for the eigenvalues. These estimates vary in accuracy according to the degree of smoothness assumed on the coefficient p. We shall write Z 1

p−1/2 (t) dt

σ=

0

and recall that ν denotes half the number of Dirichlet conditions in (1.2, 1.3). Theorem 3.4 (i) We have λ± n = ±

nπ + o(n) σ

as n → ±∞.

(ii) If p is absolutely continuous, then ± ± ± λ± n ∈ In := [an , bn ]

as n → ±∞,

where (n + ν)π s¯ − + o(1), σ 2 (n + ν)π s − + o(1). = ± σ 2

a± = ± n b± n

(iii) If p′ is absolutely continuous, the o(1) in (ii) may be replaced by O(1/n). (iv) If p is constant, so that p = σ −2 , then     Z 1 1 1 s¯2 σ s¯ (n + ν)π ∗ ∗ + − cot β + cot α + +o q+ + an = − + 2 σ (n + ν)π 2 0 8 n ± with corresponding expressions for a− n , bn .

Proof: For λ+ n , combining (2.1) and (3.3) we obtain (i), while (2.2) and (3.3) give (ii) and (iii). The case of λ− n is similar. For (iv) we have again that   √ σγ 1 (n + ν)π , + +o τn = σ (n + ν)π n where

1 γ = − cot β + cot α + 2 ∗



10

Z

0

1

q(t) dt,

and so s¯ 1 p 2 a+ s¯ + 4τn n = − + 2 2    s¯2 1 s¯ √ +O = − + τn 1 + 2 8τn n4   s¯2 s¯ √ 1 = − + τn + √ + O 2 8 τn n3   s¯ (n + ν)π s¯2 σγ 1 h i = − + + + +o (n+ν)π 2 σ (n + ν)π n −1 8 + O(n ) σ   σγ s¯2 σ 1 s¯ (n + ν)π + + +o = − + 2 σ (n + ν)π 8(n + ν)π n from which the result follows. The analysis for the other end points of the intervals In± is similar. Theorem 3.5 Assume that p is absolutely continuous. Let C be the number of non-real eigenvalues counted according to algebraic multiplicity and let N be the number of negative eigenvalues of T . Then C ≤ 2N. Proof: We employ a homotopy argument by replacing the coefficient s in (1.1) by δs, where δ, the homotopy parameter, ranges through [0, 1]. The eigenvalues now depend on δ and we consequently write λ(δ) to display this. For δ = 0 it is clear that the eigenvalues are the solutions to λ2 (0) = τk , k ≥ 0,

so that the number of non-real eigenvalues λ(0) is twice the number of values of k for which τk < 0, i.e. for δ = 0, C = 2N. The Richardson index, K, of Theorem 3.3 now depends on δ, but with s¯, s replaced by δ¯ s, δs respectively. It is easy to see that the index takes its maximum value at δ = 1 and this maximum will be taken as the index, K, here. Take n ≥ K and large enough to ensure − that the circle |λ| = max{λ+ n , λn } contains all non-real √ eigenvalues of (1.1 - 1.3), i.e. we − −τ0 . Since n ≥ K it follows that take n large enough to ensure that max{λ+ , λ } > n n + − − λ+ n+1 (δ) > λn (δ) and λn+1 (δ) < λn (δ) for all 0 ≤ δ ≤ 1.

We now construct a contour Rδ in the complex plane, rectangular in shape, with the two sides parallel to the imaginary axis given by ± λ± n (δ) + λn+1 (δ) + iθ, 2

√ √ − −τ0 − 1 ≤ θ ≤ −τ0 + 1.

Then Rδ contains all the complex eigenvalues for (1.1 - 1.3) for any value of δ and the total number of eigenvalues counted according to algebraic multiplicity within Rδ remains 11

constant in δ. At δ = 0 this number is readily seen to be 2(n + 1). Now at δ = 1, since + n ≥ K, the curves µ = −λ2 and µ = µn (λ) for λ− n ≤ λ ≤ λn form the boundary of a closed region, say B, in the (λ, µ)-plane. Since µj (λ) < µj+1 (λ) for real λ, the points (0, −τk ), k = N, N + 1, . . . , n − 1, are in B and the continuous curves (λ, µk (λ)), k = N, N + 1, . . . , n − 1, + − each intersect µ = −λ2 in at least two points with λ− n < λ < λn . Hence, along with λn and λ+ n , we have located at least 2(n + 1 − N) real eigenvalues in R1 . Thus the number of non-real eigenvalues is the number of non-real eigenvalues in R1 which is less than or equal to 2(n + 1) − 2(n + 1 − N). Finally, we come to our main oscillation theorem. Theorem 3.6 Assume that p is absolutely continuous and that T ≥ 0. Then for each − n ≥ 1 there are exactly two real simple eigenvalues λ+ n > 0, λn < 0 with oscillation number n. Additionally, one of the following holds: − if T > 0: there are two real simple eigenvalues λ+ 0 > 0, λ0 < 0 with oscillation number 0;

