Research article

Spectral parameter power series method for Kurzweil–Henstock integrable functions

  • Received: 18 May 2024 Revised: 25 July 2024 Accepted: 30 July 2024 Published: 07 August 2024
  • MSC : 26A39, 34B24, 34L16, 41A58, 65L10

  • In this paper, the convergence of the spectral parameter power series method, proposed by Kravchenko, is performed for the Sturm–Liouville equation with Kurzweil–Henstock integrable coefficients. Numerical simulations of some examples are also presented to validate the performance of the method.

    Citation: Israel A. Cordero-Martínez, Salvador Sánchez-Perales, Francisco J. Mendoza-Torres. Spectral parameter power series method for Kurzweil–Henstock integrable functions[J]. AIMS Mathematics, 2024, 9(9): 23598-23616. doi: 10.3934/math.20241147

    Related Papers:

    [1] Juan H. Arredondo, Genaro Montaño, Francisco J. Mendoza . A new characterization of the dual space of the HK-integrable functions. AIMS Mathematics, 2024, 9(4): 8250-8261. doi: 10.3934/math.2024401
    [2] Mukhamed Aleroev, Hedi Aleroeva, Temirkhan Aleroev . Proof of the completeness of the system of eigenfunctions for one boundary-value problem for the fractional differential equation. AIMS Mathematics, 2019, 4(3): 714-720. doi: 10.3934/math.2019.3.714
    [3] Imran Talib, Md. Nur Alam, Dumitru Baleanu, Danish Zaidi, Ammarah Marriyam . A new integral operational matrix with applications to multi-order fractional differential equations. AIMS Mathematics, 2021, 6(8): 8742-8771. doi: 10.3934/math.2021508
    [4] Erdal Bas, Ramazan Ozarslan . Theory of discrete fractional Sturm–Liouville equations and visual results. AIMS Mathematics, 2019, 4(3): 593-612. doi: 10.3934/math.2019.3.593
    [5] Ahmed M.A. El-Sayed, Eman M.A. Hamdallah, Hameda M. A. Alama . Multiple solutions of a Sturm-Liouville boundary value problem of nonlinear differential inclusion with nonlocal integral conditions. AIMS Mathematics, 2022, 7(6): 11150-11164. doi: 10.3934/math.2022624
    [6] Jiaye Lin . Evaluations of some Euler-type series via powers of the arcsin function. AIMS Mathematics, 2025, 10(4): 8116-8130. doi: 10.3934/math.2025372
    [7] Tuba Gulsen, Sertac Goktas, Thabet Abdeljawad, Yusuf Gurefe . Sturm-Liouville problem in multiplicative fractional calculus. AIMS Mathematics, 2024, 9(8): 22794-22812. doi: 10.3934/math.20241109
    [8] Anwar Ahmad, Dumitru Baleanu . On two backward problems with Dzherbashian-Nersesian operator. AIMS Mathematics, 2023, 8(1): 887-904. doi: 10.3934/math.2023043
    [9] Juya Cui, Ben Gao . Symmetry analysis of an acid-mediated cancer invasion model. AIMS Mathematics, 2022, 7(9): 16949-16961. doi: 10.3934/math.2022930
    [10] M. Mossa Al-Sawalha, Khalil Hadi Hakami, Mohammad Alqudah, Qasem M. Tawhari, Hussain Gissy . Novel Laplace-integrated least square methods for solving the fractional nonlinear damped Burgers' equation. AIMS Mathematics, 2025, 10(3): 7099-7126. doi: 10.3934/math.2025324
  • In this paper, the convergence of the spectral parameter power series method, proposed by Kravchenko, is performed for the Sturm–Liouville equation with Kurzweil–Henstock integrable coefficients. Numerical simulations of some examples are also presented to validate the performance of the method.



    The spectral parameter power series (SPPS) method, introduced in [1], expresses the solution of the Sturm–Liouville equation

    (ρy)+qy=λy

    as a power series with respect to the spectral parameter λ, where the coefficients are given in terms of a particular solution of the homogeneous equation

    (ρy)+qy=0.

    In [2], the SPPS method was used for solving spectral problems for Sturm–Liouville equations. This method is an important and efficient tool for solving a variety of problems involving Sturm–Liouville equations. In most publications devoted to the SPPS method, the coefficients of the differential equations are assumed to be continuous. In [3], it is shown that the SPPS method is valid for Sturm–Liouville equations with coefficients in the Lebesgue space L([a,b]).

    There are problems that are described by Sturm–Liouville differential equations with highly oscillatory coefficients, and the Lebesgue integral is not enough to integrate some coefficients of this type. The Kurzweil–Henstock integral is more general than the Lebesgue integral, and it is well known that it integrates highly oscillating functions, so in this work we study the SPPS method for solving spectral problems for Sturm–Liouville equations with coefficients in the space of Kurzweil–Henstock integrable functions.

    The Kurzweil–Henstock integral was discovered independently by J. Kurzweil in the context of differential equations and R. Henstock, who made a systematic study. It is an integral whose definition is as simple as the Riemann integral but is more general than the Lebesgue integral. It has good convergence criteria and the advantage that it is not necessary to introduce improper integrals. A particular difference with the Lebesgue integral is that it is not an absolute integral.

    We denote by KH([a,b]) the space of Kurzweil–Henstock integral functions. The Alexiewicz seminorm for this space is given by

    f[a,b]=sup{|dcf|:[c,d][a,b]}.

    The space KH([a,b]) with the seminorm [a,b] is not complete, contrary to what happens with the Lebesgue space L([a,b]) with its norm f1=ba|f|. The space of all functions of bounded variation on [a,b] is denoted by BV([a,b]), and the variation of a function φBV([a,b]) is denoted by V[a,b]φ. It is well known that the multipliers of KH integrable functions are functions of bounded variation, i.e., if fKH([a,b]) and gBV([a,b]), then

    fgKH([a,b]). (2.1)

    Moreover, even though the integral is not absolute, we have an estimate for the integral of this product:

    |bafg||baf|inf[a,b]|g|+f[a,b]V[a,b]g. (2.2)

    This is a Holder-type inequality for the Kurzweil–Henstock integral. See [4, Lemma 24].

