705
Views
18
CrossRef citations to date
0
Altmetric
Articles

Trace formula and inverse nodal problem for a conformable fractional Sturm-Liouville problem

&
Pages 524-555 | Received 30 Jul 2018, Accepted 29 Apr 2019, Published online: 25 May 2019

ABSTRACT

In this paper, we have developed the spectral theory for a conformable fractional Sturm-Liouville problem Lα(q(x),h,H) with boundary conditions which include conformable fractional derivatives of order α, 0<α1, and prove a completeness theorem and an expansion theorem. We obtain the canonical infinite product constructed by the zero set of the characteristic function of Lα. Also, we calculate the regularized trace formula of the eigenvalues and investigate the inverse nodal problem for this problem with real-valued coefficients on a finite interval. The oscillation of the eigenfunctions for sufficiently large n is established, and an asymptotic formula for elements constructed by of nodal points is obtained. The uniqueness theorem is proved, and an effective procedure for solving the inverse problem is given. Finally, we present two examples to illustrate our theoretical findings.

1. Introduction

The beginning of the fractional calculus is considered to be Leibniz's letter to L'Hôspital in 1695. L'Hôspital asked, ‘what does it mean dnfdxn if n=12?’. Since then, various types of fractional derivatives have been introduced, most of which use an integral form with different singular kernels for the fractional derivative. Due to the same reason, these fractional derivatives inherit some non-local behaviours which lead them to numerous interesting applications, including memory effects and future dependence. However, almost all the fractional derivatives used in this literature, such as Grünwald-Letnikov, Riemann-Liouville, Caputo and Jumarie, Marchaud and Riesz, fail to satisfy some of the basic properties owned by usual derivatives, e.g. the product rule, chain rule, Rolle's theorem, mean value theorem, composition rule and semigroup property. These inconsistencies lead to the development of the local fractional derivative whose most properties coincide with classical integer derivative. Recently, Khalil, Al Horani, Yousef and Sababhehb [Citation1] introduced a new local fractional derivative and fractional integral. This simple, well-behaved definition enables us to prove many properties analogous to those of integer-order derivative, probably because it includes a limit, instead of an integral. More basic properties and main results on conformable derivative can be found in [Citation2,Citation3]. Despite the numerous attractive properties the conformable derivative has, it has drawbacks and some unusual properties, e.g. the zero order derivative of a function does not return the function and the index law does not hold, that is, DαDβf(t)Dα+βf(t) for any α and β. Some authors (See [Citation4]) have argued that conformable fractional derivative is not a truly fractional operator. This question seems today to still be open and perhaps it is a philosophical issue. Because of its effectiveness and applicability, conformable derivative has been employed in various field such as the control theory of dynamical systems [Citation5], Newton mechanics [Citation6,Citation7], quantum mechanics [Citation8], variational calculus [Citation9], arbitrary time scale problems [Citation10,Citation11], anomalous diffusion [Citation12], diffusive transport [Citation13,Citation14] and stochastic process [Citation15]. Recently, a physical interpretation of the conformable derivative is provided by Zhao and Luo [Citation16]. They generalized the definition of the conformable derivative to general conformable derivative by means of linear extended Gaˆteaux derivative, and used this definition to demonstrate that the physical interpretation of the conformable derivative is a modification of classical derivative in direction an dmagnitude. So, in our opinion, the interest in the conformable derivative has been increasing and it is worthwhile to explore in this new research area.

In recent years, much effort has focussed on a fractional generalization of the well-known Sturm-Liouville problems (See [Citation17–20]). Fractional Sturm-Liouville problems (FSLPs) are different from those usually defined in this literature, i.e. the ordinary derivatives in a traditional Sturm-Liouville problem are replaced with fractional derivatives or derivatives of fractional order. These types of FSLPs arise in various areas of science and in many fields in engineering, we refer to [Citation21–25]. Now, we consider on the closed interval [0,π] the boundary value problem Lα(q(x),h,H): (1) αy:=DxαDxαy+q(x)y=λy,0<x<π,(1) with boundary conditions (2) U(y):=Dxαy(0)hy(0)=0,V(y):=Dxαy(π)+Hy(π)=0,(2) known as a conformable fractional Sturm-Liouville problem (CFSLP). Here, λ is the spectral parameter, q(x), h and H are real, and q(x)Lα2(0,π). Where Dxα is the CF derivative of order α, 0<α1. The operator α is called the CFSL operator.

The present paper is composed of two main parts. In the first part, we prove the existence of infinitely many real eigenvalues of the CFSLP which are simple, and that the corresponding eigenfunctions are orthogonal (See parts 1, 2 and 3 of Theorem 3.7), and then, we provide the asymptotic behaviour of eigenvalues and the corresponding eigenfunctions (Theorem 3.10). The proof concerning the expansion theorem is established by means of contour integration and the calculus of residues (Theorem 3.12). Also, we obtain the canonical infinite product constructed by the zero set of the characteristic function of Lα (Theorem 3.11). Then, we present an explicit formula of the regularized trace regarding the boundary condition parameters and the potential for the problem (Equation1)–(Equation2) by using Levitan's method as in [Citation26] (See Theorem 4.3). The study of regularized traces for the Sturm-Liouville operator on a finite interval was initiated in [Citation27]. Later on, some authors turned their attention to trace theory and obtained exciting results [Citation28–31]. This research continued in many respects, such as Dirac operators, the matrix Schrödinger equation with energy-dependent potential, the linear damped wave equation, differential operators with abstract operator-valued coefficients, and the case of matrix-valued Sturm-Liouville operators, etc. (See for instance, [Citation32–36]). In the second part, we investigate the inverse nodal problem for the CFSLP (Equation1) with boundary conditions (Equation2) which includes the CF derivative of order α, 0<α1. At first, we obtain a detailed asymptotic formula for elements constructed by of nodal points (Lemma 5.1). Then, we demonstrate that the n-th eigenfunction of Lα has precisely n nodes in the interval (0,π) for sufficiently large n, which is an analogue of the classical Sturm's oscillation theorem for the Sturm-Liouville operator (Theorem 5.2). Further, we prove a uniqueness theorem and provide two algorithms to reconstruct the potential function q and the parameters h, H in the boundary conditions from elements constructed by of nodes and the length of those elements (See Theorem 5.5 and Theorem 5.9). The inverse nodal problem, which is different from the classical inverse spectral theory of Gelfand and Levitan [Citation37], was started by McLaughlin [Citation38]. Later, Hald and McLaughlin [Citation39] and Browne and Sleeman [Citation40] proved that it is sufficient to know the nodal points to determine the potential function of the regular Sturm-Liouville problem uniquely. Yang [Citation41] presented an algorithm to recover q from a dense subset of nodal points. Furthermore, the inverse nodal Sturm-Liouville problems have been considered recently in [Citation42–48].

This paper is organized as follows. In Section 2 some basic definitions and properties of CF calculus are given. Section 3 is devoted to the study of the spectrum of the CFSL operator. The regularized traces for the α operator with boundary conditions is calculated in Section 4. Our main results are given in Section 5. We discuss the inverse nodal problem of reconstructing the potential from an arbitrary dense set of elements constructed by of nodes of the boundary value problem Lα. Finally, we summarize our results in Section 6.

2. Conformable fractional preliminaries

Here, we give some basic definitions and properties of the conformable fractional calculus theory which can be found in [Citation1,Citation2].

Definition 2.1

Let f:[0,)R be a given function. Then, the conformable fractional derivative of f of order α is defined by: (3) Dαf(x)=limh0f(x+hx1α)f(x)h,Dαf(0)=limx0+Dαf(x),(3) for all x>0, α(0,1]. Note that if f is differentiable, then (4) Dαf(x)=x1αf(x),(4) where f(x)=limh0[f(x+h)f(x)]/h. If Dαf(x0) exists and is finite, we say that f is α-differentiable at x0. From now on, assume that DxαDα (i.e. the CF derivative of order α with respect to x).

Theorem 2.2

If a function f:[0,)R is α-differentiable at x0>0. Then, f is continuous at x0.

Definition 2.3

Let f:[0,)R be a given function. Then, the conformable fractional integral of f of order α is defined by: (5) Iαf(x)=0xf(t)dαt=0xtα1f(t)dt,(5) for all x>0. Where the integral is the usual Riemann improper integral.

Theorem 2.4

Let f, g be α-differentiable at x, x>0. Then

  1. Dxα(af+bg)=aDxαf+bDxαg,  a,bR,

  2. Dxα(xp)=pxpα,  pR,

  3. Dxα(c)=0, (c is a constant),

  4. Dxα(fg)=Dxα(f)g+fDxα(g),

  5. Dxα(f/g)=Dxα(f)gfDxα(g)g2.

Lemma 2.5

Let f:[a,)R be any continuous function. Then, for all x>a, we have DxαIαf(x)=f(x).

Lemma 2.6

Let f:(a,b)R be differentiable. Then, for all x>a, we have IαDxαf(x)=f(x)f(a).

Theorem 2.7

α-chain rule

Let f,g:(0,)R be α-differentiable functions and h(x)=f(g(x)). Then, h(x) is α-differentiable and for all x (x0 and g(x)0) and (Dxαh)(x)=(Dxαf)(g(x)).(Dxαg)(x).g(x)α1. If x=0, then (Dxαh)(0)=limx0+(Dxαf)(g(x)).(Dxαg)(x).g(x)α1.

Proposition 2.8

Let 0<α, β1 such that 1<α+β2 and f be a function defined on [0,). Then, the following semigroup properties hold:

  1. If f be twice differentiable on (0,), then (DxαDxβf)(x)=Dxα+βf(x)+(1β)xβDxαf(x).

  2. (IαIβf)(x)=xββIαf(x)+1βIα+βf(x)xβ0xtα+β2f(t)dt.