if T 6> 0: there are two simple real eigenvalues λ0 , λ′0 with oscillation number 0 one of which is 0; or λ = 0 is an eigenvalue of algebraic multiplicity 2 with oscillation number 0. Proof: We use the homotopy argument constructed for the proof of the previous theorem noting that we now have N = 0. As before the rectangle Rδ contains exactly 2(n + 1) eigenvalues all of which are real. We count at least 2n intersections of the parabola µ = −λ2 with the curves µk (λ), k = 1, . . . , n and 2 more between the parabola and µ0 (λ) in case τ0 > 0 thereby establishing the claim when T > 0. 0 If τ0 = 0, then we note that µ0 (0) = 0 and either dµ (0) is non-zero - in which case µ0 dλ has two intersections with the parabola both of which must be simple so as to produce a 0 total algebraic count of 2(n + 1) eigenvalues within Rδ - or dµ (0) = 0 in which case λ = 0 dλ is an eigenvalue of algebraic multiplicity 2 with oscillation number 0.

References [1] F. V. Atkinson, H. Langer and R. Mennicken, Sturm-Liouville problems with coefficients which depend analytically on the eigenvalue parameter, Acta Sci. Math. (Szeged), 57 (1993), 25 - 44. [2] P. A. Binding, On generalised and quadratic eigenvalue problems, Applicable Analysis, 12 (1981), 27 -45. [3] P. A. Binding and P. J. Browne, Spectral properties of two-parameter eigenvalue problems, Proc. Roy. Soc. Edinburgh, 89A (1981), 157 - 173. [4] P. A. Binding and P. J. Browne, Spectral properties of two parameter eigenvalue problems II, Proc. Roy. Soc. Edinburgh, 106A, (1987), 39 - 51. 12

[5] P. A. Binding and P. J. Browne, Eigencurves for two-parameter self-adjoint ordinary differential equations of even order, J. Differential Equations, 79 (1989), 289 - 303. [6] P. A. Binding and P. J. Browne, Asymptotics of eigencurves for second order ordinary differential equations I, J. Differential Equations , 88 (1990), 30 - 45. [7] P. A. Binding and P. J. Browne, Asymptotics of eigencurves for second order ordinary differential equations II, J. Differential Equations I, 89 (1991), 224 - 233. [8] P. A. Binding, P. J. Browne and R. Picard, Corrigendum: Spectral properties of two parameter eigenvalue problems II, Proc. Roy. Soc. Edinburgh, 115A, (1990), 87 - 90. [9] P. A. Binding, P. J. Browne and K. Seddighi, Sturm-Liouville problems with eigenparameter dependent boundary conditions, Proc. Edin. Math. Soc., 37 (1993), 57 - 72. [10] P. J. Browne, Two-parameter eigencurve theory, Proceedings 1988 Dundee Conference on Ordinary and Partial Differential Equations, Longman, Harlow, U.K., 52 59. [11] J. Eisenfeld, Quadratic eigenvalue problems, J. Math. Anal. Appl., 23 (1968), 58 70. [12] M. G. Gasymov and G. SH. Guseinov, Determination of diffusion operator on spectral data, Dokl. Akad. Nauk. Azerb. SSR, 37 (1981), 19 - 23. [13] G. SH. Guse˘ınov, On spectral analysis of a quadratic pencil of Sturm-Liouville operators, Soviet Math. Dokl., 32 (1985), 859 - 862. [14] T. Kato, Perturbation Theory for Linear Operators, Springer-Verlag, New York, 1966. [15] H. Koyunbakan and E. S. Panakhov, Half-inverse problem for diffusion operators on the finite interval, J. Math. Anal. Appl., 326 (2007), 1024 - 1030. [16] P. Lancaster and A. Shkalikov, Damped vibrations of beams and related spectral problems, Canadian Applied Mathematics Quarterly, 2 (1994), 45 - 90. [17] M. A. Naimark, Linear Differential Operators, Part I, Frederick Ungar Publishing Company, New York, 1967. [18] B. Najman, Eigenvalues of the Klein-Gordon equation, Proc. Edin. Math. Soc., 26 (1983), 181 - 190.

13

[19] G. F. Roach and B. D. Sleeman, On the spectral theory of operator bundles, Applicable Analysis, 7 (1977), 1- 14. [20] G. F. Roach and B. D. Sleeman, On the spectral theory of operator bundles II, Applicable Analysis, 9 (1979), 29 - 36. [21] A.A. Shkalikov, Boundary problems for ordinary differential equations with parameter in the boundary conditions, Trudy Seminara imeni I. G. Petrovskogo, 9 (1983), 190 - 229. [22] R. E. L. Turner, Some variational principles for a nonlinear eigenvalue problem, J. Math. Anal. Appl., 17 (1967), 151 - 160. [23] H. F. Weinberger, On a nonlinear eigenvalue problem, J. Math. Anal. Appl., 21 (1968), 506 - 509.