    A useful result, given in [5, Theorem 7.5], which we will use later, says that if fL([a,b]), then F, defined by F(x)=xaf, is of bounded variation on [a,b], and

    V[a,b]F=ba|f|. (2.3)

    The space of absolutely continuous functions (respectively, generalized absolutely continuous functions in the restricted sense) on [a,b] is denoted by AC([a,b]), respectively, by ACG([a,b]). See [6]. In these spaces, a fundamental theorem of calculus is stated in its general form.

    Theorem 2.1. [6, Fundamental Theorem of Calculus] Let f,F:[a,b]C be functions, and let x0[a,b].

    a) If fKH([a,b]) (resp. fL([a,b])) and F(x)=xx0f for all x[a,b], then FACG([a,b]), (resp. FAC([a,b])) and F=f almost everywhere on [a,b]. In particular, if f is continuous at x[a,b], then F(x)=f(x).

    b) FACG([a,b]) (resp. FAC([a,b])) if and only if F exists almost everywhere on [a,b] and xx0F=F(x)F(x0) for all x[a,b]. In the case of absolutely continuous functions, the Lebesgue integral is used.

    In the study of differential equations, Sobolev spaces become relevant. The classical Sobolev space is defined by

    W1,1([a,b])={uL([a,b]):gL([a,b]) such that bauφ=bagφ,φC1c(a,b)}, (2.4)

    where C1c(a,b) is the space of continuously differentiable functions on (a,b) with compact support in (a,b). In [7], the authors introduce a generalization of this space using the KH-integral; this space is known as the KH-Sobolev space, and it is defined by

    WKH([a,b])={uKH([a,b]):gKH([a,b]) such that bauφ=bagφ,φV}, (2.5)

    where V is a suitable test function space (for additional details, see [7]). The function g given in (2.4) and (2.5) is known as the weak derivative of the function u and is denoted by ˙u. It is well known (see [7, Theorem 3.6]) that if u=v a.e on [a,b] and vAC([a,b]) (resp. vACG([a,b])), then uW1,1([a,b]) (resp. uWKH([a,b])). Furthermore, the weak derivative coincides pointwise with the classical derivative whenever it exists, i.e.,

    ˙u(x)=v(x), (2.6)

    for all x in which v(x) exists. For example, if u is defined by u(x)=xcos(πx) for all x(0,1] and u(0)=0, then uACG([0,1]). Thus, uWKH([0,1])W1,1([0,1]), and ˙u(x)=cos(πx)+πxsin(πx) for all x(0,1], and ˙u(0)=0.

    Blancarte et al. [3] study the SPPS method for functions that are Lebesgue integrable. In this section, the convergence of the SPPS method is performed for the Sturm–Liouville differential equation with Kurzweil–Henstock integrable coefficients. First, let us introduce some concepts that will be necessary to deal with this generalization.

    Definition 3.1. Let f,g:[a,b]C be functions, and t[a,b]. If there exists Ut an open neighborhood in [a,b] of t[a,b], where

    a) f is continuous, then we say that f is locally continuous at t;

    b) f is of class C1, then we say that f is locally of class C1 at t;

    c) f(x)=g(x) for all xUt, then we say that f is locally equal to g at t.

    The set of points x[a,b], where 1ρ is locally continuous at x, is denoted by Eρ. While the space of functions that are locally of class C1 at every point of Eρ is denoted by C1p.

    Consider the Sturm–Liouville (S-L) equation

    (ρ˙y)˙+qy=λrya.e. on [a,b]. (3.1)

    Through this paper, we assume that q and r are functions such that q,rKH([a,b]) and ρ is a function such that 1ρL([a,b]). Also, we assume that Eρ. Define the space

    A={yAC([a,b])C1ρ:ρ˙y=ga.e. on [a,b]for somegACG([a,b])}. (3.2)

    Sánchez–Perales et al. in [8] use this space to guarantee the existence of the solutions of the Sturm–Liouville type differential equations with KH-integrable coefficients. It is clear that if yA, then y exists a.e. on [a,b], especially at every point of Eρ and ˙y=y. Also, since there exists gACG([a,b]) such that ρ˙y=g a.e. on [a,b], it follows that (ρ˙y)˙=gKH([a,b]).

    In Example 4.2, ρ is defined by ρ(x)=π+x. Note that 1ρL([π,0]) and Eρ=(π,0]. An example of an element of A is the function y defined by y(x)=4(x+π)3. Indeed, y is absolutely continuous on [π,0] and y is locally of class C1 at every point of (π,0], so yAC([π,0])C1ρ; moreover, ρ(x)˙y(x)=34x+π for all x(π,0], and this function is ACG on [π,0].

    By the form of the S-L Eq (3.1), we consider the differential operator L:AKH([a,b]) defined as L[y]=(ρ˙y)˙+qy. The next lemma lets us rewrite the operator L in terms of a non-vanishing solution of the associated homogeneous equation of the S-L equation.

    Lemma 3.2 (Polya factorization). Let y0A be a solution of the homogeneous equation L[y]=0 a.e. on [a,b], with y0(x)0 for all x[a,b]. Then,

    L[y]=1y0[ρy20(yy0)]a.e. on [a,b].

    Proof. Let yA. Then, y,y0,ρ˙y,ρ˙y0 are derivable in the weak sense. Since y0 is continuous and does not vanish on [a,b], there exists α>0 such that α|y0(x)| for all x[a,b]. Therefore, yy0 also has weak derivative and

    (yy0)(x)=˙y(x)y0(x)y(x)˙y0(x)y20(x),for all x[a,b].