Note that if α,β1, then (D1D1f)(x)=D2f(x)=f(x), (I1I1f)(x)=I2f(x).

Theorem 2.9

α-integration by parts

Let f,g:[a,b]R be two functions such that fg is differentiable. Then (6) abf(x)Dxαg(x)dαx=fg|ababg(x)Dxαf(x)dαx.(6)

Definition 2.10

The space Cαn[a,b] consists of all functions defined on the interval [a,b] which are continuously α-differentiable up to order n.

Definition 2.11

Let 1p<, a>0. The space Lαp(0,a) consists of all functions f:[0,a]R satisfying the condition (0a|f(x)|pdαx)1/p<.

Lemma 2.12

The space Lαp(0,a) associated with the norm function fp,α:=0a|f(x)|pdαx1/p, is a Banach space. Moreover, if p=2, then Lα2(0,a) associated with the inner product f,g:=0af(x)g(x)¯dαx, (f,gLα2(0,a)), is a Hilbert space (See [Citation49]).

Definition 2.13

Let pR be such that p1. The Sobolev space Wαp(0,a) consists of all functions defined on the interval [0,a], such that f(x) is absolutely continuous, and Dxαf(x)Lαp(0,a) (See [Citation49] for more details).

Lemma 2.14

α-Leibniz rule

Let tα1f(x,t) and tα1fx(x,t) be continuous in x on some regions of the (x,t)-plane, including a(x)tb(x), x0xx1. If a(x) and b(x) are both α-differentiable for x0xx1, then Dxαa(x)b(x)f(x,t)dαt=Dxαb(x)f(x,b(x))b(x)α1Dxαa(x)f(x,a(x))a(x)α1+a(x)b(x)Dxαf(x,t)dαt.

Proof.

From the Leibniz rule, the proof is clear.

Lemma 2.15

Assume that u(x) and v(x) be real-valued, nonnegative on an interval [0,π], u(x) is a continuous function, and v(x)Lα(0,π). If u(x)c+0xv(x)u(x)dαt, where c is a non negative constant, then u(x)cexp(0xv(x)dαt).

Proof.

This lemma can be proved by using an argument similar to that for Lemma 3.2.1. in [Citation30].

3. Conformable fractional Sturm-Liouville problems

In this section, we collect known results about the spectra of the regular conformable fractional Sturm-Liouville operator on a finite interval. Note that more general second-order equations (7) Dxαp(x)Dxαy+{l(x)+λr(x)}y=0,0<α1,(7) where the function r(x) is positive on [a,b], are reducible to the form (Equation1). If we assume that r(x)Cα2[a,b] and p(x)Cα1[a,b], then (Equation7) by the change of variables such as those of Joseph Liouville (8) tαα=0xr(s)dαs,w(t)=y(x)r(x)4exp120xp(s)dαs,(8) can be reduced to the canonical form DtαDtαw+q(t)w=λw, where q(t)=DxαDxαr(x)4r2(x)516(Dxαr(x))2r3(x)1r(x)l(x)+12Dxαp(x)+14p2(x).

3.1. Basic properties of the operator

Definition 3.1

Let y(x),z(x) be α-differentiable functions on [0,π]. The fractional Wronskian of y and z is defined as: (9) Wα[y,z]:=y(x)z(x)Dxαy(x)Dxαz(x)=y(x)Dxαz(x)z(x)Dxαy(x).(9)

Lemma 3.2

Two solutions y, z of Equation (Equation1) defined an (0,π) are linearly independent if and only if Wα[y,z](x)0 for all x(0,π) (See [Citation50]).

Theorem 3.3

For c1,c2C, Equation (Equation1) has a unique solution in Cα2[0,π] which satisfies (10) ϕ(0,λ)=c1,Dxαϕ(0,λ)=c2,(λC).(10) Moreover, ϕ(x,λ) is an entire function of λ for all x[0,π].

Proof.

The proof of this theorem is provided in Appendix A.1.

Let ϕ(x,λ),ψ(x,λ) be the solutions of (Equation1) under the conditions (11) ϕ(0,λ)=1,Dxαϕ(0,λ)=h,ψ(π,λ)=1,Dxαψ(π,λ)=H.(11) Clearly, (12) U(ϕ):=Dxαϕ(0,λ)hϕ(0,λ)=0,V(ψ):=Dxαψ(π,λ)+Hψ(π,λ)=0.(12) We denote (13) Δ(λ)=Wα[ψ(x,λ),ϕ(x,λ)].(13) Using the formula for the fractional Wronskian (See [Citation51]), Wα[y,z](x)=Wα[y,z](x0), x(0,π), for two solutions y and z of (Equation1) and some x0(0,π). Hence, Wα[ψ(x,λ),ϕ(x,λ)] does not depend on x. The function Δ(λ) is called the characteristic function of Lα. Now, putting x=0 and x=π in (Equation13), we get (14) Δ(λ)=V(ϕ)=U(ψ).(14)

Theorem 3.4

The zeros {λn} of the characteristic function coincide with the eigenvalues of the boundary value problem Lα. The functions ϕ(x,λn) and ψ(x,λn) are eigenfunctions and there exists a sequence {βn} such that (15) ψ(x,λn)=βnϕ(x,λn),βn0.(15)

Proof.

The proof is very similar to the one proved in [Citation52, Theorem 1.1.1.] and thus omitted.

Lemma 3.5

Denote αn=ϕ(x,λn)2,α2. Then, we have βnαn=Δ˙(λn), where the numbers βn are defined by (Equation15), and Δ˙(λ)=ddλΔ(λ). The data {λn,αn} are called the spectral data of Lα.

Proof.

Since DxαDxαψ(x,λ)+q(x)ψ(x,λ)=λψ(x,λ),DxαDxαϕ(x,λn)+q(x)ϕ(x,λn)=λnϕ(x,λn), we get DxαWα[ψ(x,λ),ϕ(x,λn)]=(λλn)ψ(x,λ)ϕ(x,λn). Now, the α-integration over the interval [0,π] and using (Equation5) and (Equation14), we obtain (λλn)0πψ(x,λ)ϕ(x,λn)dαx=Wαψ(x,λ),ϕ(x,λn)0π=Dxαϕ(π,λn)+Hϕ(π,λn)+Dxαψ(0,λ)hψ(0,λ)=Δ(λ). For λλn, this yields 0πψ(x,λn)ϕ(x,λn)dαx=limλλnΔ(λ)λλn=Δ˙(λn). Using (Equation15) and αn concludes the proof.

Lemma 3.6

Let α(0,1]. Then the CFSL operator α is a self-adjoint on Lα2(0,π). In other words, (16) αu,v=u,αv,u,vLα2(0,π).(16)

Proof.

After applying α-integration by parts twice, we can write (17) αu,v=Wα[u,v]x=πWα[u,v]x=0+u,αv.(17) It follows from the boundary conditions (Equation2) that Wα[u,v]x=πWα[u,v]x=0=0. Therefore, α is a self-adjoint operator on Lα2(0,π).

Theorem 3.7

The eigenvalues and the eigenfunctions of the boundary value problem (Equation1), (Equation2) have the following properties:

  1. The eigenvalues are real.

  2. Eigenfunctions related to different eigenvalues are orthogonal in Lα2(0,π).

  3. All eigenvalues are simple.

Proof.

Let λn and λk(λnλk) be eigenvalues with eigenfunctions φn(x) and φk(x), respectively. Lemma 3.6 imply that αφn,φk=φn,αφk, and consequently, λnφn,φk=λkφn,φk, and as λnλk, we get φn,φk=0. Then, φn(x) and φk(x) are orthogonal on Lα2(0,π). Also, let λ1=u+iv,v0 be a complex eigenvalue with an eigenfunction φ(x,λ1)0. Since q(x),h and H are real, we conclude that λ2=λ¯1=uiv is also the eigenvalue with the eigenfunction φ¯(x,λ1). Since λ1λ2, we derive as before φ2,α2=φ,φ¯=0. Hence, φ(x,λ1)=0. Thus, all eigenvalues {λn} of Lα are real, and consequently the eigenfunctions ϕ(x,λn) and ψ(x,λn) are real as well. Since αn0, βn0, by Lemma 3.5 we have Δ˙(λn)0.

Property 3.1

Dxαcosραxα=ρsinραxα,Dxαsinραxα=ρcosραxα,DxαDxαcosραxα=ρ2cosραxα,DxαDxαsinραxα=ρ2sinραxα.

Property 3.2

Dxαcosρα(παxα)=ρsinρα(παxα),Dxαsinρα(παxα)=ρcosρα(παxα),Dtαcosρα(xαtα)=ρsinρα(xαtα),Dtαsinρα(xαtα)=ρcosρα(xαtα).

3.2. Asymptotic behaviour of the eigenvalues and eigenfunctions

In this part, similar to the classical case (e.g. see [Citation30,Citation52] for more details), we can obtain the following asymptotic formulae for the eigenvalues and eigenfunctions of Lα, and the conformable fractional derivative makes no essential complications.

Lemma 3.8

Let λ=ρ2. Then (19) ϕ(x,λ)=cosραxα+hρsinραxα+1ρ0xsinρα(xαtα)q(t)ϕ(t,λ)dαt,(19) (20) ψ(x,λ)=cosρα(παxα)+Hρsinρα(παxα)+1ρxπsinρα(tαxα)q(t)ψ(t,λ)dαt.(20)

Proof.