14

Oscillation Theory for a Quadratic Eigenvalue Problem

Sep 15, 2008 - For example, Roach and Sleeman [19, 20] recast (1.1. - 1.3) as a linked two parameter system in L2(0, 1)⊗C2 and set their completeness results in this space. Binding [2] establishes the equivalence of L2(0, 1)⊗C2 with L2(0, 1)⊕L2(0, 1) and gives abstract eigenvector completeness results in this latter ...

152KB Sizes 0 Downloads 214 Views

Recommend Documents

Quadratic eigenvalue problems for second order systems
We consider the spectral structure of a quadratic second order system ...... [6] P.J. Browne, B.A. Watson, Oscillation theory for a quadratic eigenvalue prob-.

RENORMALIZED OSCILLATION THEORY FOR ...
solutions, and then turn to renormalized oscillation theory in terms of .... Schmidt operator with a semi-separable integral kernel depending on an energy parameter ...... Next, we recall the structure of the resolvent and Green's function of Ta,.

A SECOND EIGENVALUE BOUND FOR THE ...
will be presented in Section 5. .... 5 operators are defined by their quadratic forms. (5) h±[Ψ] = ∫. Ω. |∇Ψ(r)|2 ..... The first eigenfunction of −∆+V on BR is radi-.

Schwarz symmetric solutions for a quasilinear eigenvalue ... - EMIS
We begin with the abstract framework of symmetrization following Van Schaftingen [13] and in Section ...... quasi-linear elliptic problems, Electron. J. Differential ...

The Quadratic Gaussian CEO Problem with Byzantine ...
H. If there is exactly one such a sequence, call it ˆUn k [H] for ... among all H k, then call this sequence ˆUn k . ..... 46th Allerton Conf. on Comm.,. Control and ...

EIGENVALUE ESTIMATES FOR STABLE MINIMAL ...
KEOMKYO SEO. (Communicated by H. Martini). Abstract. In this article ... negative constants, we can extend Theorem 1.1 as follows. THEOREM 1.3. Let M be an ...

Quadratic Transformations
Procedure: This activity is best done by students working in small teams of 2-3 people each. Develop. 1. Group work: Graphing exploration activity. 2.

A quadratic string adapted barrier exploring method for ...
Feb 8, 2007 - This paper proposes a novel method for locating relevant transition states that contain crucial information on rare events of transition from the energy barriers, such as ionic diffusion in a crystalline material with vacancies and adat

Reduced Order Models for Eigenvalue Problems
number of input and output variables are denoted by m and p respectively. ... that sC − G is a regular pencil and that s0 is chosen such that G + s0C is.

Truncated Power Method for Sparse Eigenvalue ...
Definition 1 Given a vector x and an index set F, we define the truncation ..... update of x and the update of z; where the update of x is a soft-thresholding ...

Galois Theory and Noether's Problem
We are particularly interested in finite algebraic extensions obtained by adding roots of irreducible polynomials to the field of rational numbers. Galois groups give information about such field extensions and thus, information about the roots of th

The generalized principal eigenvalue for Hamilton ...
large, then his/her optimal strategy is to take a suitable control ξ ≡ 0 which forces the controlled process Xξ to visit frequently the favorable position (i.e., around ..... this section, we collect several auxiliary results, most of which are f

Using the Eigenvalue Relaxation for Binary Least ...
Abstract. The goal of this paper is to survey the properties of the eigenvalue relaxation for least squares binary problems. This relaxation is a convex program ...

Specialized eigenvalue methods for large-scale model ...
High Tech Campus 37, WY4.042 ..... Krylov based moment matching approaches, that is best .... from sound radiation analysis (n = 17611 degrees of free-.

Autonomous Oscillation Control Loop Design for ... - IEEE Xplore
Abstract—This paper suggests an autonomous oscillation con- trol loop for frequency read-out-type resonant sensors that pro- duces outputs of variable ...

Multitask Generalized Eigenvalue Program
School of Computer Science. McGill University ... Although the GEP has been well studied over the years [3], to the best of our knowledge no one has tackled the ...

A constrained quadratic spline as a model for the ...
Probability distributions traditionally used in reliability analysis (e.g., exponential, Weibull, ..... to be a better life variable for obvious engineering reasons.

Four limit theorems for quadratic functionals of ...
Dedicated to Miklós Csörg˝o on the occasion of his 70th birthday. Abstract. We generalize and give new proofs of four limit theorems for quadratic functionals of ...

Four limit theorems for quadratic functionals of ...
well as classic theorems about weak convergence of Brownian function- als. ... as well as some stochastic version of Fubini's theorem (see [2]), and asymptotic.

Sparse Filter Design Under a Quadratic Constraint: Low ...
Examples in wireless channel equalization and minimum-variance ... area, or the supply voltage may be lowered to take advantage of a slower computation rate ...

The Madden–Julian Oscillation in CCSM4
Dec 15, 2011 - Unlike its predecessor, CCSM4 reproduces a number of aspects of MJO ...... U* is the zonal wind anomaly, averaged over the region from 408E to .... verse relationship between the two modes (Webster and. Yang 1992; Ju ...