    Moreover, as L[y0]=0 a.e. on [a,b], there exists a set B[a,b] with measure zero such that for every x[a,b]B, L[y0](x)=0. Let x[a,b]B, then

    1y0(x)[ρ(x)y20(x)(yy0)(x)]=1y0(x)[ρ(x)y20(x)(˙y(x)y0(x)y(x)˙y0(x)y20(x))]=1y0(x)[(ρ˙y)(x)y0(x)(ρ˙y0)(x)y(x)]=1y0(x)[(ρ˙y)˙(x)y0(x)(ρ˙y0)˙(x)y(x)]=1y0(x)[((ρ˙y)˙(x)+q(x)y(x))y0(x)((ρ˙y0)˙(x)+q(x)y0(x))y(x)]=1y0(x)[L[y](x)y0(x)L[y0](x)y(x)]=1y0(x)[L[y](x)y0(x)0y(x)]=L[y](x).

    Now, we will define a family of functions that allow us to write the representation of the general solution of the S-L Eq (3.1) as a spectral power series. Let x0[a,b], and y0A be such that y0(x)0 for all x[a,b]. Define

    ˜X(0)1,˜X(n)(x)={xx0˜X(n1)(s)r(s)y20(s)ds, if n es odd;xx0˜X(n1)(s)dsρ(s)y20(s), if n es even; (3.3)
    X(0)1,X(n)(x)={xx0X(n1)(s)dsρ(s)y20(s), if n is odd;xx0X(n1)(s)r(s)y20(s)ds, if n is even. (3.4)

    These functions are bound by the coefficients of the series of an exponential, as illustrated by the following proposition:

    Proposition 3.3. Let x0[a,b] and y0A be such that y0(x)0 for all x[a,b]. Then, for each nN{0},

    ˜X(2n+1)ACG([a,b]) and ˜X(2n)AC([a,b]), (3.5)
    X(2n+1)AC([a,b]) and X(2n)ACG([a,b]). (3.6)

    Furthermore, the following inequalities are satisfied:

    |˜X(2n)(x)|ry20n[a,b](sgn(xx0)Q(x))nn!,|˜X(2n+1)(x)|ry20n+1[a,b](sgn(xx0)Q(x))nn!, (3.7)
    |X(2n)(x)|ry20n[a,b](sgn(xx0)Q(x))nn!,&|X(2n+1)(x)|ry20n[a,b](sgn(xx0)Q(x))n+1(n+1)!, (3.8)

    for all x[a,b], where Q(x)=xx0ds|ρ(s)y20(s)|.

    Proof. We will only show (3.5) and (3.7). Since y0A, it follows that y0AC([a,b])BV([a,b]). Then by (2.1), y20rKH([a,b]), and hence, by Theorem 2.1, ˜X(1)=()x0y20(s)r(s)dsACG([a,b]). Now, suppose that ˜X(2n+1)ACG([a,b]). By hypothesis 1ρL([a,b]), then ˜X(2n+1)ρy20L([a,b]), which implies by Theorem 2.1 that ˜X(2n+2)=()x0˜X(2n+1)(s)ρ(s)y20(s)dsAC([a,b]). Thus, y20˜X(2n+2)BV([a,b]) and so by (2.1), ˜X(2n+2)y20rKH([a,b]). Hence, by Theorem 2.1, ˜X(2n+2+1)=()x0˜X(2n+2)(s)y20(s)r(s)dsACG([a,b]). By induction (3.5) holds.

    To prove (3.7), first note that for each x[a,x0) and nN,

    |˜X(2n+1)(x)|ry20[a,b]x0x|˜X(2n1)(s)ρ(s)y20(s)|ds. (3.9)

    Indeed, by the inequality (2.2),

    |˜X(2n+1)(x)|=|xx0˜X(2n)(s)r(s)y20(s)ds|=|x0x˜X(2n)(s)r(s)y20(s)ds||x0xr(s)y20(s)ds|inft[x,x0]|˜X(2n)(t)|+ry20[x,x0]V[x,x0]˜X(2n)=ry20[x,x0]V[x,x0]˜X(2n). (3.10)

    Since ˜X(2n)(x)=x0x˜X(2n1)(s)ρ(s)y20(s) ds, it follows from the equality (2.3) that

    V[x,x0]˜X(2n)=x0x|˜X(2n1)(s)ρ(s)y20(s)|ds. (3.11)

    Substituting (3.11) into Eq (3.10), we obtain that the inequality (3.9) holds.

    Now, we show again, by induction, that

    |˜X(2n)(x)|ry20n[a,b](Q(x))nn!,|˜X(2n+1)(x)|ry20n+1[a,b](Q(x))nn!, x<x0. (3.12)

    For n=0, we have that for every x[a,x0),

    |˜X(1)(x)|=|xx0r(s)y20(s)ds|ry20[a,b]. (3.13)

    By the induction hypothesis, we assume that the second inequality given in (3.12) is valid for the natural n. Let x<x0 and observe that

    |˜X(2n+2)(x)|=|x0x˜X(2n+1)(s)ρ(s)y20(s)|dsx0x|˜X(2n+1)(s)ρ(s)y20(s)|dsx0xry20n+1[a,b]n!(Q(s))n|ρ(s)y20(s)|ds=(1)nry20n+1[a,b]x0xQn(s)n!1|ρ(s)y20(s)|ds. (3.14)

    On the other hand, [Qn+1(s)n!(n+1)]=(n+1)Qn(s)Q(s)n!(n+1)=Qn(s)n!1|ρ(s)y20(s)| for almost all s[x,x0]. Therefore,

    x0xQn(s)n!1|ρ(s)y20(s)|ds=Qn+1(x)(n+1)!.

    Consequently,

    |˜X(2(n+1))(x)|x0x|˜X(2n+1)(s)ρ(s)y20(s)|ds(1)n+1ry20n+1[a,b]Qn+1(x)(n+1)!. (3.15)

    This proves the first inequality in (3.12) for the natural n+1. The second inequality in (3.12) for the natural n+1 is obtained by (3.9) and (3.15):

    |˜X(2(n+1)+1)(x)|ry20[a,b]x0x|˜X(2n+1)(s)ρ(s)y20(s)|ds(1)n+1ry20n+1+1[a,b]Qn+1(x)(n+1)!. (3.16)

    This completes the induction process. Estimates for the case where x>x0 and for the function X(n), with nN, are shown similarly.