Equations (Equation18) and (Equation19) are called the conformable fractional Volterra integral equations. To prove (Equation18), since ϕ(x,λ) satisfies (Equation1), we see that 0xsinρα(xαtα)q(t)ϕ(t,λ)dαt=0xsinρα(xαtα)DtαDtαϕ(t,λ)dαt+ρ20xsinρα(xαtα)ϕ(t,λ)dαt. After applying α-integration by parts twice, and using (Equation11) and Property 3.2 yields 0xsinρα(xαtα)DtαDtαϕ(t,λ)dαt=hsinραxα+ρϕ(x,λ)ρcosραxαρ20xsinρα(xαtα)ϕ(t,λ)dαt, and consequently 0xsinρα(xαtα)q(t)ϕ(t,λ)dαt=hsinραxα+ρϕ(x,λ)ρcosραxα, which yields (Equation18). The formula in (Equation19) is similarly proved.

Lemma 3.9

Let λ=ρ2, ρ=σ+iτ. Then, for |ρ|, the following asymptotic formulae are valid: (20) ϕ(x,λ)=cosραxα+O1|ρ|exp|τ|αxα=Oexp|τ|αxα,Dxαϕ(x,λ)=ρsinραxα+Oexp|τ|αxα=O|ρ|exp|τ|αxα,(20) and (22) ψ(x,λ)=cosρα(παxα)+O1|ρ|exp|τ|α(παxα),=Oexp|τ|α(παxα),Dxαψ(x,λ)=ρsinρα(παxα)+Oexp|τ|α(παxα),=O|ρ|exp|τ|α(παxα).(22) All estimates are uniform with respect to x[0,π].

Proof.

Put ϕ(x,λ)=exp(|τ|αxα)f(x). Since |sin(ραxα)|exp(|τ|αxα),|cos(ραxα)|exp(|τ|αxα), we have from (Equation18) that for |ρ|1, x[0,π], f(x)=cosραxα+hρsinραxαexp|τ|αxα+1ρ0xsinρα(xαtα)exp|τ|α(xαtα)q(t)f(t)dαt. Denote μ(λ)=max0xπ|f(x)|. It follows form the last equality that f(x)1+1|ρ|h+μ(λ)0x|q(t)|dαtc1+c2|ρ|μ(λ). Hence, μ(λ)c1+c2|ρ|μ(λ). For sufficiently large |ρ|, this yields μ(λ)=O(1), i.e. ϕ(x,λ)=O(exp(|τ|αxα)). Substituting ϕ(x,λ)=cos(ραxα) into the right-hand side of (Equation18), we get ϕ(x,λ)=cosραxα+Ohρsinραxα+1ρ0xsinρα(xαtα)q(t)cosραtαdαt,=cosραxα+O1|ρ|1+0x|q(t)|dαtexp|τ|αxα. Therefore, we arrive at the first formula in (Equation20). Further, with CF differentiating (Equation18), we calculate (22) Dxαϕ(x,λ)=ρsinραxα+hcosραxα+0xcosρα(xαtα)q(t)ϕ(t,λ)dαt.(22) Thus, similarly, the second estimate (Equation20) holds. It can be proved similarly for ψ(x,λ), using (Equation19).

The most important results concerning the existence and the asymptotic behaviour of the eigenvalues and the eigenfunctions and the properties of spectral characteristics of Lα are given in the next theorems.

Theorem 3.10

The boundary value problem Lα has a countable set of eigenvalues {λn}n0. For n0, the following estimates hold: (23) ρn=λn=απα1n+ωαπn+καnn,{καn}lα2,(23) (24) ϕ(x,λn)=cosnπα1xα+ξαn(x)n,|ξαn(x)|C,(24) where (25) ωα=h+H+120πq(t)dαt.(25) Here, the same symbol {καn} denotes various sequences from lα2, and symbol C denotes various positive constants which do not depend on x,λ and n.

Proof.

By setting the asymptotic formula for ϕ(x,λ) from (Equation20) into the right-hand side of (Equation18) and (Equation22), we calculate (26) ϕ(x,λ)=cosραxα+q1(x)ρsinραxα+12ρ0xq(t)sinρα(xα2tα)dαt+O1ρ2exp|τ|αxα,(26) (27) Dxαϕ(x,λ)=ρsinραxα+q1(x)cosραxα+120xq(t)cosρα(xα2tα)dαt+O1ρexp|τ|αxα,(27) where q1(x)=h+120xq(t)dαt. In accordance with (Equation14), the relation Δ(λ)=Dxαϕ(π,λ)+Hϕ(π,λ) with the help of (Equation26) and (Equation27) is simplified to the form (28) Δ(λ)=ρsinραπα+ωαcosραπα+kα(ρ),(28) where kα(ρ)=120πq(t)cos(ρα(πα2tα))dαt+O(1ρexp(|τ|απα)).

We define Gδ={ρ:|ρ(απα1)k|δ, k=0,±1,±2,}, δ>0. Let us show that (29) sinραπαCδexp|τ|απα,ρGδ,(29) (30) |Δ(λ)|Cδ|ρ|exp|τ|απα,ρGδ,|ρ|ρ,(30) for sufficiently large ρ=ρ(δ). Let ρ=σ+iτ. It is sufficient to prove (Equation29) for the domain Dδ={ρ:σ[α2πα1,α2πα1], τ0, |ρ|δ}.

We denote θ(ρ)=|sin(ραπα)|exp(|τ|απα). Let ρDδ. When ταπα1, we have θ(ρ)Cδ, and if ταπα1, then θ(ρ)=|1exp(2iσαπα)exp(2ταπα)|/214. Thus, (Equation29) is proved. Also, combining (Equation28), (Equation29) for ρGδ gives Δ(λ)=ρsinραπα1ωαcos(ραπα)ρsin(ραπα)=ρsinραπα1+O1ρ, and consequently (Equation30) is valid. Take a circle Γn={λ:|λ|=(απα1)2(n+12)2} in the λ plan. From (Equation28), it follows that Δ(λ)=f(λ)+g(λ), f(λ)=ρsin(ραπα), |g(λ)|Cexp(|τ|απα), and using (Equation29), we obtain |f(λ)|>|g(λ)|, λΓn, for sufficiently large n(nn). Then, by Rouchè's theorem, there are as many zeros of Δ(λ) inside Γn as of the function f(λ)=ρsin(ραπα), i.e. it equals n+1. Thus, in the circle |λ|<(απα1)2(n+12)2 there exist exactly n+1 eigenvalues of Lα:λ0,,λn. Consequently, in the circle γn(δ)={ρ:|ραπα1n|δ}, there is exactly one zero of Δ(ρ2),n and ρn=λn. Since δ>0 is arbitrary, we conclude that ρn=απα1n+εn, εn=o(1) as n. By substituting this into (Equation28), we obtain 0=Δ(ρn2)=απα1n+εnsinαπα1n+εnπαα+ωαcosαπα1n+εnπαα+καn. This implies (31) απα1nsinεnαπα+ωαcosεnαπα+καn=0.(31) Then, sin(εnαπα)=O(1n), i.e. εn=O(1n). Using (Equation31), once more we more precisely obtain εn=ωαπn+καnn, i.e. (Equation23) is valid. Setting (Equation23) into (Equation26), we arrive at (Equation24), where (26) ξαn(x)=πα1αh+120xq(t)dαtxααωαπ+καnsinnπα1xα+πα12α0xq(t)sinnπα1(xα2tα)dαt+O1n.(26) In conclusion, |ξαn(x)|C, and this completes the proof.

Theorem 3.11

The specification of the spectrum {λn}n0 uniquely determines the characteristic function Δ(λ) by the formula (33) Δ(λ)=π3α2α3(λ0λ)n=1λnλn2.(33)

Proof.

It follows from (Equation28) that Δ(λ) is entire in λ of order 1/2, hence by Hadamard's factorization theorem, we have Δ(λ)=Cn=0(1λλn). Consider the function Δ~(λ)=ρsin(ραπα)=λπααn=1(1λ(α2/π2α2)n2). Then Δ~(λ)Δ(λ)=Cλλ0λλ0α3π3α2n=1n2λnn=1(1+λn(α2/π2α2)n2(α2/π2α2)n2λ). Now, combining (Equation23) and (Equation28) yields limλΔ~(λ)/Δ(λ)=1,limλn=1(1+λn(α2/π2α2)n2(α2/π2α2)n2λ)=1, and hence C=π3α2α3λ0n=1λnn2. Combined with Δ(λ), this implies (Equation33).

3.3. Completeness and expansion theorems

In this subsection, we provide that the system of the eigenfunctions of the CFSL boundary value problem Lα is complete and forms an orthogonal basis in Lα2(0,π). The completeness and expansion theorems are important for solving various problems in mathematical physics by the Fourier method, and also for the spectral theory itself.

Theorem 3.12

  1. The system of eigenfunctions {ϕ(x,λn)}n0 of the boundary value problem Lα is complete in Lα2(0,π).

  2. Let f(x), x[0,π] be an absolutely continuous function. Then (34) f(x)=n=0cnϕ(x,λn),cn=1αn0πf(t)ϕ(t,λn)dαt,(34) and the series converges uniformly on [0,π].

  3. Let f(x)Lα2(0,π), the series (Equation34) converges in Lα2(0,π), and 0π|f(x)|2dαx=n=0αn|cn|2.

Proof.

The proof of this theorem is similar to the proof of Theorem 1.2.1. from [Citation52, p.15] and will be given in Appendix A.2.

4. Regularized trace formulas

In this section, we obtain the regularized sum from the eigenvalues for problem Lα. We follow the idea employed in the proof of the similar trace formula for the Sturm-Liouville operators in [Citation26,Citation31]. We first present a lemma that gives more precise asymptotic formulae for characteristic function Δ(λ) and other formulas (See the proof in Appendix A.3).