    The next result shows us that we can bound the functions ˜X(n) and X(n) at any point by constants.

    Corollary 3.4. Under the conditions of Proposition 3.3, the functions ˜X(2n),˜X(2n+1),X(2n), and X(2n+1) are bounded by the following constants:

    |˜X(2n)(x)|Cn1Cn2n!,|˜X(2n+1)(x)|Cn+11Cn2n!,|X(2n)(x)|Cn1Cn2n!,&|X(2n+1)(x)|Cn1Cn+12(n+1)!, (3.17)

    where C1=ry20[a,b] and C2=1ρy201.

    Proof. Without loss of generality, we assume that ax<x0. Then,

    0<Q(x)=xx0ds|ρ(s)y20(s)|=x0xds|ρ(s)y20(s)|bads|ρ(s)y20(s)|=1ρy201=C2.

    Thus,

    0<[Q(x)]nCn2 (3.18)

    for all nN{0}. By Proposition 3.3 and the inequalities in (3.18), we have that

    |˜X(2n)(x)|ry20n[a,b](Q(x))nn!Cn1Cn2n!,|˜X(2n+1)(x)|ry20n+1[a,b](Q(x))nn!Cn+11Cn2n!,|X(2n)(x)|ry20n[a,b](Q(x))nn!Cn1Cn2n!,&|X(2n+1)(x)|ry20n[a,b](Q(x))n+1n!Cn1Cn+12n!.

    In the following result, we define several functions (related to the representation of the solution of the S-L Eq (3.1) as a power spectral series), the space to which these functions belong, and some aspects related to their derivative.

    Theorem 3.5. Let y0A be such that y0(x)0 for all x[a,b]. If ˜X(n) and X(n) are the functions given in (3.3)(3.4), and u,v,w,z:[a,b]C are defined as

    u=n=0λn˜X(2n),v=n=0λn+1˜X(2n+1),w=n=0λnX(2n+1), z=n=0λnX(2n). (3.19)

    Then,

    a) u,wAC([a,b]), and

    ˙u=vρy20&˙w=zρy20a.e. on [a,b]. (3.20)

    Moreover, the equalities

    ˙u=k=0λk[˜X(2k)]=vρy20and˙w=k=0λk[X(2k+1)]=zρy20, (3.21)

    are locally satisfied at every point of Eρ, and so, u,wC1ρ.

    b) v,zACG([a,b]), and

    ˙v=n=0λ(n+1)[˜X(2n+1)]=λry20uand˙z=n=0λn[X(2n)]=λry20w (3.22)

    a.e. on [a,b].

    Proof. (a) First, by Corollary 3.4, we have that n=0|λn˜X(2n)|n=0(|λ|C1C2)nn!<. Note that the right-hand side of this inequality is the expansion in power series of the function exp(|λ|C1C2). Then, by the Weiersstras M-test, we can actually see that the series n=0λn˜X(2n)uu. Similarly, this occurs with the functions v, w, and z.

    The next part of this proof relies on [3, Proposition 3]. We have a sequence (λn˜X(2n)) in AC([a,b]) such that n=0λn˜X(2n) converges uniformly to u on [a,b]. Now, we show that n=0λn˜X(2n) converges to the function vρy20 in L([a,b]) with the norm 1. By Theorem 2.1 and the definition of X(2n), we have that (˜X(2n))=˜X(2n1)ρy20 a.e. on [a,b]. Then,

    Nn=0(λn˜X(2n))vρy201=ba|Nn=1λn˜X(2n1)ρy20vρy20|=ba|1ρy20||Nn=1λn˜X(2n1)v|.sup[a,b]|N1n=0λn+1˜X(2n+1)v|C20,

    as N. Thus, by [3, Proposition 3], there exists a function FAC([a,b]) such that n=0λn˜X(2n) converges uniformly to F and ˙F=vρy20 a.e. on [a,b]. But as mentioned earlier, n=0λn˜X(2n) converges uniformly to u, so by uniqueness of the limits, u=F, which means that uAC([a,b]) and ˙u=vρy20 a.e. on [a,b]. This proves one of the equalities in Eq (3.20).

    Now, let xEρ, then the functions 1ρ,˜X(2n1)ρy20 are locally continuous at x, which means that there exist δx>0 such that 1ρ,˜X(2n1)ρy20 are continuous on [xδx,x+δx][a,b]. Then, by Theorem 2.1,

    [˜X(2n)](s)=˜X(2n1)(s)ρ(s)y20(s)for all s[xδx,x+δx][a,b]. (3.23)

    Let Um=mn=0λn˜X(2n). Then, Umu over [xδx,x+δx][a,b], and ˙Um(s)=mn=1λn˜X(2n1)(s)ρ(s)y20(s) for all s[xδx,x+δx][a,b]. Observe that

    |˙Um(s)v(s)ρ(s)y20(s)|=|mn=1λn˜X(2n1)(s)ρ(s)y20(s)v(s)ρ(s)y20(s)|=|1ρ(s)y20(s)||mn=1λn˜X(2n1)(s)v(s)|maxs[xδx,x+δx][a,b]|1ρ(s)y20(s)||m1n=0λ(n+1)˜X(2n+1)(s)v(s)|u0 (3.24)

    as m. Thus, ˙Um converges uniformly to vρy20 on [xδx,x+δx][a,b], i.e.,

    n=0λn[˜X(2n)](s)=v(s)ρ(s)y20(s) (3.25)

    for all s[xδx,x+δx][a,b]. Consequently, by [9, Theorem 7.17] and (3.25),

    ˙u(s)=n=0λn[˜X(2n)](s)=v(s)ρ(s)y20(s) (3.26)

    for all s[xδx,x+δx][a,b]. From (3.26) and being arbitrary xEρ, we obtain that uC1ρ.

    (b) Define the functions V and Z by

    V(x)=xx0λry20uandZ(x)=xx0λry20w.

    Then, V,ZACG([a,b]), v=V and z=Z+1. We will only show that V=v. Let x[a,x0). Note that UmBV([x,x0]), ˙Um=mn=1λn˜X(2n1)ρy20 a.e. on [a,b], and ˙UmL([a,b]). Consequently, by (2.3) and from Corollary 3.4, we have that

    V[x,x0]Umx0x|˙Um(s)|dsx0xmn=1|λn||˜X(2n1)(s)ρ(s)y20(s)|ds=mn=1|λ|nx0x|˜X(2(n1)+1)(s)ρ(s)y20(s)|dsmn=1|λ|nx0xCn1Cn12(n1)!|ρ(s)y20(s)|dsmn=1|λ|nCn1Cn2(n1)!ds<. (3.27)

    Therefore, (Um) is a uniformly bounded variation on [x,x0]. Moreover, Umuu on [x,x0] and ry0KH([x,x0]). Consequently, by [10, Corollary 3.2], it follows that

    x0xr(s)y20(s)mn=0λn˜X(2n)(s)dsx0xr(s)y20(s)u(s)ds, (3.28)

    when m. Therefore,

    |v(x)V(x)|=|limmmn=0λ(n+1)˜X(2n+1)(x)xx0λr(s)y20(s)u(s)ds|=|λ||limmmn=0λnxx0˜X(2n)(s)r(s)y20(s)dsxx0r(s)y20(s)u(s)ds|=|λ||limmx0xr(s)y20(s)mn=0λn˜X(2n)(s)dsx0xr(s)y20(s)u(s)|0. (3.29)

    The same occurs for xx0. Thus, v=V on [a,b], and so vACG([a,b]). By Theorem 2.1 (a),

    ˙v=λry20u (3.30)

    a.e. on [a,b]. Now, let Vm=λ(m+1)˜X(2m+1). Then, ˙Vm=λ(m+1)˜X(2m)ry20 a.e. on [a,b]. For each mN, take Em a set of measure zero, such that for every x[a,b]Em, ˙Vm(x)=λ(m+1)˜X(2m)(x)r(x)y20(x). Define E=mNEm, then m(E)=0, and for every x[a,b]E,

    |mn=0˙Vn(x)λr(x)y20(x)u(x)|=|mn=0λ(n+1)˜X(2n)(x)r(x)y20(x)λr(x)y20(x)u(x)|=|λ||r(x)y20(x)||mn=0λn˜X(2n)(x)u(x)|0, (3.31)

    when m. Thus,

    λry20u=n=0λ(n+1)[˜X(2n+1)] (3.32)

    a.e. on [a,b]. From Eqs (3.30) and (3.32), we conclude Eq (3.22). For the function z, the proof follows similarly.

    Theorem 3.6. Let x0Eρ. If y0A is a solution of the homogeneous equation

    (ρ˙y)˙+qy=0a.e. on [a,b] (3.33)

    with y0(x)0 for all x[a,b]. Then, the general solution of the equation

    (ρ˙y)˙+qy=λrya.e. on [a,b] (3.34)

    has the form

    y=c1y1+c2y2, (3.35)

    where c1,c2 are arbitrary constants, and

    y1=y0n=0λn˜X(2n)=y0uandy2=y0n=0λnX(2n+1)=y0w. (3.36)

    Proof. Let A be the set of all solutions to Eq (3.34). This is a linear space, because

    A={yA:(ρ˙y)˙+(qλr)y=0}.

    From conditions x0Eρ, qλrKH([a,b]), and 1ρL([a,b]), we have by [8, Corollary 3.3] that dim(A)=2. On the other hand, y0AC([a,b])C1ρ, and there exists g0ACG([a,b]) such that

    ρ˙y0=g0a.e. on [a,b].

    By Theorem 3.5, uAC([a,b])C1ρ, ˙u=vρy20 a.e. on [a,b], and vACG([a,b]). Therefore, y1=y0uAC([a,b])C1ρ, and

    ρ˙y1=ρ(y0u)˙=ρ[˙y0u+vρy20y0]=(ρ˙y0)u+vy0=g0u+vy0, (3.37)

    a.e. on [a,b]. Since y0(x)0 for all x[a,b], it follows that 1y0AC([a,b]), thus g0u+vy0ACG([a,b]). Consequently, y1A. In the same way, it is shown that y2A. Now, as y0 is a non-vanishing solution of the homogeneous equation, we can apply Lemma 3.2 and Theorem 3.5, and obtain that

    L[y1]=1y0[ρy20(y0uy0)]=1y0[ρy20vρy20]=1y0˙v=λry0u=λry1, (3.38)

    a.e. on [a,b]. For y2, the proof follows similarly. Thus, y1,y2A. Now, we will verify that y1,y2 are linearly independent. From the definitions of the functions ˜X(2n) and X(2n+1) observe that

    y1(x0)=y0(x0)u(x0)=y0(x0)n=0λn˜X(2n)(x0)=y0(x0)1=y0(x0),w(x0)=n=0λnX(2n+1)(x0)=0,y2(x0)=y0(x0)w(x0)=y0(x0)0=0, andz(x0)=n=0λnX(2n)(x0)=1.

    Also, as x0Eρ, then by Theorem 3.5, ˙w(x0)=z(x0)ρ(x0)y20(x0). Therefore,

    ˙y2(x0)=(˙y0(x0)w(x0)+˙w(x0)y0(x0))=z(x0)ρ(x0)y20(x0)y0(x0)=1ρ(x0)y0(x0). (3.39)

    Then, the generalized Wronskian at x0 is

    [ρW(y1,y2)](x0)=ρ(x0)(y1(x0)˙y2(x0)y2(x0)˙y1(x0))=ρ(x0)(y0(x0)˙y2(x0))=1 (3.40)

    Therefore, y1 and y2 are linearly independent. Moreover, since dimA=2, then {y1,y2} is a basis for A.

    As we can see in Eq (3.36), y1 and y2 are given by the infinite series, but only the terms ˜X(2n) and X(2n+1) are needed. Now, the following result will allow us to avoid computing the unnecessary terms ˜X(2n+1) and X(2n), respectively. This will make the code more numerically efficient.