Lemma 4.1

If q(x)Wα2(0,π), then one can obtain more precise asymptotic formulae as before: (36) ϕ(x,λ)=cosραxα+q1(x)ρsinραxα+q20(x)ρ2cosραxα14ρ20xDtαq(t)cosρα(xα2tα)dαt+O1ρ3exp|τ|αxα,(36) (37) Dxαϕ(x,λ)=ρsinραxα+q1(x)cosραxα+q21(x)ρsinραxα+14ρ0xDtαq(t)sinρα(xα2tα)dαt+O1ρ2exp|τ|αxα,(37) (38) Δ(λ)=ρsinραπα+ωαcosραπα+ω0ρsinραxα+kα0(ρ)ρ,(38) where (38) kα0(ρ)=140πDtαq(t)sinρα(πα2tα)dαt+O1ρexp|τ|απα,q1(x)=h+120xq(t)dαt,ω0=q21(π)+Hq1(π),q2j(x)=14q(x)+(1)j+1q(0)+(1)j+120xq(t)q1(t)dαt,j=0,1.(38)

Subsequent estimates are based on the following lemma.

Lemma 4.2

If |(α2π2α2)n2λn|b, then (39) n=1|(α2π2α2)n2λn|k(μ2+(α2π2α2)n2)kπα2αbkμ2k1.(39)

Proof.

Let λ0<λ1<λ2<<λn< be the eigenvalues of the problem (Equation1)–(Equation2). Applying Theorem 3.10, we thus obtain (40) λn=ρn2=α2π2α2n2+2αωαπα+O1n2,(40) this yields (α2π2α2)n2λn=O(1), i.e. there exists b>0 such that |(α2π2α2)n2λn|b, and consequently n=1(α2π2α2)n2λnkμ2+(α2π2α2)n2kbkn=11μ2+(α2π2α2)n2kbk0dtμ2+(α2π2α2)t2kbππα1αμ2k10dz1+z2kπα2αbkμ2k1, and the lemma is thus proved.

Theorem 4.3

Suppose that q(x)Wα2(0,π), and let {λn}n=0 be the sequence of the eigenvalues of the problem Lα(q(x),h,H). Then, the following trace formula is valid: (45) sλ:=λ0+n=1λnα2π2α2n22αωαπα,=14q(π)+q(0)απαh+H+120πq(t)dαt12h+H2.(45)

Proof.

Let λ=μ2, here μ is real, we obtain the sum of (Equation41) with comparing the coefficients of the same powers of μ on the formulas (Equation33) and (Equation37) for μ+. We conclude from Theorem 3.11 that (42) Δ(μ2)=π3α2α3(μ2+λ0)n=1λn+μ2n2=(μ2+λ0)sinh(μαπα)μΦ(μ),(42) where Φ(μ)=n=1λn+μ2μ2+(α2π2α2)n2=n=1(1(α2π2α2)n2λnμ2+(α2π2α2)n2). To study the asymptotic behaviour of Φ(μ) for large positive μ, we consider (43) lnΦ(μ)=n=1ln1(α2π2α2)n2λnμ2+(α2π2α2)n2=k=11kn=1(α2π2α2)n2λnμ2+(α2π2α2)n2k.(43) The estimate (Equation39) leads to (44) k=21kn=1(α2π2α2)n2λnkμ2+(α2π2α2)n2kπα2αb2μ3k=0bμ2k=O1μ3=o1μ2.(44) Furthermore, (49) n=1(α2π2α2)n2λnμ2+(α2π2α2)n2=2αωαπαn=11μ2+(α2π2α2)n2+1μ2n=1λnα2π2α2n22απαωα1μ2n=1λnα2π2α2n22απαωα(α2π2α2)n2μ2+(α2π2α2)n2.(49) Since supn|λn(α2π2α2)n2+2απαωα|(α2π2α2)n2<, it follows from (Equation39) that (50) 1μ2n=1λnα2π2α2n22απαωα(α2π2α2)n2μ2+α2π2α2n2=1μ2n=1O(1)μ2+(α2π2α2)n2πα2αμ3=O1μ3=o1μ2.(50) Also, we know that (47) n=11μ2+(α2π2α2)n2=πα12α(παα)μcoth(πααμ)1μ2(πα1α)=πα2α1μ12μ2+o1μ2.(47) Now, putting (Equation44)–(Equation47) and (Equation41) into (Equation43), we get lnΦ(μ)=ωαμ+1μ2sλλ0+αωαπα+o1μ2. Therefore (53) Φ(μ)=expωαμ+1μ2sλλ0+αωαπα+o1μ2,=1+ωαμ+1μ2sλλ0+αωαπα+ωα22+o1μ2.(53) Combining (Equation48) with (Equation42), and using sinhμ=(eμeμ)/2 as μ+ yields (49) Δ(μ2)=exp(μπαα)2μ+ωα+1μsλ+αωαπα+ωα22+o1μ.(49) Substituting ρ=μ2=±iμ into (Equation37) as μ+ gives us (55) Δ(μ2)=μsinhμαπα+ωαcoshμαπα+ω0μsinhμαπα+o1μ,=exp(μπαα)2μ+ωα+ω0μ+o1μ.(55) Moreover, comparing the coefficients (Equation49) with (Equation50) allows us to conclude that sλ=ω0απαωα12ωα2, and using (Equation25) and (Equation38), we can write sλ=14q(π)+q(0)12h+H2+120πq(t)q1(t)dαt+Hh+120πq(t)dαtαπαh+H+120πq(t)dαt12h+H0πq(t)dαt180πq(t)dαt2. Finally, since 0xq(t)0tq(s)dαsdαt=12(0xq(t)dαt)2 holds, we have proved (Equation41).

5. Inverse nodal problems

The main results of this section are the asymptotic formula for elements constructed by of nodal points, oscillation of the eigenfunctions, uniqueness theorem and reconstruction formula for any qLα1(0,π) as a limit of nodal data.

5.1. Oscillation theorem. Asymptotic formula

Let {λn}n0 be the set of eigenvalues of (Equation1)–(Equation2) and ϕ(x,λn) be the eigenfunction corresponding to the eigenvalue λn. Using an analogue of the Sturm oscillation theorem, for sufficiently large n, we find that ϕ(x,λn) has exactly n nodal points locating in (0,π). Define Y={ynj:ynj=(xnj)α/α, xnjX},n>0, j=1,n¯ and the set X={xnj:nN} is called the nodal set of the boundary value problem Lα. In other words, ϕ(xnj,λn)=0. Also, let Inj=[ynj, ynj+1] be the j-th domain of the n-th element of Y and let lnj=ynj+1ynj be the length of the j-th domain, which is constructed by the n-th nodes. We also define the function jn(y) to be the largest index j such that 0ynjy. Thus, j=jn(y) if and only if y[ynj,ynj+1).

The following lemma is, in fact, a refinement of [Citation53, Lemma 2.2]. Its proof is similar and will be given in Appendix A.4.

Lemma 5.1

Assume that qLα1(0,π). Then, as n, (9) ynj=j12πρn+1ρn2h+120ynj1+cos(2ρnt)q(αt)1αt1αdαt+o1ρn3,(9) (30) lnj=πρn+12ρn2ynjynj+11+cos(2ρnt)q(αt)1αt1αdαt+o1ρn3,(30) for the problem (Equation1)–(Equation2), where ynj=(xnj)α/α, xnjX, for j=1,n¯.

Theorem 5.2

For sufficiently large n, the eigenfunction ϕ(x,λn) of the problem Lα(q(x),h,H) has exactly n nodes in the interval (0,π), and 0<xn1<xn2<<xnn<π, for n1. Moreover, suppose that q(x)Lα1(0,π). Then, the elements Y constructed by of nodal points have the following asymptotic form, as n (60) ynj=γnj+πα1αn2πα1αhγnjωαπ+καn0γnj+πα12α0γnj1+cos2αnπα1tq(αt)1αt1αdαt+o1n3,(60) uniformly with respect to j, where γnj:=παα(j1/2n) for j=1,n¯.

Proof.

Let λn=ρn2 be the eigenvalues of the problem Lα. From (Equation23), we obtain 1ρn=πα1αnπ2α2α2n3ωαπ+καn+o1n3,1ρn2=π2α2α2n2+O1n4. Substituting back into (Equation51), and using the following relation 0ynjcos2αnπα1tcos(2ρnt)q(αt)1αdt=o1n,n, we arrive at (Equation53), which, in turn, gives ynj+1ynj=πααn+O(1n2), n. Observing that for large n, ynj<ynj+1 for positive n. Also, for j=1,n¯ the formula (Equation53) gives yn1=πα2αn+O1n2,,ynn=πααπα2αn+O1n2. Thus, we conclude that 0<yn1<yn2<<ynn<παα, for n>0. Since 0<α1, then α11, we can easily obtain 0<xn1<xn2<<xnn<π, for n>0. Hence, precisely n zeros of the eigenfunction ϕ(x,λn) lie in (0,π), and this completes the proof.

Corollary 5.3

From Theorem 5.2 it follows that the set Y is dense in [0,πα/α].

5.2. Uniqueness theorem. A solution of the inverse nodal problem

Now, we are ready to formulate two explicit algorithms for the potential function and the parameters h, H in the boundary conditions in the CFSLP (Equation1)–(Equation2) by using elements constructed by of nodes and the length of those elements. We point out that our results are extensions to those in [Citation38,Citation44,Citation54–58]. We consider the following inverse problem.

Problem 5.1

Given a set Y, constructed by of nodal points X, or a subset Y0={ynkj} of a set Y, where {ynkj} is dense in (0,πα/α), find the parameters h,H in the boundary conditions and the function q(x)q, where q=απα0πq(t)dαt.