    Proposition 3.7. Let x0[a,b] and y0A be such that y0(x)0 for all x[a,b]. If P(x)=xx01ρy20 and ˜X(2n), X(2n+1) are the functions defined in (3.3)(3.4), then, for each nN, it is satisfied that

    ˜X(2n)(x)=xx0[P(x)P(t)]y20(t)r(t)˜X(2n2)(t)dt, (3.41)

    and

    X(2n+1)(x)=xx0[P(x)P(t)]y20(t)r(t)˜X(2n1)(t)dt. (3.42)

    Proof. Without loss of generality, let xx0. Note that

    ˜X(2n)(x)=xx01ρ(s)y20(s)˜X(2n1)(s)ds=xx01ρ(s)y20(s)sx0r(t)y20(t)˜X(2n2)(t)dtds. (3.43)

    Define

    g(t,s)={1ρ(s)y02(s), if x0tsx0, if xt>sx0, (3.44)

    and

    f(t)=y20(t)r(t)˜X(2n2)(t). (3.45)

    Then, fKH([a,b]) and

    V[x0,x]g(,s)=1ρ(s)y20(s) (3.46)

    for all s[x,x0]. Since 1ρy20L([a,b]), it follows from [11, Theorem 57] that

    xx0xx0f(t)g(t,s)dtds=xx0xx0f(t)g(t,s)dsdt. (3.47)

    With the left-hand side of this equality, we obtain that

    xx0xx0f(t)g(t,s)dtds=xx0sx0f(t)1ρ(s)y20(s)dtds=xx01ρ(s)y20(s)sx0f(t)dtds=xx01ρ(s)y20(s)sx0y20(t)r(t)˜X(2n2)(t)dtds. (3.48)

    Whereas on the right-hand side of Eq (3.47), we have that

    xx0xx0f(t)g(t,s)dsdt=xx0xtf(t)1ρ(s)y20(s) dsdt=xx0f(t)xtdsρ(s)y20(s) dt.=xx0[P(x)P(t)]y20(t)r(t)˜X(2k2)(t)dt. (3.49)

    Therefore, by Eqs (3.43), (3.48), and (3.49), we can conclude that Eq (3.41) is satisfied. Equality (3.42) is proved in an analogous way.

    Remark 3.8. Note that almost all results depend on the existence of y0, a solution of the homogeneous Eq (3.33), such that y0(x)0 for all x[a,b]. The construction of this function does not represent any difficulty. As a matter of fact, if Ph(x)=xx01ρ(s)ds and

    ˜X(0)h1,  ˜X(2n)h(x)=xx0˜X(2n2)h(t)q(t)[Ph(t)Ph(x)]dt; (3.50)
    X(1)h=Ph,X(2n+1)h(x)=xx0X(2n1)h(t)q(t)[Ph(t)Ph(x)]dt; (3.51)

    then,

    y=c1n=0˜X(2n)h+c2n=0X(2n+1)h (3.52)

    is the general solution of the homogeneous Eq (3.33), and

    y0=n=0˜X(2n)h+in=0X(2n+1)h (3.53)

    is a particular solution of the homogeneous equation such that y0(x)0 for all x[a,b]. Indeed, let us first rewrite the homogeneous Eq (3.33) as (ρ˙ψ)˙=1(q)ψ. Note that ψ01 is a non-vanishing solution of the homogeneous equation (ρ˙ψ)˙=0. Then, by Theorem 3.6, the general solution of the equation (ρ˙ψ)˙=1(q)ψ has the form

    ψ=c1[ψ0n=01n˜Ψ(2n)]+c2[ψ0n=01nΨ(2n+1)]=c1n=0˜Ψ(2n)+c2n=0Ψ(2n+1), (3.54)

    where

    ˜Ψ(0)1,˜Ψ(2n)(x)=xx0˜Ψ(2n1)1ρ(t)ψ20(t)dt,Ψ(1)=R,X(2n+1)(x)=xx0˜Ψ(2n)dtρ(t)ψ20(t)and R(x)=xx01ρψ02=xx01ρ. (3.55)

    Moreover, by Proposition 3.7,

    ˜Ψ(2n)(x)=xx0[R(x)R(t)]ψ02(t)(q)(t)˜Ψ(2n2)(t)dt=xx0[R(t)R(x)]q(t)˜Ψ(2n2)(t)dt,X(2n+1)(x)=xx0[R(x)R(t)]ψ02(t)(q)(t)˜Ψ(2n1)(t)dt=xx0[R(t)R(x)]q(t)˜Ψ(2n1)(t)dt.

    Notice that actually R=Ph, Ψ(2n)=˜X(2n)h, and Ψ(2n+1)=X(2n+1)h. Therefore, y=c1n=0˜X(2n)+c2n=0X(2n+1) is the general solution in A of the homogeneous equation (ρ˙y)˙+qy=0.

    We show now that y0(x)0 for all x[a,b]. Let ψ1=n=0˜Ψ(2n), and ψ2=n=0Ψ(2n+1). Then, y0=ψ1+iψ2. Since ψ1,ψ2A, there exist functions g1,g2ACG([a,b]) such that ρ˙ψ1=g1, and ρ˙ψ2=g2 a.e. on [a,b], specially at every point of Eρ. Therefore, as x0Eρ, we have that

    ρ(x0)˙ψ1(x0)=g1(x0)&ρ(x0)˙ψ2(x0)=g2(x0).

    The functions 1ρ˜Ψ(2n1) and 1ρΨ(2n) are locally continuous at x0, thus by (3.55) and Theorem 2.1,

    [˜Ψ(2n)](x0)=1ρ(x0)˜Ψ(2n1)(x0)=0nN,

    and

    [Ψ(2n+1)](x0)=1ρ(x0)Ψ(2n)(x0)={1ρ(x0),if n=00,if nN.

    Therefore,

    ˙ψ1(x0)=n=0[Ψ(2n)](x0)=0and˙ψ2(x0)=n=0[Ψ(2n+1)](x0)=1ρ(x0).

    Note also that ψ1(x0)=1 and ψ2(x0)=0. Then,

    [ψ1(x0)g2(x0)ψ2(x0)g1(x0)]=g2(x0)=ρ(x0)˙ψ2(x0)=ρ(x0)1ρ(x0)=1.

    It s clear that ψ1,ψ2AC([a,b])ACG([a,b]), therefore, ψ1g2ψ2g1ACG([a,b]). Let x[a,b]. Then,

    ψ1(x)g2(x)ψ2(x)g1(x)1=ψ1(x)g2(x)ψ2(x)g1(x)[ψ1(x0)g2(x0)ψ2(x0)g1(x0)]=xx0[ψ1g2ψ2g1]=xx0[ψ1(ρ˙ψ2)ψ2(ρ˙ψ1)]=xx0[ψ1(ρ˙ψ2)˙ψ2(ρ˙ψ1)˙]=xx0[ψ1(ρ˙ψ2)˙ψ2(ρ˙ψ1)˙]+ψ1qψ2ψ2qψ1=xx0[ψ1[(ρ˙ψ2)˙+qψ2]ψ2[(ρ˙ψ1)˙+qψ1]]=xx0[ψ10ψ20]=xx0 0=0,for all x[a,b].

    Thus, for every x[a,b],

    ψ1(x)g2(x)ψ2(x)g1(x)=1.

    Then, for each x[a,b], ψ1(x)0 or ψ2(x)0. Consequently, y0(x)0 for all x[a,b].

    The representation of the solution of the S-L equation (see (3.35)) allows us to solve spectral problems in a simple way. In this section, we present some aspects related to the numerical implementation of the SPPS method, as well as some examples illustrating how we can apply it to find the eigenvalues of spectral problems.

    Example 4.1. Consider the equation (ρ˙y)˙+qy=λry a.e. on [0,π], where

    ρ(x)=1,q(x)=2πxsin(πx2)andr(x)=1

    with boundary conditions

    y(0)=y(π)=0.

    The function q is highly oscillating, and qKH([0,π])L([0,π]). This example clearly shows that the results of this paper cover wider cases than those results using the Lebesgue integral. Let

    D={yA:y(0)=0=y(π)},

    and define L:DKH([0,π]) as L(y)=(ρ˙y)˙+qy. We will find an approximation for the point spectrum σp(L).

    From Theorem 3.6, the general solution of the equation (ρ˙y)˙+qy=λry has the form

    y=c1[y0s=0λs˜X(2s)]+c2[y0s=0λsX(2s+1)], (4.1)

    where ˜X(2s) and X(2s+1) are defined as in (3.41) by taking x0=0. Using the boundary condition y(0)=0 in (4.1), we obtain that

    0=c1y0(0)+c2y0(0)0.

    Therefore, c1=0. Now, using the boundary condition y(π)=0, we find that

    0=c2y0(π)s=0λsX(2s+1)(π).

    Since the homogeneous solution satisfies y0(π)0, it follows that

    Ω(λ):=s=0λsX(2s+1)(π)=0.

    The zeros of Ω form the point spectrum of L. To find the zeros of this function, we truncate this series after m terms. Then, we have to find the zeros of

    Ωm(λ):=ms=0λsX(2s+1)(π).

    According to Theorem 3.8, to find the values of X(2s+1)(π), we first need to find the values of y0 over [0,π]. Then, in order to build the non-vanishing homogeneous solution y0, as Remark 3.8 states, we must find the values of ˜X(2s)h and X(2s+1)h over [0,π]. From Eq (3.50), we have that

    ˜X(2s)h(x)=xx0[Ph(t)Ph(x)]q(t)˜X(2s2)(t)dt,X(2s+1)h(x)=xx0X(2s1)h(t)q(t)[Ph(t)Ph(x)]dt.

    These integrals are from Kurzweil–Henstock because qKH([0,π])L([0,π]), so we have to use an appropriate method to estimate the integrals since uniform partitions do not work. Instead, we have to use unequal partitions. For this, we will use the method described by Yang et al. in [12]. As q has a singularity at x=0, and near this point it oscillates quite a bit, we take the following sequence of points that approaches x=0 without reaching it:

    xi=π(5(i1)),wherei{1,2,,t+1}.

    Then, on each subinterval [xi+1,xi], we take an unequal partition of size n generated by the points

    u(i)n,0=xi+1,u(i)n,k=xi+1+kj=1a(i)n,j, (4.2)

    where

    a(i)n,j=2(xixi+1)jn(n+1)forj{1,2,,n}. (4.3)

    Therefore, we have that

    0<xt+1=u(t)n,0<u(t)n,1<u(t)n,2<<u(t)n,n=xt=u(t1)n,0<u(t1)n,1<u(t1)n,2<u(t1)n,n=xt1< (4.4)
    <x2=u(1)n,0<u(1)n,1<u(1)n,2<<u(1)n,n=x1=π. (4.5)

    The quadrature that will be used to estimate the value of the integral of a function f over each subinterval [xi+1,xi] is given by

    xixi+1fQ2n(f)=nj=1a(i)n,j2(f(u(i)n,j)+f(u(i)n,j1)). (4.6)

    This quadrature allows us to estimate the values of ˜X(2s)h,X(2s+1)h for s{0,1,,m} at each node u(i)n,j. As mentioned in Remark 3.8, we take y0=ms=0λs˜X(2s)h+ims=0λsX(2s+1)h. This allows us to calculate the values of P and X(2s+1) over [0,π], especially for s=m and x=π, which leads us to the characteristic polynomial Ωm(λ), where the roots of Ωm must be the point spectrum σp(L).

    The eigenvalues in this example were calculated with Python using t=30, n=3000, and m=100.

    Example 4.2. Consider the equation (ρ˙y)˙+qy=λry a.e. on [π,0], where

    ρ(x)=π+x,q(x)=2x+2π+12π+xandr(x)=csc(x+π2)sin(csc(π+x2))

    with boundary conditions

    y(π)=0,˙y(0)=0.