Theorem 5.4

For each x[0,π]. Let {ynkj}Y be chosen such that limnynj=y. Then, the following finite limit exists and the corresponding equality holds, for j=1,n¯. (54) limnαn2πα1ynjγnj=deff(x),(54) where (55) f(x)=πα1αh+120xq(t)dαtxααωαπ.(55)

Proof.

Thanks to the Theorem 5.2 and the fact that limnynj=y implies that limnγnj=y, we have (69) limnαn2πα1ynjγnj=limnαn2πα1πα1αn2πα1αhγnjωαπ+καn0γnj0γnj+πα12α0γnj1+cos2ntπα1q(αt)1αt1αdαt+o1n3,=πα1αh+120yq(αt)1αt1αdαtyωαπ=g(y).(69) Consequently, (Equation55) follows from (EquationA15) and (Equation56).

Based on Theorem 5.4 we can now prove a uniqueness theorem and provide an effective procedure for the solution of the inverse nodal problem. We agree that if a particular symbol ϒ denotes an object related to Lα=Lα(q(x),h,H), then Υ~ will denote an analogous object related to L~α=Lα(q~(x),h~,H~).

Theorem 5.5

Let Y0Y be dense set on (0,πα/α). Let Y0=Y~0 then q(x)q=q~(x)q~ a.e. on (0,π), h=h~, H=H~. Thus, the specification of Y uniquely determines the potential q(x)q on (0,π) and the coefficients of the boundary conditions. Furthermore, the function q(x)q and the numbers h, H can be constructed by the formulae (70) q(x)q=2α2π2α2f(0)f(π)π+2απα1Dxαf(x),(70) (71) h=απα1f(0),H=απα1f(π),(71) where f(x) is calculated by (Equation55).

Proof.

Taking α-CF derivative of (Equation55), we calculate Dxαf(x)=πα12αq(x)ωαπ, and using (Equation25), we arrive at (Equation57)–(Equation58). If Y0=Y~0, then (Equation54) implies that f(x)=f~(x) for x[0,π]. By virtue of (Equation57)–(Equation58), we get q(x)q=q~(x)q~ a.e. on (0,π), h=h~, H=H~.

Corollary 5.6

Let X0={xnkj} and X0X be a subset of nodal points which satisfy that {(xnkj)α/α} is dense in (0,πα/α). Then, the X0 uniquely determines the parameters h, H in the boundary conditions and the function q(x)q. Furthermore, the numbers h, H and the function q(x)q can be found by the following algorithm:

Algorithm 5.1

Let a subset X0 of the nodal points be given. Then

  1. Create a dense subset Y0={ynj} in (0,πα/α) by ynj=(xnj)α/α, xnjX0.

  2. For each y(0,πα/α), choose a sequence {ynj}Y0 such that limnynj=y.

  3. Find the function f(x) by (Equation54), and calculate Dxαf(x),f(0),f(π).

  4. Finally, calculate the function q(x)q by the formula (Equation57), and determine h,H by (Equation58).

Let us point out an alternative algorithm using the notion of the length of the elements constructed by of nodal points lnj:=ynj+1ynj, which allows one to approximate q(x) directly by a pointwise limit. First, we need the following assertion.

Theorem 5.7

Let f:[x,y]R be a continuous function. Then, there exists ξ in (x,y), such that (59) xyf(t)dαt=f(ξ)yαxαα.(59)

Proof.

Similar to the mean value theorem for integrals, we can easily obtain (Equation59).

Lemma 5.8

Let qLα1[0,π]. Then limnρnπxnjxnj+1q(u)dαu=limnρnπynjynj+1q(αt)1αt1αdαt=q(x),limnρnπxnjxnj+1cos2ρnαuαq(u)dαu=limnρnπynjynj+1cos(2ρnt)q(αt)1αt1αdαt=0, for almost every x(0,π), with j=jn(x) for j=1,n¯.

Proof.

The proof follows from the same arguments as in the proof of Lemma 3.1 and Theorem 3.2 in [Citation53].

Theorem 5.9

The potential function qLα1(0,π) satisfies (60) q(x)=limn2ρn2ρnlnjπ1,(60) for almost every x(0,π), with j=jn(y), for j=1,n¯.

Proof.

From (Equation52), it is clear that (88) 2ρn2ρnlnjπ1=ρnπynjynj+1cos(2ρnt))q(αt)1αt1αdαt+ρnπynjynj+1q(αt)1αt1αdαt+o(1).(88) Passing to the limit as n in (Equation61) and Lemma 5.8, we arrive at (Equation60).

Lemma 5.10

For the set Y={ynj}, constructed by of nodal points X, the following limits exist for all j=1,n¯:

  1. limnαnπα1[(αnπα1)ynj(j12)π],

  2. limnαnπα1[(αnπα1)ynnj(nj12)π].

Corollary 5.11

Let j be fixed. If (i) exists, then (62) h=limnαnπα1αnπα1ynjj12π,(62) and if (ii) exists, then (63) H=limnαnπα1αnπα1ynnjnj12π.(63)

Proof.

Combining Lemma 5.10 with (Equation53) concludes the proof.

Remark 5.1

In practice, if the third term (i.e. γnj(ωαπ+καn)) appears in the asymptotic formula of ynj as defined in Theorem 5.2 for j=1,n¯, then the limit (Equation60) with the approximation ρn(απα1)n takes the following form: (64) limn2α2n2π2α2αnπαlnj1=q(x)q2απαh+H.(64) If the third term does not appear in (Equation53), then limn2α2n2π2α2[αnπαlnj1]=q(x).

Thus, we arrive at the following alternative algorithm for solving Problem 5.1, which, unlike the first one contains no differentiation.

Algorithm 5.2

Let a subset X0 of the nodal points be given. Then

  1. Generate a dense subset Y0={ynj} in (0,πα/α) by ynj=(xnj)α/α, xnjX0.

  2. Create a sequence {lnj} from the length of the elements Y0 by lnj=ynj+1ynj.

  3. Find the function q(x) by the formula (Equation64).

  4. Calculate the parameters h,H in the boundary conditions by (Equation62) and (Equation63).

Example 5.1

Let the potential function q(x)=cos(2xαπα1) be given. Then, the eigenvalues and the eigenfunctions for the boundary value problem (Equation1)–(Equation2) with h=H=1 obtained from the relations (Equation40) and (EquationA14), respectively, are formulated in the following forms, for nn0, sufficiently large n0, (32) λn=α2π2α2n2+4απα+O1n2,(32) (33) ϕ(x,λ)=cosnxαπα1+πα1αn1+πα14αsin2xαπα1sinnxαπα1π3α34α3n21+πα18αsin2xαπα1sin2xαπα1cosnxαπα1+π2α28α2n1n+1cos2(n+1)xαπα1+1n1cos2(n1)xαπα12nn21cosnxαπα1+π2α28α2n1n+1sin2(n+1)xαπα1+1n1sin2(n1)xαπα1×sinnxαπα1+o1n2.(33) The growth of eigenvalues of the CFSLP, λn, is plotted in Figure , corresponding to six values of α=0.1,0.2,0.5 (a) and α=0.6,0.8 (b). As we see, the growth of the eigenvalues in CFSLP w.r.t. n is dependent on the fractional derivative order α, as shown in (Equation65). Since α(0,1], there are several growth modes, the sublinear growth corresponding to α0, the subquadratic growth mode which corresponds to α1, and in other points of this interval, where a superlinear-subquadratic growth is observed. The case α=1 leads to an exact quadratic growth. To visually get more sense of how the eigensolutions look like, in Figure  we plot the eigenfunctions of this example, ϕ(x,λ) of different orders and corresponding to different values of α used in Figure . We realize that, for α values very close to 0, e.g. α=0.1, most of the roots of the eigenfunctions are accumulated on a around of zero and have minimal values and so close to 0. Also, for α values sufficiently away from 0, especially as α approaches 1, e.g. α=0.8, the roots (the nodal points) are distributed in the interval [0,π]. We can see that the conditions of oscillation theorem 5.2 are provided.

Figure 1. Growth of the eigenvalues of CFSLP in Example 5.1, versus n, corresponding to different fractional-order α=0.1; (a): sublinear growth, α=0.2 and 0.5; (a): superlinear-subquadratic growth, and α=0.6; (b): subquadratic growth, α=0.8; (b): quadratic growth. Here we compare the growth of the eigenvalues to the classical case, i.e. α=1.0 with quadratic growth. The blue line denotes the linear growth. (a) The first ten eigenvalues for α=0.1,0.2,0.5. (b) The first ten eigenvalues for α=0.6,0.8,1.0.

Figure 1. Growth of the eigenvalues of CFSLP in Example 5.1, versus n, corresponding to different fractional-order α=0.1; (a): sublinear growth, α=0.2 and 0.5; (a): superlinear-subquadratic growth, and α=0.6; (b): subquadratic growth, α=0.8; (b): quadratic growth. Here we compare the growth of the eigenvalues to the classical case, i.e. α=1.0 with quadratic growth. The blue line denotes the linear growth. (a) The first ten eigenvalues for α=0.1,0.2,0.5. (b) The first ten eigenvalues for α=0.6,0.8,1.0.

Figure 2. Eigenfunctions of CFSLP in Example 5.1, versus x, corresponding to the fractional order α=0.1 (first row), α=0.2 (second row), α=0.5 (third row), and α=0.8 (last row), for n=1,3,5 (left column), n=10 (right column).

Figure 2. Eigenfunctions of CFSLP in Example 5.1, versus x, corresponding to the fractional order α=0.1 (first row), α=0.2 (second row), α=0.5 (third row), and α=0.8 (last row), for n=1,3,5 (left column), n=10 (right column).

The nodal points of the Equation (Equation1)–(Equation2) with h=H=1 obtained from the relation (Equation51), are given as (67) xnj=αynj1α, n>1,j=1,n¯,(67) where ynj=πααj1/2n+πα1αn2πα1α+π2α2n24α2(n21)sin2π(j1/2)n+o1n3. We calculate in Table  the numerical results of true nodal points and their predictions by the asymptotic formula (Equation67). The true nodal points are calculated by Newton's method, and the values are precise in the first eight digits (which is set to 1.0×1012 in our computation). We observe that, as α values away from 0, the asymptotic formula (Equation67) is a good approximation not only for sufficiently large n but also for small spectral numbers, especially for α values close to 1. This is in stark coincidence with the classical case α=1, where the asymptotic formula is very accurate even for relatively small n, (See Figure  and Table ). Also, we obtain the absolute errors, enj, between the true (Λnj, the roots of Equation (Equation66) obtained by Newton's method) and explicit nodal points xnj, for n=2,5 and n=10, j=1,n¯, corresponding to the fractional-order α=0.1,0.5,0.8, and α=1.0 which are seen in Table .

Table 1. Numerical values of nodal points and absolute errors between the true and predicted values, for n=2 (Part I), n=5 (Part II), n=10 (Part III), corresponding to the fractional order α=0.1,0.5,0.8 and α=1.0 in Example 5.1.

Example 5.2

Let {ynj}Y0 be the dense subset of nodal points in (0,πα/α) given by the following asmptotics: (86) ynj=πααj1/2n+πα1αn2πα1αj1/2nωα+πα12απ2α2α(π2α2+4)×expπα(j1/2n)2πα1sin2π(j1/2)n+cos2π(j1/2)n1+o1n3,j=1,n¯,(86) where ωα=1+π2α22α(π2α2+4)(exp(πα)1). From (Equation68), we deduce (87) lnj=πααn+πα1αn2πα1αnωα+πα12απ2α2α(π2α2+4)expπα(j+1/2n)×2πα1sin2π(j+1/2)n+cos2π(j+1/2)nexpπα(j1/2n)×2πα1sin2π(j1/2)n+cos2π(j1/2)n+o1n3,j=1,n¯.(87) Therefore, using (Equation54) and (Equation55), and Algorithm 5.1, we get f(x)=ωαπxαα+π3α32α2(π2α2+4)×expxα2πα1sin2xαπα1+cos2xαπα11. By virtue of (Equation57)–(Equation58), we compute q(x)=exp(xα)cos(2xαπα1), h=0, H=1. Another method for calculating the potential function q(x) is using Algorithm 5.2. Theorem 5.9 with (Equation69) imply limn2α2n2π2α2[αnπαlnj1]=exp(xα)cos(2xαπα1)q2απα, which concludes q(x)=exp(xα)cos(2xαπα1), and using (Equation68) and Corollary 5.11, we obtain h=0, H=1.

6. Concluding remarks

We expanded the spectral theory for CFSL operator α, and proved that the eigenvalues are all real, simple and the eigenfunctions corresponding to distinct eigenvalues are orthogonal in the Lα2(0,π)-space. Also, we presented asymptotic formulae for the eigenvalues and eigenfunctions, and then, we showed that every nontrivial solution u of (Equation1) and its CF-derivative Dxαu in [0,π] are entire functions of λ of order 1/2, and consequently by Hadamard's factorization theorem we obtain the infinite product representation for the characteristic function Δ(λ) of Lα. Next, we indicated that the obtained eigenfunctions are complete in different Hilbert space, and provided sufficient conditions under which the Fourier series for the eigenfunctions converges uniformly on [0,π]. Moreover, we have presented a first analytical study of the theory of regularized traces and inverse nodal problems for CFSL operator α. The experimental and numerical observations lead to some interesting conjectures for the CFSLP:

  1. Thanks to the asymptotic formula (Equation40) of the eigenvalues, there are several modes of growth referred to as a sublinear mode corresponding to α0, a quadratic mode which corresponds to α1, and a superlinear-subquadratic mode corresponding to other values α of (0,1).

  2. The presented nodal points xn1,xn2,,xnn set up the increasing sequence. On the other hand, one can observe that the nodal points decrease with decreasing the order of derivative α. Also, for sufficiently large n, numerical example confirm the validity of the oscillation theorem for this type of the FSLP. Therefore, only the location of the zeros of eigenfunctions depends on the conformable derivative order, i.e. when the fractional-order α sufficiently away from 0 and approaches 1, similar to the classic case, the distribution of the zeros of the eigenfunctions in the interval (a,b) is uniform.

  3. For each α(0,1], any dense subset of the elements constructed by of nodal points, and the length of those elements completely determines the potential q and the coefficients of the boundary conditions in the CFSLP (Equation1)–(Equation2). Also, by increasing the order of derivative α, the asymptotic formulas obtained for the nodal points are in good agreement with the exact ones.

The authors believe that the new conceptions of the theory of CFSLPs not only can be applied to solve other problems with CF derivatives such as CF partial differential equations, CF integro-differential equations, CF partial integro-differential equations and CF optimal control problems and so on but can be easily used to solve the mentioned problems with integer order derivatives. Apart from these theoretical aspects, the rigorous analysis of appropriate numerical schemes is also of immense interest. In particular, the error estimates for the eigenvalue approximations, the rate of convergence for numerical schemes, numerical solution of inverse nodal problems, etc.

Disclosure statement

No potential conflict of interest was reported by the authors.

References

  • Khalil R, Al Horani M, Yousef A, et al. A new definition of fractional derivative. J Comput Appl Math. 2014;264:65–70. doi: 10.1016/j.cam.2014.01.002
  • Abdeljawad T. On conformable fractional calculus. J Comput Appl Math. 2015;279:57–66. doi: 10.1016/j.cam.2014.10.016
  • Atangana A, Baleanu D, Alsaedi A. New properties of conformable derivative. Open Math. 2015;13:889–898. doi: 10.1515/math-2015-0081
  • Ortigueira MD, Tenreiro Machado JA. What is a fractional derivative?. J Comput Phys. 2015;293:4–13. doi: 10.1016/j.jcp.2014.07.019
  • Makhlouf AB, Naifar O, Hammami MA, et al. FTS and FTB of conformable fractional order linear systems. Math Probab Eng. 2018;2018: 2572986:5.
  • Chung WS. Fractional Newton mechanics with conformable fractional derivative. J Comput Appl Math. 2015;290:150–158. doi: 10.1016/j.cam.2015.04.049
  • Ebaid A, Masaedeh B, El-Zahar E. A new fractional model for the falling body problem. Chin Phys Lett. 2017;34:020201–1. doi: 10.1088/0256-307X/34/2/020201
  • Anderson DR, Ulness DJ. Properties of the Katugampola fractional derivative with potential application in quantum mechanics. J Math Phys. 2015;56:063502. doi: 10.1063/1.4922018
  • Lazo MJ, Torres DFM. Variational calculus with conformable fractional derivatives. IEEE/CAA J Automat Sinica. 2017;4:340–352. doi: 10.1109/JAS.2016.7510160
  • Nwaeze ER, Torres DFM. Chain rules and inequalities for the BHT fractional calculus on arbitrary timescales. Arab J Math. 2017;6:13–20. doi: 10.1007/s40065-016-0160-2
  • Benkhettou N, Hassani S, Torres DFM. A conformable fractional calculus on arbitrary time scales. J King Saud Univ Sci. 2016;28:93–98. doi: 10.1016/j.jksus.2015.05.003
  • Zhou HW, Yang S, Zhang SQ. Conformable derivative approach to anomalous diffusion. Phys A. 2018;491:1001–1013. doi: 10.1016/j.physa.2017.09.101
  • Iyiola O, Tasbozan O, Kurt A, et al. On the analytical solutions of the system of conformable time-fractional Robertson equations with 1-D diffusion. Chaos Solitons Fractals. 2017;94:1–7. doi: 10.1016/j.chaos.2016.11.003
  • Avci D, Iskender Eroglu B, Ozdemir N. The Dirichlet problem of a conformable advection-diffusion equation. Thermal Sci. 2017;21:9–18. doi: 10.2298/TSCI160421235A
  • Cenesiz Y, Kurt A, Nane E. Stochastic solutions of conformable fractional Cauchy problems. Statist Probab Lett. 2017;124:126–131. doi: 10.1016/j.spl.2017.01.012
  • Zhao D, Luo M. General conformable fractional derivative and its physical interpretation. Calcolo. 2017;54:903–917. doi: 10.1007/s10092-017-0213-8
  • Al-Towailb MA. A q-fractional approach to the regular Sturm-Liouville problems. Electron J Differ Equ. 2017;88:1–13.
  • Khosravian-Arab H, Dehghan M, Eslahchi MR. Fractional Sturm-Liouville boundary value problems in unbounded domains, theory and applications. J Comput Phys. 2015;299:526–560. doi: 10.1016/j.jcp.2015.06.030
  • Klimek M, Agrawaln OP. Fractional Sturm-Liouville problem. Comput Math Appl. 2013;66:795–812. doi: 10.1016/j.camwa.2012.12.011
  • Rivero M, Trujillo JJ, Velasco MP. A fractional approach to the Sturm-Liouville problem. Centr Eur J Phys. 2013;11:1246–1254.
  • Monje CA, Chen Y, Vinagre BM, et al. Fractional-order systems and controls: fundamentals and applications. London (UK): Springer-Verlag; 2010.
  • Baleanu D, Guvenc ZB, Machado JT. New trends in nanotechnology and fractional calculus applications. New York (US): Springer; 2010.
  • Mainardi F. Fractional calculus and waves in linear viscoelasticity: an introduction to mathematical models. London, UK: Imperial College Press; 2010.
  • Silva MF, Machado JAT. Fractional order PD α joint control of legged robots. J Vib Control. 2006;12:1483–1501. doi: 10.1177/1077546306070608
  • Pálfalvi A. Efficient solution of a vibration equation involving fractional derivatives. Int J Nonlin Mech. 2010;45:169–175. doi: 10.1016/j.ijnonlinmec.2009.10.006
  • Levitan BM. Calculation of the regularized trace for the Sturm-Liouville operator. Uspekhi Mat Nauk. 1964;19:161–165.
  • Gelfand IM, Levitan BM. On a simple identity for the characteristic values of a differential operator of the second order. Dokl Akad Nauk SSSR. 1953;88:593–596.
  • Dikii LA. On a formula of Gelfand-Levitan. Uspekhi Mat Nauk. 1953;8:119–123.
  • Halberg CJ, Kramer VA. A generalization of the trace concept. Duke Math J. 1960;27:607–617. doi: 10.1215/S0012-7094-60-02758-7
  • Levitan BM, Sargsjan IS. Sturm-Liouville and Dirac operators. Vol. 59. Dordrecht: Kluwer Academic Publishers Group; 1991.
  • Levitan BM, Sargsjan IS. Introduction to spectral theory: selfadjoint ordinary differential operators. Vol. 39. Providence: American Mathematical Society; 1975.
  • Sadovnichii VA, Podol'skii VE. Traces of differential operators. Differ Equ. 2009;45:477–493. doi: 10.1134/S0012266109040028
  • Borisov D, Freitas P. Eigenvalue asymptotics, inverse problems and a trace formula for the linear damped wave equation. J Differential Equations. 2009;247:3028–3039. doi: 10.1016/j.jde.2009.07.029
  • Hıra F. The regularized trace of Sturm-Liouville problem with discontinuities at two points. Inverse Probl Sci Eng. 2017;25:785–794. doi: 10.1080/17415977.2016.1197921
  • Carlson R. Eigenvalue estimates and trace formulas for the matrix Hill's equation. J Differential Equations. 2000;167:211–244. doi: 10.1006/jdeq.2000.3785
  • Yang CF. Trace formulae for the matrix Schrödinger equation with energy-dependent potential. J Math Anal Appl. 2012;393:526–533. doi: 10.1016/j.jmaa.2012.03.003
  • Gelfand IM, Levitan BM. On the determination of a differential equation from its spectral function. Amer Math Soc Trans. 1951;1:253–304.
  • McLaughlin JR. Inverse spectral theory using nodal points as data–a uniqueness result. J Differential Equations. 1988;73:354–362. doi: 10.1016/0022-0396(88)90111-8
  • Hald OH, McLaughlin JR. Solution of inverse nodal problems. Inverse Problems. 1989;5:307–347. doi: 10.1088/0266-5611/5/3/008
  • Browne PJ, Sleeman BD. Inverse nodal problem for Sturm-Liouville equation with eigenparameter depend boundary conditions. Inverse Problems. 1996;12:377–381. doi: 10.1088/0266-5611/12/4/002
  • Yang XF. A solution of the inverse nodal problem. Inverse Problems. 1997;13:203–213. doi: 10.1088/0266-5611/13/1/016
  • Cheng YH, Law CK, Tsay J. Remarks on a new inverse nodal problem. J Math Anal Appl. 2000;248:145–155. doi: 10.1006/jmaa.2000.6878
  • Ozkan AS, Keskin B. Inverse nodal problems for Sturm-Liouville equation with eigenparameter-dependent boundary and jump conditions. Inverse Probl Sci Eng. 2015;23:1306–1312. doi: 10.1080/17415977.2014.991730
  • Shieh CT, Yurko VA. Inverse nodal and inverse spectral problems for discontinuous boundary value problems. J Math Anal Appl. 2008;347:266–272. doi: 10.1016/j.jmaa.2008.05.097
  • Yang CF, Yang XP. Inverse nodal problems for the Sturm-Liouville equation with polynomially dependent on the eigenparameter. Inverse Probl Sci Eng. 2011;19:951–961. doi: 10.1080/17415977.2011.565874
  • Yang CF. Inverse nodal problems of discontinuous Sturm-Liouville operator. J Differential Equations. 2013;254:1992–2014. doi: 10.1016/j.jde.2012.11.018
  • Wang YP, Yurko VA. On the inverse nodal problems for discontinuous Sturm-Liouville operators. J Differential Equations. 2016;260:4086–4109. doi: 10.1016/j.jde.2015.11.004
  • Wang YP, Lien KY, Shieh CT. Inverse problems for the boundary value problem with the interior nodal subsets. Appl Anal. 2017;96:1229–1239. doi: 10.1080/00036811.2016.1183770
  • Wang Y, Zhou J, Li Y. Fractional sobolev's spaces on time scales via conformable fractional calculus and their application to a fractional differential equation on time scales. Adv Math Phys. 2016;2016:9636491:21.
  • Pospíšil M, Škripková LP. Sturm's theorems for conformable fractional differential equations. Math Commun. 2016;21:273–281.
  • Hammad MA, Khalil R. Abel's formula and wronskian for conformable fractional differential equations. Int J Differ Equ Appl. 2014;13:177–183.
  • Freiling G, Yurko VA. Inverse Sturm-Liouville problems and their applications. New York: Nova Science Publishers; 2001.
  • Law CK, Shen CL, Yang CF. The inverse nodal problem on the smoothness of the potential function. Inverse Problems. 1999;15:253–263. doi: 10.1088/0266-5611/15/1/024
  • Yang XF. A new inverse nodal problem. J Differential Equations. 2001;169:633–653. doi: 10.1006/jdeq.2000.3911
  • Guo Y, Wei G. Inverse problems: dense nodal subset on an interior subinterval. J Differential Equations. 2013;255:2002–2017. doi: 10.1016/j.jde.2013.06.006
  • Law CK, Yang CF. Reconstructing the potential function and its derivatives using nodal data. Inverse Problems. 1998;14:299–312. doi: 10.1088/0266-5611/14/2/006
  • Shen CL, Tsai TM. On a uniform approximation of the density function of a string equation using EVs and nodal points and some related inverse nodal problems. Inverse Problems. 1995;11:1113–1123. doi: 10.1088/0266-5611/11/5/014
  • Yang CF. Trace and inverse problem of a discontinuous Sturm-Liouville operator with retarded argument. J Math Anal Appl. 2012;395:30–41. doi: 10.1016/j.jmaa.2012.04.078

Appendix

A.1. Proof of Theorem 3.3

The functions ϕ1(x,λ)=cos(ραxα) and ϕ2(x,λ)=ρ1sin(ραxα) are the solutions of DxαDxαy+ρ2y=0 with zero potential and Wα[ϕ1(,λ),ϕ1(,λ)]=1, where ρ:=λ is defined with respect to the principal branch. For all x[0,π], λC, we define the sequence {yn(,λ)}n=1 of successive approximations by (A1) yn+1(x,λ)=y1(x,λ)+0x{ϕ2(x,λ)ϕ1(t,λ)ϕ1(x,λ)ϕ2(t,λ)}q(t)yn(t,λ)dαt,(A1) where y1(x,λ)=c1ϕ1(x,λ)+c2ϕ2(x,λ). We prove that for each fixed λC the uniform limit of yn(,λ) as n exists and defines a solution of (Equation1) and (Equation10). Let λC be fixed. There exist positive numbers K(λ) and K~(λ) such that |ϕi(x,λ)|K(λ)/2 for i=1,2 and |y1(x,λ)|K~(λ), where K~(λ):=(|c1|+|c2|)K(λ)/2, for all x[0,π]. Let us show by induction that (A2) |yn+1(x,λ)yn(x,λ)|K~(λ)K(λ)Q0(x)nn!,x[0,π],nN,(A2) where Q0(x)=0x|q(t)| dαt. Indeed, for n=1, (EquationA2) is obvious. Suppose that (EquationA2) is valid for a certain fixed n1. From (EquationA1), we get (A3) |yn+2(x,λ)yn+1(x,λ)|K(λ)0x|q(t)||yn+1(t,λ)yn(t,λ)|dαt.(A3) Substituting (EquationA2) into the right-hand side of (EquationA3), we calculate |yn+2(x,λ)yn+1(x,λ)|K~(λ)(K(λ))n+1n!0x|q(t)|(Q0(t))ndαt=K~(λ)(K(λ))n+1n!0x(Q0(t))nDtαQ0(t)dαt=K~(λ)K(λ)Q0(x)n+1(n+1)!. It follows from [Citation49, Theorem 26] and q(x)Lα2(0,π) that q1,αα1παq2,α<. Consequently by Weierstrass M-test the series (A4) y(x,λ)=y1(x,λ)+n=1yn+1(x,λ)yn(x,λ).(A4) Since the n-th partial sums of the series is nothing but yn+1(,λ), then yn+1(,λ) approaches a function ϕ(,λ) uniformly on [0,π] as n, where ϕ(,λ) is the sum of the series. Similarly, we can also prove by induction on n that Dxαyn+1(,λ) approaches a function Dxαϕ(,λ) uniformly on [0,π] as n. Hence, both ϕ(,λ) and Dxαϕ(,λ) are continuous, i.e. ϕ(,λ)Cα2[0,π]. Because of the uniform convergence, letting n in (EquationA1), we obtain ϕ(x,λ)=c1ϕ1(x,λ)+c2ϕ2(x,λ)+0x{ϕ2(x,λ)ϕ1(t,λ)ϕ1(x,λ)ϕ2(t,λ)}q(t)ϕ(t,λ)dαt. Then, by α-Leibniz rule 2.14, clearly, ϕ(,λ) satisfies (Equation1) and (Equation10). That ϕ(,λ) is an entire function of the variable λ follows from the uniform convergence of the series (EquationA4) and the structure of the functions yn(,λ). To prove that problem (Equation1), (Equation10) has a unique solution, suppose on the contrary that ψi(,λ), i=1,2, are two solutions of (Equation1), (Equation10). Therefore, for x[0,π], we can write ψ1(x,λ)ψ2(x,λ)0xK(λ)|q(t)ψ1(t,λ)ψ2(t,λ)dαt, letting ψ(x,λ)=|ψ1(x,λ)ψ2(x,λ)|, we have ψ(x,λ)0xK(λ)|q(t)ψ(t,λ)dαt. Thus, the fractional version of Gronwall's inequality (See Lemma 2.15) implies ψ(x,λ)0, and consequently ψ1(x,λ)=ψ2(x,λ) for all x[0,π]. This completes a proof of the theorem.

A.2. Proof of Theorem 3.12

We denote Gα(x,t,λ)=1Δ(λ)ϕ(x,λ)ψ(t,λ),xtϕ(t,λ)ψ(x,λ),xt, and consider the function Y(x,λ)=0πGα(x,t,λ)f(t)dαt,=1Δ(λ)ψ(x,λ)0xϕ(t,λ)f(t)dαt+ϕ(x,λ)xπψ(t,λ)f(t)dαt. The function Gα(x,t,λ) is called Green's function for Lα. Gα(x,t,λ) is the kernel of the inverse operator for the CFSL operator, i.e. Y(x,λ) is the solution of the boundary value problem αYλY+f(x)=0,U(Y)=V(Y)=0. From (Equation15), Theorem 3.7 and using Lemma 3.5, we obtain (A5) Resλ=λnY(x,λ)=1αnϕ(x,λn)0πf(t)ϕ(t,λn)dαt.(A5) Let f(x)Lα2(0,π) be such that 0πf(t)ϕ(t,λn)dαt=0,n0, then (EquationA5) implies that Resλ=λnY(x,λ)=0, and consequently for each x[0,π], the function Y(x,λ) is entire in λ. Moreover, it follows from Lemma 3.9 and (Equation30) that for a fixed δ>0, and sufficiently large ρ>0, |Y(x,λ)|Cδ|ρ|, ρGδ, |ρ|ρ. Using the maximum principle and Liouville's theorem, we conclude that Y(x,λ)0. Hence,  f(x)=0 a.e. on (0,π). Thus (1) is proved. Let now f(x)AC[0,π] be an arbitrary absolutely continuous function. Applying the same arguments as in the proof of [Citation52, Theorem 1.2.1.], we obtain (A6) Y(x,λ)=f(x)λ1λZ1(x,λ)+Z2(x,λ),(A6) where Z1(x,λ)=1Δ(λ)ψ(x,λ)0xg(t)Dtαϕ(t,λ)dαt+ϕ(x,λ)xπg(t)Dtαψ(t,λ)dαt,Z2(x,λ)=1Δ(λ)hf(0)ψ(x,λ)+Hf(π)ϕ(x,λ)+ψ(x,λ)0xq(t)ϕ(t,λ)f(t)dαt+ϕ(x,λ)xπq(t)ψ(t,λ)f(t)dαt,g(t):=Dtαf(t). Using Lemma 3.9 and (Equation30), we get for a fixed δ>0, and sufficiently large ρ>0, (A7) max0xπ|Z2(x,λ)|C|ρ|,ρGδ,|ρ|ρ.(A7) Let now g(t)Lα1(0,π). Fix ε>0 and choose an absolutely continuous function gε(t) such that 0π|g(t)gε(t)|dαt<ε2C+ (See [Citation49, Definition 16]), where C+=max0xπsupρGδ1|Δ(λ)||ψ(x,λ)|0x|Dtαϕ(t,λ)|dαt+|ϕ(x,λ)|xπ|Dtαψ(t,λ)|dαt. Then, we have |Z1(x,λ)||Z1(x,λ,gε)|+|Z1(x,λ,ggε)|ε/2+C(ε)/|ρ|, for ρGδ, |ρ|ρ. Hence (A8) lim|ρ|ρGδmax0xπ|Z1(x,λ)|=0.(A8) Consider the contour integral IN(x)=12πiΓNY(x,λ)dλ,ΓN=λ:|λ|=απα12N+122. It follows from (EquationA6)–(EquationA8) that IN(x)=f(x)+εN(x), limNmax0xπ|εN(x)|=0. On the other hand, we can calculate IN(x) with the help of the residue theorem. From (EquationA5), we get IN(x)=n=0Ncnϕ(x,λn),cn=1αnf,ϕ(t,λn). If N, we arrive at (Equation34), where the series converges uniformly on [0,π], i.e. (2) is proved. Moreover, by virtue of (1), the Parseval's equality is clear.

A.3. Proof of Lemma 4.1

Let q(x)Wα2(0,π). Then, α-integration by parts yields (A9) 120xq(t)cosρα(xα2tα)dαt=14ρsinραxα(q(x)+q(0))+14ρ0xDtαq(t)sinρα(xα2tα)dαt,(A9) (A10) 12ρ0xq(t)sinρα(xα2tα)dαt=14ρ2cosραxα(q(x)q(0))14ρ20xDtαq(t)cosρα(xα2tα)dαt.(A10) It follows from (Equation26), (EquationA10) that ϕ(x,λ)=cosραxα+q1(x)ρsinραxα+O1ρ2exp|τ|αxα. By putting this asymptotic into the right-hand side of (Equation18) and from Property 3.2 and (Equation6), we conclude that (A11) ϕ(x,λ)=cosραxα+q1(x)ρsinραxα+12ρ0xq(t)sinρα(xα2tα)dαt+I+II,(A11) where (A12) I=12ρ2cosραxα0xq(t)q1(t)dαt+12ρ20xcosρα(xα2tα)q(t)q1(t)dαt,=12ρ2cosραxα0xq(t)q1(t)dαt+14ρ3sinραxαq(x)q1(x)q(0)q1(0)+14ρ30xsinρα(xα2tα)Dtαq(t)q1(t)dαt,=12ρ2cosραxα0xq(t)q1(t)dαt+O1ρ3exp|τ|αxα.(A12) Since q(x)Lα2(0,π), then (A13) II=O1ρ30xsinρα(xαtα)exp|τ|αtαq(t)dαt=O1ρ3exp|τ|αxα.(A13) Hence, by substituting (EquationA12), (EquationA13) into (EquationA11), and applying (EquationA10) we arrive at (Equation35). The formula in (Equation36) is proved similarly. Finally, by using (Equation35), (Equation36) and (Equation14), the desired result for Δ(λ) can be concluded, i.e. (Equation37) are valid.

A.4. Proof of Lemma 5.1

For arbitrary λ>0, let ρ=λ. By virtue of Lemma 3.8, the function ϕ(x,λ) with the initial conditions ϕ(0,λ)=1 and Dxαϕ(0,λ)=h is a unique solution of the CF integral equation (Equation18). Setting ϕ(t,λ) in Lemma 3.8 into the right-hand side of (Equation18), we have ϕ(x,λ)=cosραxα+hρsinραxα+1ρ0xsinρα(xαtα)cosραtαq(t)dαt+1ρ20x0tsinρα(xαtα)sinρα(tαyα)q(y)q(t)ϕ(y,λ)dαydαt+hρ20xsinρα(xαtα)sinραtαq(t)dαt. Thus, by considering ϕ(x,λ)=cos(ραxα)+O(ρ1) and trigonometric calculations, and using (EquationA9) and (EquationA10), we can write (A14) ϕ(x,λ)=cosραxα+sin(ραxα)ρh+120x1+cos2ραtαq(t)dαtcos(ραxα)2ρ0xsin2ραtαq(t)dαthcos(ραxα)2ρ20xq(t)dαtcos(ραxα)4ρ20x0tq(y)q(t)dαydαt+o1ρ2.(A14) If ϕ(x,λ) equal to zero and sin(ραxα) is not close to zero, then by using (EquationA14), we get 0=cotραxα+1ρh+120x1+cos2ραtαq(t)dαtcot(ραxα)2ρ0xsin2ραtαq(t)dαthcot(ραxα)2ρ20xq(t)dαtcot(ραxα)4ρ20x0tq(y)q(t)dαydαt+o1ρ2. Since cot(ρx)=O(ρ1), the above equality yields cotραxα1+o1ρ=1ρh+120x1+cos2ραtαq(t)dαt+o1ρ2, and obtain cot(ραxα)=1ρ(h+120x(1+cos(2ραtα))q(t)dαt)+o(1ρ2). Set y=α1xα, by the change of variable t=α1uα, dαu=t1αdαt, we obtain (A15) 0x1+cos2ραuαq(u)dαu=0y1+cos(2ρt)q(αt)1αt1αdαt.(A15) Therefore cot(ρy)=1ρ(h+120y(1+cos(2ρt))q((αt)1α)t1αdαt)+o(1ρ2). Now, we take ρ=ρn and y=ynj. Since the Taylor's expansion for the arccotangent function is given by Arccoty=(j1/2)πk=0(1)2ky2k+1/(2k+1), for some integer j, we then have ρnynj=(j12)π+1ρn(h+120ynj(1+cos(2ρnt))q((αt)1α)t1αdαt)+o(1ρn2), which yields (Equation51). Finally, the asymptotic formula for lnj follows from (Equation51) for j=1,2,,n.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.