    It is clear that 1ρ,qL([π,0]) and have a singularity at x=π; also, the function r is highly oscillating, and rKH([π,0]). Let

    D={yA:y(π)=0=˙y(0)},

    and define L:DKH([π,0]) as L(y)=(ρ˙y)˙+qy. We will find an approximation for the point spectrum σp(L). From Theorem 3.6, the general solution of the equation (ρ˙y)˙+qy=λry has the form

    y=c1[y0s=0λs˜X(2s)]+c2[y0s=0λsX(2s+1)]=c1y0u+c2y0w, (4.7)

    where ˜X(2s) and X(2s+1) are defined as in (3.41) by taking x0=0. The choice of this point is due to the fact that 0Eρ, note that πEρ. Using the boundary condition y(π)=0 in (4.7), we obtain that

    0=c1[y0(π)s=0λs˜X(2s)(π)]+c2[y0(π)s=0λsX(2s+1)(π)]. (4.8)

    Now, using the boundary condition ˙y(x0)=0 in Eq (4.7), together with the fact that u(x0)=1, v(x0)=0, w(x0)=0, z(x0)=1, ˙u(x0)=v(x0)ρ(x0)y20(x0), and ˙w(x0)=z(x0)ρ(x0)y20(x0), we have that

    0=˙y(x0)=c1(˙y0(x0)u(x0)+y0(x0)˙u(x0))+c2(˙y0(x0)w(x0)+y0(x0)˙w(x0))=c1(˙y0(x0)1+y0(x0)v(x0)ρ(x0)y20(x0))+c2(˙y0(x0)0+y0(x0)z(x0)ρ(x0)y20(x0))=c1˙y0(x0)+c21ρ(x0)y0(x0).

    Then, c2=πy0(0)˙y0(0)c1. When we substitute the value of c2 into Eq (4.8), considering that c1 and y0(π) must not be zero, we will arrive at the next equation

    [s=0λs˜X(2s)(π)]πy0(0)˙y0(0)[s=0λsX(2s+1)(π)]=0.

    Thus, to find the point spectrum of L, we have to find the zeros of

    Ωm(λ):=ms=0λs[˜X(2s)(π)πy0(0)˙y0(0)X(2s+1)(π)].

    The eigenvalues for this example were calculated using t=30, n=3000, and m=100.

    In this paper, we show the convergence of the spectral parameter power series method, proposed by Kravchenko, for the Sturm–Liouville equation with Kurzweil–Henstock integrable coefficients. By incorporating the Kurzweil–Henstock integral into the SPPS method, we have significantly expanded the scope and applicability of the method, allowing us to tackle a wider variety of problems, including those containing highly oscillating functions that are not Lebesgue integrable. The result given by Blancarte et al. in [3, Theorem 7] remains a particular case of the results presented here when q,rL([a,b])(KH([a,b])). The numerical implementation of the method was reasonably tractable and has proven to be a powerful tool for solving Sturm–Liouville problems. This is clearly shown in the examples in Section 4, where we were able to find the point spectrum for problems with Kurzweil–Henstock integrable functions.

    All authors, I. A. Cordero-Martínez, S. Sánchez-Perales and F. J. Mendoza-Torres, have contributed equally to this work. The authors have read and accepted the published version of the manuscript.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The authors are grateful for the support of VIEP-BUAP, within the framework of Project 100100300-VIEP2024, the Universidad Tecnológica de la Mixteca and CONAHCYT.

    The authors declare no conflicts of interest.



    [1] V. V. Kravchenko, A representation for solutions of the Sturm–Liouville equation, Complex Var. Elliptic, 53 (2008), 775–789. https://doi.org/10.1080/17476930802102894 doi: 10.1080/17476930802102894
    [2] V. V. Kravchenko, R. M. Porter, Spectral parameter power series for Sturm–Liouville problems, Math. Method. Appl. Sci., 33 (2010), 459–468. https://doi.org/10.1002/mma.1205 doi: 10.1002/mma.1205
    [3] H. Blancarte, H. M. Campos, K. V. Khmelnytskaya, Spectral parameter power series method for discontinuous coefficients, Math. Method. Appl. Sci., 38 (2015), 2000–2011. https://doi.org/10.1002/mma.3282 doi: 10.1002/mma.3282
    [4] E. Talvila, Henstock–Kurzweil Fourier transforms, Illinois J. Math., 46 (2002), 1207–1226. https://doi.org/10.1215/ijm/1258138475 doi: 10.1215/ijm/1258138475
    [5] R. G. Bartle, A modern theory of integration, Providence: American Mathematical Society, 2001. https://doi.org/10.1090/gsm/032
    [6] R. A. Gordon, The integrals of Lebesgue, Denjoy, Perron, and Henstock, Providence: American Mathematical Society, 1994.
    [7] T. Pérez-Becerra, S. Sánchez-Perales, J. J. Oliveros-Oliveros, The HK-Sobolev space and applications to one-dimensional boundary value problems, J. King Saud Univ. Sci., 32 (2020), 2790–2796. https://doi.org/10.1016/j.jksus.2020.06.016 doi: 10.1016/j.jksus.2020.06.016
    [8] S. Sánchez-Perales, I. A. Cordero-Martínez, H. Kalita, T. Pérez-Becerra, Sturm–Liouville differential equations with Kurzweil–Henstock integrable functions as coefficients, in press.
    [9] W. Rudin, Principles of mathematical analysis, 3 Eds., New York: McGraw-Hill, 1976.
    [10] E. Talvila, Limits and Henstock integrals of products, Real Anal. Exchange, 25 (1999), 907–918.
    [11] V. G. Čelidze, A. G. Džvaršeǐšvili, The theory of Denjoy integral and some applications, Singapore: World Scientific, 1989. https://doi.org/10.1142/0935
    [12] W. C. Yang, P. Y. Lee, X. F. Ding, Numerical integration on some special Henstock–Kurzweil integrals, The Electronic Journal of Mathematics and Technology, 3 (2009), 205–223.
  • Reader Comments
  • © 2024 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1054) PDF downloads(59) Cited by(0)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog