Processing math: 17%
Research article Special Issues

Problems involving combinations of coefficients for the inverse of some complex-valued analytical functions

  • ORCID Number: https://orcid.org/0000-0003-1484-7643
  • Received: 11 August 2024 Revised: 20 September 2024 Accepted: 23 September 2024 Published: 14 October 2024
  • MSC : 30C45, 30C50

  • Inequalities are essential in solving mathematical problems in many different areas of mathematics. Among these, problems involving coefficient combinations that occurred in the Taylor–Maclaurin series of the inverse of complex-valued analytic functions are the challenging ones to solve. In the current article, our aim is to study certain coefficient-related problems that construct from coefficients of the inverse of specific analytic functions. These problems include the Zalcman and Fekete–Szegö inequalities, as well as sharp estimates of the second and third-order Hankel determinants with inverse function coefficients. Also, one of the obtained results gives an improvement of the problem that has been recently published in the journal "AIMS Mathematics".

    Citation: Huo Tang, Muhammad Abbas, Reem K. Alhefthi, Muhammad Arif. Problems involving combinations of coefficients for the inverse of some complex-valued analytical functions[J]. AIMS Mathematics, 2024, 9(10): 28931-28954. doi: 10.3934/math.20241404

    Related Papers:

    [1] Chuanhua Wu, Ziqiang Wang . The spectral collocation method for solving a fractional integro-differential equation. AIMS Mathematics, 2022, 7(6): 9577-9587. doi: 10.3934/math.2022532
    [2] Ali H. Tedjani, Sharifah E. Alhazmi, Samer S. Ezz-Eldien . An operational approach for one- and two-dimension high-order multi-pantograph Volterra integro-differential equation. AIMS Mathematics, 2025, 10(4): 9274-9294. doi: 10.3934/math.2025426
    [3] Obaid Algahtani, M. A. Abdelkawy, António M. Lopes . A pseudo-spectral scheme for variable order fractional stochastic Volterra integro-differential equations. AIMS Mathematics, 2022, 7(8): 15453-15470. doi: 10.3934/math.2022846
    [4] Xiaojun Zhou, Yue Dai . A spectral collocation method for the coupled system of nonlinear fractional differential equations. AIMS Mathematics, 2022, 7(4): 5670-5689. doi: 10.3934/math.2022314
    [5] Mahmoud A. Zaky, Weam G. Alharbi, Marwa M. Alzubaidi, R.T. Matoog . A Legendre tau approach for high-order pantograph Volterra-Fredholm integro-differential equations. AIMS Mathematics, 2025, 10(3): 7067-7085. doi: 10.3934/math.2025322
    [6] Chuanli Wang, Biyun Chen . An hp-version spectral collocation method for fractional Volterra integro-differential equations with weakly singular kernels. AIMS Mathematics, 2023, 8(8): 19816-19841. doi: 10.3934/math.20231010
    [7] Yumei Chen, Jiajie Zhang, Chao Pan . Numerical approximation of a variable-order time fractional advection-reaction-diffusion model via shifted Gegenbauer polynomials. AIMS Mathematics, 2022, 7(8): 15612-15632. doi: 10.3934/math.2022855
    [8] Kanagaraj Muthuselvan, Baskar Sundaravadivoo, Kottakkaran Sooppy Nisar, Suliman Alsaeed . Discussion on iterative process of nonlocal controllability exploration for Hilfer neutral impulsive fractional integro-differential equation. AIMS Mathematics, 2023, 8(7): 16846-16863. doi: 10.3934/math.2023861
    [9] Saowaluck Chasreechai, Sadhasivam Poornima, Panjaiyan Karthikeyann, Kulandhaivel Karthikeyan, Anoop Kumar, Kirti Kaushik, Thanin Sitthiwirattham . A study on the existence results of boundary value problems of fractional relaxation integro-differential equations with impulsive and delay conditions in Banach spaces. AIMS Mathematics, 2024, 9(5): 11468-11485. doi: 10.3934/math.2024563
    [10] Chang Phang, Abdulnasir Isah, Yoke Teng Toh . Poly-Genocchi polynomials and its applications. AIMS Mathematics, 2021, 6(8): 8221-8238. doi: 10.3934/math.2021476
  • Inequalities are essential in solving mathematical problems in many different areas of mathematics. Among these, problems involving coefficient combinations that occurred in the Taylor–Maclaurin series of the inverse of complex-valued analytic functions are the challenging ones to solve. In the current article, our aim is to study certain coefficient-related problems that construct from coefficients of the inverse of specific analytic functions. These problems include the Zalcman and Fekete–Szegö inequalities, as well as sharp estimates of the second and third-order Hankel determinants with inverse function coefficients. Also, one of the obtained results gives an improvement of the problem that has been recently published in the journal "AIMS Mathematics".



    The study of various phenomena in science and engineering often leads to the formulation of systems of fractional integro-differential equations (FIDEs). Examples of such systems can be found in electric circuit analysis, the activity of interacting inhibitory and excitatory neurons [1], glass-forming processes, non-hydrodynamics, drop-wise condensation, and wind ripple formation in deserts [2,3]. Due to the presence of fractional derivative operators, obtaining analytical solutions for these functional equations is generally infeasible, like the fractional derivative model of viscoelasticity, the so-called fractional Zener model [4,5]. Consequently, researchers have refined existing methods to provide semi-analytical or numerical solutions for FIDEs. For instance, Wang et al. proposed the use of Bernoulli wavelets and operational matrices to solve coupled systems of nonlinear fractional-order integro-differential equations [6]. Dief and Grace developed a new technique based on iterative refinement to approximate the analytical solution of a linear system of FIDEs [7]. In [3], the single term Walsh series (STWS) method was employed to handle second-order Volterra integro-differential equations. The Muntz-Legendre wavelets were applied to find approximate solutions for systems of fractional integro-differential Volterra-Fredholm equations [8]. Heydari et al. presented a Chebyshev wavelet method for solving a class of nonlinear singular fractional Volterra integro-differential equations [9]. Mohammed and Malik applied a modified series algorithm to solve systems of linear FIDEs [10]. In [11], Genocchi polynomials combined with the collocation method were utilized to numerically solve a system of Volterra integro-differential equations. Youbi et al. introduced an iterative reproducing kernel algorithm to investigate approximate solutions for fractional systems of Volterra integro-differential equations in the Caputo-Fabrizio operator sense [12]. Akbar et al. extended the optimal homotopy asymptotic method to systems of fractional order integro-differential equations [13]. Wang et al. combined a mixed element method with the second-order backward difference scheme to numerically solve a class of two-dimensional nonlinear fourth-order partial differential equations [14].

    Spectral methods offer semi-analytic approximate solutions to various functional equations by expressing them as linear combinations of basis functions. The three main spectral methods are collocation, Galerkin, and tau methods. Orthogonal polynomials serve as fundamental basis functions in spectral methods. Classic polynomials such as Jacobi, Hermite, and Laguerre polynomials are commonly employed for numerical solutions of diverse equations. For instance, Legendre polynomials and Chebyshev polynomials of the first to sixth kind, which are special cases of Jacobi polynomials, are applied in spectral methods for solving multidimensional partial Volterra integro-differential equations, nonlinear fractional delay systems, distributed-order fractional differential equations, fractional-order wave equations, population balance equations, fractional-order diffusion equations, and Volterra-Fredholm integral equations [15,16,17,18,19,20,21].

    While Gegenbauer polynomials, a specific form of Jacobi polynomials, are less commonly used compared to their orthogonal counterparts, they have found applications in various areas. Usman et al. utilized Gegenbauer polynomials to approximate solutions for multidimensional fractional-order delay problems arising in mathematical physics and engineering [22]. Shifted Gegenbauer polynomials have been employed in extracting features from color images [23]. Alkhalissi et al. introduced the generalized Gegenbauer-Humbert polynomials for solving fractional differential equations [24]. Faheem and Khan proposed a wavelet collocation method based on Gegenbauer polynomials for solving fourth-order time-fractional integro-differential equations with a weakly singular kernel [25]. In [26], a Gegenbauer wavelets method was utilized to solve FIDEs.

    In this study, we focus on solving a system of Caputo fractional-order Volterra integro-differential equations with variable coefficients in the following form:

    C0Dμr,nxyr(x)+n1k=1ξr,k(x)C0Dμr,nkxyr(x)+ξr,nyr(x)2j=1νr,jx0κr,j(x,t)C0Dγr,2jtyj(t)dtfr(x)=0 (1.1)

    where xI=[0,1] and r=1,2 and the initial conditions are

    dmrdxyr(0)=yr,mr,mr=0,1,,Mr1,r=1,2 (1.2)

    In (1.1), ξr,kC(R),ξr,0(x)=1,r=1,2,k=0,1,,n, μr,k,γr,j(0,1], such that μr,n>μt,n1>>μr,1>0,γr,1>γr,0>0,Mr=max{pr,j,qr,k},qr,k1<μr,kqr,k, pr,j1<γr,jpr,j,r=1,2,j=1,2,k=1,2,,n, yr(x),r=1,2 are real continuous functions, frC[0,1], κr,jC([0,1]×[0,1]) are known functions, νr,jR,r,j=1,2, and C0Dμx is the Caputo fractional derivative operator. A rigid plate submerged in Newtonian fluid can be modeled using FIDEs like (1.1) [27]. Among the important special cases of (1.1), the Bagley-Torvik equation with fractional-order derivative describes the motion of physical systems in Newtonian fluids [28]. Structural dynamics most frequently use the Kelvin-Voigt model, which is another special case [29].

    The use of Gegenbauer polynomials in addressing various types of fractional equations, particularly FIDEs, has been relatively limited (readers can refer to [22,23,24,25,26]). In this study, our objective is to develop a combined approach using the tau method and Gegenbauer polynomials. Initially, we investigate the existence and uniqueness of solutions to these equations by leveraging Krasnoselskii's fixed point theorem. Subsequently, we derive integral operational matrices of both integer and fractional orders, as well as the operational matrix of the product, specifically tailored to Gegenbauer polynomials. Through appropriate approximations utilizing these operational matrices, the original system is transformed into an algebraic system. By employing the tau spectral method and the inner product by basis functions, we eliminate the independent variable x and obtain a system of 2(N+1) algebraic equations that determine the coefficients of the series solutions. Solving this system allows us to approximate the solutions of the main system. Additionally, we estimate an error bound for the residual function within a Gegenbauer-weighted Sobolev space, which reveals that increasing the number of basis functions in the series solutions leads to smaller errors.

    The objectives of this paper can be summarized as follows:

    ● Investigating the existence and uniqueness of solutions for the system of FIDEs with variable coefficients.

    ● Expressing the approximate solutions of the model (1.1) as linear combinations of shifted Gegenbauer polynomials.

    ● Developing operational matrices for integration and product operations associated with Gegenbauer polynomials.

    ● Estimating error bounds for the approximate solutions and residual functions within a Gegenbauer-weighted Sobolev space.

    The structure of this paper is as follows: Section 2 provides a brief review of relevant definitions in fractional calculus. The existence and uniqueness of solutions to the system (1.1) are investigated in Section 3. In Section 4, the shifted Gegenbauer polynomials are introduced, their connection to Jacobi polynomials is stated, and operational matrices for integration and product operations are derived. Section 5 outlines the necessary approximations for the functions involved in the system (1.1). Error bounds for the approximate solutions and residual functions are estimated in Section 6. The effectiveness of the proposed method is demonstrated through two numerical examples in Section 7. Finally, Section 8 provides concluding remarks.

    In this section, some definitions and properties of the fractional derivative and integral operators are recalled.

    Definition 2.1. The well-known non-integer derivative operator in the Caputo sense of the differentiable function g of the order μ(0,1) is defined below [30]:

    C0Dμxg(x)=1Γ(1μ)x0(xs)μg(s)ds,0<μ<1 (2.1)

    The following properties are achieved directly from Definition 2.1:

    1) C0Dμxϱ=0,ϱR;

    2) C0Dμx(λ1g1(x)+λ2g2(x))=λ1C0Dμxg1(x)+λ2C0Dμxg2(x),λ1,λ2R;

    3) C0Dμxxν={Γ(ν+1)Γ(νμ+1)xνμ,μ>ν,0,otherwise

    Definition 2.2. The Riemann-Liouville integral operator of the function gC[0,1] of the order μ is defined below [30]:

    RL0Iμxg(x)=1Γ(μ)x0(xs)μ1g(s)ds,μ>1 (2.2)

    The following properties follow from Definition 2.2:

    1) RL0Iμ1x(RL0Iμ2xg(x))=RL0Iμ1+μ2xg(x),μ1,μ2R;

    2) RL0Iμxg(x)=RL0Isμx(Dsg(x)),s=μ,Ds=dsdxs;

    3) RL0Iμxxν={Γ(ν+1)Γ(ν+μ+1)xν+μ,ν,μR+,0,otherwise;

    4) RL0Iμx(C0Dμxg(x))=g(x)g(0)

    This section deals with the existence of unique solutions to system (1.1). By applying the Riemann-Liouville integral operator of order μr,n(0,1) to Eq (1.1), the following equivalent equation is achieved:

    yr(x)=yr,0n1k=11Γ(μr,n)x0(xt)μr,n1ξr,k(t)C0Dμr,nktyr(t)dt1Γ(μr,n)x0(xt)μr,n1ξr,n(t)yr(t)dt+2j=1νr,jΓ(μr,n)x0(xt)μr,n1t0κr,j(t,s)C0Dγr,2jsyj(s)dsdt+1Γ(μr,n)x0(xt)μr,n1fr(t)dt (3.1)

    Now, suppose that C(I,X) is a Banach space of real-valued continuous functions from I=[0,1] into XR equipped with the following norm

    zC=max{supxI|z(x)|,supxI|C0Dμxz(x)|},zC(I,X)

    and the following assumptions hold for any xI and (x,t)I×I:

    Mr,j=sup(x,t)I|κr,j(x,t)|,r,j=1,2,Pr,k=sup(x)I|ξr,k(x)|,r=1,2,k=1,,n,Fr=sup(x)I|fr(x)|,r=1,2 (3.2)

    Theorem 3.1. Suppose that the assumptions in (3.2) and the following inequality hold

    Pr,nΓ(μr,n+1)<1,nk=1Pr,kΓ(μr,n+1)+2j=1|νr,j|Mr,jΓ(μr,n+2)<1 (3.3)

    then, problems (1.1)–(1.2) have a unique solution on C(I,X).

    Proof. Let Dq={zC(I,X)|zCq} subject to

    q|yr,0|+FrΓ(μr,n+1)1{nk=1Pr,kΓ(μr,n+1)+2j=1|νr,j|Mr,jΓ(μr,n+2)} (3.4)

    Then, Dq is a closed, bounded, and convex subset of C(I,X). The operators B1 and B2 are defined as shown below:

    B1yr(x)=yr,01Γ(μr,n)x0(xt)μr,n1ξr,n(t)yr(t)dt+1Γ(μr,n)x0(xt)μr,n1fr(t)dt
    B2yr(x)=2j=1νr,jΓ(μr,n)x0(xt)μr,n1t0κr,j(t,s)C0Dγr,2jsyj(s)dsdtn1k=11Γ(μr,n)x0(xt)μr,n1ξr,k(t)C0Dμr,nktyr(t)dt

    It's necessary to be shown that B1+B2 has a fixed point in Dq. The process of the proof is divided into four stages:

    Stage 1. It is shown that B1yr(x)+B2ur(x)Dq for every yr,urDq. Using (3.2) and (3.4), one obtains the following:

    B1yr+B2urC|yr,0|+Pr,nyrCΓ(μr,n+1)+FrΓ(μr,n+1)+2j=1|νr,j|Mr,jurCΓ(μr,n+2)+n1k=1Pr,kurCΓ(μr,n+1)|yr,0|+FrΓ(μr,n+1)+{Pr,nyrCΓ(μr,n+1)+n1k=1Pr,kurCΓ(μr,n+1)+2j=1|νr,j|Mr,jurCΓ(μr,n+2)}qq

    Therefore, B1yr+B2urCq, which implies that B1yr+B2urDq for any yr,urDq.

    Stage 2. It is shown that the operator B1 is a contraction mapping on Dq. For each yr,urDq and each xI, one gets the following

    B1yrB1urC=1Γ(μr,n)x0(xt)μr,n1ξr,n(t)(ur(t)yr(t))dtCPr,nΓ(μr,n+1)yrurC

    From (3.3), this shows that B1 is a contraction mapping on Dq.

    Stage 3. It's shown that the operator B2 is compact and continuous. For yrDq, one has

    B2yrC=2j=1νr,jΓ(μr,n)x0(xt)μr,n1t0κr,j(t,s)C0Dγr,2jsyj(s)dsdtn1k=11Γ(μr,n)x0(xt)μr,n1ξr,k(t)C0Dμr,nktyr(t)dtC2j=1|νr,j|Mr,jΓ(μr,n+1)yjC+n1k=1Pr,kΓ(μr,n+2)yrC{2j=1|νr,j|Mr,jΓ(μr,n+1)+n1k=1Pr,kΓ(μr,n+2)}q

    This shows that B2 is uniformly bounded on Dq. It remains to prove the compactness of the operator B2. For x1,x2I such that x1<x2 and yrDq, one has

    B2yr(x2)B2yr(x1)=2j=1νr,jΓ(μr,n)x20(x2t)μr,n1t0κr,j(t,s)C0Dγr,2jsyj(s)dsdtn1k=11Γ(μr,n)x20(x2t)μr,n1ξr,k(t)C0Dμr,nktyr(t)dt2j=1νr,jΓ(μr,n)x10(x1t)μr,n1t0κr,j(t,s)C0Dγr,2jsyj(s)dsdt+n1k=11Γ(μr,n)x10(x1t)μr,n1ξr,k(t)C0Dμr,nktyr(t)dt=2j=1νr,jΓ(μr,n)x20[(x2t)μr,n1(x1t)μr,n1]t0κr,j(t,s)C0Dγr,2jsyj(s)dsdt+2j=1νr,jΓ(μr,n)x2x1(x1t)μr,n1t0κr,j(t,s)C0Dγr,2jsyj(s)dsdtn1k=11Γ(μr,n)x20[(x2t)μr,n1(x1t)μr,n1]ξr,k(t)C0Dμr,nktyr(t)dtn1k=11Γ(μr,n)x2x1(x1t)μr,n1ξr,k(t)C0Dμr,nktyr(t)dt

    By taking norm, one gets

    B2yr(x2)B2yr(x1)C2j=1|νr,j|Mr,jqΓ(μr,n)x20[(x2t)μr,n1(x1t)μr,n1]tdt+2j=1|νr,j|Mr,jqΓ(μr,n)x20(x1t)μr,n1tdt+n1k=1Pr,kqΓ(μr,n)x20[(x2t)μr,n1(x1t)μr,n1]dt+n1k=1Pr,kqΓ(μr,n)x2x1(x1t)μr,n1dt

    Using a change of variable, one gets:

    B2yr(x2)B2yr(x1)C2j=1|νr,j|Mr,jqΓ(μr,n){0x2(x2u)uμr,n1du+x1x2x1(x1u)uμr,n1dux1x20(x1u)uμr,n1du}+n1k=1Pr,kqΓ(μr,n){0x2uμr,n1dux1x2x1uμr,n1du+x1x20uμr,n1du}=2j=1|νr,j|Mr,jqΓ(μr,n)(xμr,n+12μr,n(μr,n+1)+x1(x1x2)μr,nμr,n(x1x2)μr,n+1μr,n+1xμr,n+11μr,n(μr,n+1)x1(x1x2)μr,nμr,n+(x1x2)μr,n+1μr,n+1)+n1k=1Pr,kqΓ(μr,n)(xμr,n2μr,n+(x1x2)μr,nμr,nxμr,n1μr,n(x1x2)μr,nμr,n)=2j=1|νr,j|Mr,jqΓ(μr,n+2)(xμr,n+12xμr,n+11)+n1k=1Pr,kqΓ(μr,n+1)(xμr,n1xμr,n2)

    As x2 tends to x1, the righthand side of the above inequality tends to zero and B2 is equicontinuous. From Stages 1–3 along with the Arzela-Ascoli theorem, one deduces that the operator B2 is compact and continuous [31]. Finally, by Krasnoselskii's fixed-point theorem [32], problem (1.1) has at least a set of solutions as (y1(x),y2(x)) on I.

    Stage 4. Define Eyr(x)=B1yr(x)+B2yr(x). For any yr,urC(I,X) and xI one has:

    EyrEurC=n1k=11Γ(μr,n)x0(xt)μr,n1ξr,k(t)C0Dμr,nks(ur(t)yr(t))dt+1Γ(μr,n)x0(xt)μr,n1ξr,n(t)(ur(t)yr(t))dt+2j=1νr,jΓ(μr,n)x0(xt)μr,n1t0κr,j(t,s)C0Dγr,2js(uj(s)yj(s))dsdtCn1k=1Pr,kΓ(μr,n+1)yrurC+Pr,nΓ(μr,n+1)yrurC+2j=1|νr,j|Mr,jΓ(μr,n+2)yrurC

    Imposing the hypothesis in (3.3) implies that E is a contraction mapping. It follows that E has a unique fixed point, which is a solution of problems (1.1)–(1.2).

    The shifted Gegenbauer polynomials (SGPs) Gσi(x),i=0,1,2, are orthogonal polynomials on the interval [0,1]

    textcolorred, with respect to the weight function ωσ(x)=xσ12(1x)σ12,σ>12, that can be defined by the following recurrence relation:

    Gσi+1(x)=2(i+σ)i+1(2x1)Gσi(x)i+2σ1i+1Gσi1(x),i=1,2,,xI,Gσ0(x)=1,Gσ1(x)=2σ(2x1)

    where I=[0,1]. If Gσi(x) and Gσj(x) are the Gegenbauer polynomials of degrees i and j, respectively, then they satisfy the following relation:

    10Gσi(x)Gσj(x)ωσ(x)dx={0,ij,hσi,i=j (4.1)

    where hσi is the normalizing factor as follows:

    hσi=Γ(σ+12)2Γ(i+2σ)Γ(2σ)2(2i+2σ)i!,i=0,1,2, (4.2)

    These polynomials can be represented in series form as

    Gσi(x)=ik=0ρσk,ixk,xI (4.3)

    where the coefficients ρσk,i,k=0,1,,i are computed below

    ρσk,i=(1)ikΓ(σ+12)Γ(i+k+2σ)Γ(2σ)Γ(i+σ+12)(ik)!k!,i=0,1,,k=0,1,,i (4.4)

    Remark 4.1. The shifted Gegenbauer polynomials are classified as an especial case of the classic Jacobi polynomials. The relation between these polynomials is as [34]:

    Gσi(x)=i!Γ(σ+12)Γ(i+σ+12)Jσ12,σ12i(x) (4.5)

    where Jα,βi(x) is the shifted Jacobi polynomials of the degree i. The normalizing factor of the shifted Jacobi polynomials regarding the weight function wα,β(x)=xβ(1x)α is

    ςα,βi=Γ(i+α+1)Γ(i+β+1)(2i+α+β+1)Γ(i+1)Γ(i+α+β+1),i=0,1,2, (4.6)

    and the l-th derivative of Jα,βi(x) w.r.t. x is

    dlJα,βi(x)dxl=Γ(l+i+α+β)Γ(i+α+β+1)Jα+l,β+lil(x) (4.7)

    The square-integrable function yL2(I) can be considered as a linear combination of the SGPs, that is

    y(x)=k=0ykGσk(x) (4.8)

    where the coefficients yk,k=0,1, are calculated by the following relation:

    yk=1hσk10y(x)Gσk(x)ωσ(x)dx. (4.9)

    To use the shifted Gegenbauer polynomials as basis functions and present approximations in matrix form, a finite form of the series in (4.8) is considered:

    y(x)yN(x)=Nk=0ykGσk(x)=YTΛσ(x)=ΛσT(x)Y, (4.10)

    where Y and Λ(x) are (N+1)×1 vectors as follows

    Y=[y0,y1,y2,,yN]T,Λσ(x)=[Gσ0(x),Gσ1(x),Gσ2(x),,GσN(x)]T (4.11)

    To deduce computational time and avoid multiplying or integrating the basis functions, introducing and using operational matrices are recommended.

    Consider the vector Λσ(x) in (4.11); the integral of this vector can be represented in a matrix form. To compute the integral operational matrix, the i-th component of Λσ(x) is first considered. Using the series given by (4.3), the integral of Gσi(x) can be computed as follows:

    x0Gσi(s)ds=ik=0ρσk,ix0skds=ik=0ρσk,ixk+1k+1 (4.12)

    Now, xk+1 is approximated in terms of the Gegenbauer polynomials:

    xk+1Nj=0dσj,k+1Gσj(x)

    where

    dσl,k+1=1hσj10xk+1Gσj(x)ωσ(x)dx=1hσjjl=0ρσl,j10xk+1xlxσ12(1x)σ12dx=1hσjjl=0ρσl,jΓ(k+l+σ+32)Γ(σ+12)Γ(k+l+2σ+2)

    Therefore, the integral in (4.12) will be as

    x0Gσi(s)dsNj=0{ik=0jl=0ρσk,iρσl,jΓ(k+l+σ+32)Γ(σ+12)hσj(k+1)Γ(k+l+2σ+2)}Gσj(x),i=0,1,,N (4.13)

    Thus, (4.13) can be written in matrix form as

    x0Λσ(s)dsΘ(σ)Λσ(x) (4.14)

    where Θ(σ) is called the operational matrix of the integration and its entries are calculated below:

    Θ(σ)i,j=ik=0jl=0ρσk,iρσl,jΓ(k+l+σ+32)Γ(σ+12)hσj(k+1)Γ(k+l+2σ+2),i,j=0,1,,N

    Similar to what was stated in the previous subsection, the Riemann-Liouville integral of the i-th Gegenbauer polynomial, Gσi(x), can be calculated as follows:

    RL0Iμx(Gσi(x))=ik=0ρσk,iRL0Iμx(xk)=ik=0ρσk,iΓ(k+1)Γ(k+μ+1)xk+μ (4.15)

    The function xk+μ can be approximated as follows

    xk+μNj=0bσj,k+μGσj(x)

    where

    bσj,k+μ=1hσj10xk+μGσj(x)ωσ(x)dx=1hσjjl=0ρσl,j10xk+l+μ+σ12(1x)σ12dx=1hσjjl=0ρσl,jΓ(k+l+μ+σ+12)Γ(σ+12)Γ(k+l+μ+2σ+1)

    So, (4.15) is written as

    RL0Iμx(Gσi(x))Nj=0{ik=0jl=0ρσk,iρσl,jΓ(k+1)Γ(k+l+μ+σ+12)Γ(σ+12)hσjΓ(k+μ+1)Γ(k+l+μ+2σ+1)}Gσj(x),i=0,1,,N

    Therefore, one gets

    RL0Iμx(Λσ(x))Θ(μ,σ)Λσ(x) (4.16)

    where Θ(μ,σ) is called the operational matrix of the integration of the order σ and its entries are calculated as

    Θ(μ,σ)i,j=ik=0jl=0ρσk,iρσl,jΓ(k+1)Γ(k+l+μ+σ+12)Γ(σ+12)hσjΓ(k+μ+1)Γ(k+l+μ+2σ+1),i,j=0,1,,N

    After substituting appropriate approximations into Eq (1.1), terms like x0Λσ(t)ΛσT(t)Vdt are appeared where V is an (N×1) vector. To reduce the computational time of the product of the vectors Λσ(x) and ΛσT(x), an algorithm is suggested and this product is accomplished in a matrix form:

    Λσ(x)ΛσT(x)VVΛσ(x) (4.17)

    where V is the (N×N) operational matrix of the product corresponding to the vector V. As an approximate way of calculating its entries, the following steps can be proceeded:

    Step 1. If Gσj(x) and Gσk(x) are the Gegenbauer polynomials of degrees j and k, respectively, compute their products as follows:

    Gσj(x)Gσk(x)=j+kr=0π(j,k)rxr (4.18)

    where coefficients π(j,k)r,r=0,1,,j+k are calculated as:

      if jk then

        for r=0..k+j do

          if r>j then

            π(j,k)r=kl=rjρσrl,jρσl,k

          else

            r=min{r,k}

            π(j,k)r=rl=0ρσrl,jρσl,k

          end if

        end for

      else

        for r=0..k+j do

          if rj then

            r=min{r,j}

            π(j,k)r=rl=0ρσrl,jρσl,k

          else

            r=min{r,k}

            π(j,k)r=rl=rjρσrl,jρσl,k

          end if

        end for

      end if

    Step 2. Using (4.18), compute the integral of the product of three Gegenbauer polynomials Gσi(x)Gσj(x), and Gσk(x) as follows

    qi,j,k=10Gσi(x)Gσj(x)Gσk(x)ωσ(x)dx=j+kr=0π(j,k)r10xrGσi(x)ωσ(x)dx=j+kr=0il=0π(j,k)rρσl,iΓ(r+l+σ+12)Γ(σ+12)Γ(r+l+2σ+1) (4.19)

    Step 3. The entries of V are calculated below:

    Vj,k=1hσkNi=0Viqi,j,k,j,k=0,1,,N (4.20)

    For more details, please refer to [33].

    Remark 4.2. In the process of approximating the integral parts in system (1.1), the following integral may be appeared:

    10ΛT(x)KWΘ(σ)Λ(x)dx

    where K is an (N×N) known matrix, W is the (N×N) product operational matrix corresponding to the given W, and Θ(σ) is the integral operational matrix introduced in (4.14). The above integral can be approximated as follows:

    10ΛT(x)KWΘ(σ)Λ(x)dxΔ (4.21)

    where Δ is an (N×1) vector with the following components:

    Δs=N+1m=1N+1n=1am,nqs1,m1,n1,s=1,2,,N+1,am,n=N+1i=1N+1j=1Km,jWj,iΘ(σ)i,n,m,n=1,2,,N+1 (4.22)

    Remark 4.3. If Λ(x) is the basis vector in (4.11), one has

    10Λ(x)ΛT(x)ωσ(x)dx=Qσ (4.23)

    where Qσ is the (N+1)×(N+1) diagonal matrix such that Qσi,i=hσi,i=0,1,,N.

    In this section, we proceed to approximate the functions and integral components of Eq (1.1) by utilizing the operational matrices derived in Section 4. To this end, we apply the proposed methodology to two distinct systems of fractional Volterra integro-differential equations featuring variable coefficients.

    Consider the following system of fractional Volterra integro-differential equations with variable coefficients:

    C0D0.9xy0(x)+C0D0.6xy0(x)+2xy0(x)x05exp(x)C0D0.3ty0(t)dtf1(x)=0,C0D0.8xy1(x)+xC0D0.5xy1(x)x0(sin(x)t)C0D0.8ty0(t)dtf2(x)=0 (5.1)

    with the initial conditions y0(0)=0,y1(0)=1. According to the highest orders of derivatives in Eq (5.1), the following approximations are considered:

    C0D0.9xy0(x)YT0Λσ(x),C0D0.8xy1(x)YT1Λσ(x) (5.2)

    where Y0 and Y1 are the vectors of unknown coefficients and must be determined. Integrating Eq (5.2) along with the initial conditions leads to the following approximations:

    y0(x)YT0Θ(0.9,σ)Λσ(x)+y0(0)=WT1Λσ(x),W1=Θ(0.9,σ)TY0,y1(x)YT1Θ(0.8,σ)Λσ(x)+y1(0)=YT1Θ(0.8,σ)Λσ(x)+ET1Λ(x)=ZT1Λσ(x),Z1=E1+Θ(0.8,σ)TY1 (5.3)

    For approximating other terms in (5.1), approximations in Eq (5.2) are written as follows:

    C0D0.9xy0(x)=C0D0.3x(D0.6xy0(x))YT0Λσ(x) (5.4)

    so, one has

    C0D0.6xy0(x)YT0Θ(0.3,σ)Λσ(x)+C0D0.6xy0(0)=WT2Λσ(x),W2=Θ(0.3,σ)TY0 (5.5)

    Using the representation (5.4), one can get the following expressions:

    C0D0.9xy0(x)=C0D0.6x(D0.3xy0(x))YT0Λσ(x)D0.3xy0(x)WT3Λσ(x),W3=Θ(0.6,σ)TY0, (5.6)
    C0D0.9xy0(x)=C0D0.1x(D0.8xy0(x))YT0Λσ(x)D0.8xy0(x)WT4Λσ(x),W4=Θ(0.1,σ)TY0 (5.7)

    Similarly, one has

    C0D0.8xy1(x)=C0D0.3x(D0.5xy1(x))YT1Λσ(x)D0.5xy1(x)ZT2Λσ(x),Z2=Θ(0.3,σ)TY1 (5.8)

    Now, using (4.12) and (5.3) yields the following approximations for System I coefficients:

    2xF1Λσ(x),xF2Λσ(x),2xy0(x)F1Λσ(x)ΛσT(x)W1F1W1Λσ(x),xD0.5xy1(x)F2Λσ(x)ΛσT(x)Z2F2Z2Λσ(x) (5.9)

    where W1 and Z2 are (N+1)×(N+1) operational matrices of the product corresponding to the vectors W1 and Z2, respectively. The kernels and integral parts can be approximated as follows:

    5exp(x)ΛσT(x)K1Λσ(t),sin(x)tΛσT(x)K2Λσ(t),x05exp(x)D0.3ty0(t)dtΛσT(x)K1W3x0Λσ(t)dtΛσT(x)K1W3Θ(σ)Λσ(x),x0(sin(x)t)D0.8ty0(t)dtΛσT(x)K2W4x0Λσ(t)dtΛσT(x)K2W4Θ(σ)Λσ(x) (5.10)

    where Θ(σ) is the operational matrix of the integration in (4.14) and W3 and W4 are the operational matrices of the product. To use the tau method, the sources functions f1 and f2 should be approximated in terms of the Gegenbauer polynomials:

    f1(x)FT3Λσ(x),f2(x)FT4Λσ(x) (5.11)

    where vectors F3 and F4 are obtained using Eq (4.9). By substituting approximations (5.2)–(5.11) into system (5.1), the following algebraic system is achieved:

    R1(x)=YT0Λσ(x)+WT2Λσ(x)+FT1W1Λσ(x)ΛσT(x)K1W3Θ(σ)Λσ(x)FT3Λσ(x)0,R2(x)=YT1Λσ(x)+FT2Z2Λσ(x)ΛσT(x)K2W4Θ(σ)Λσ(x)FT4Λσ(x)0 (5.12)

    The inner product of algebraic system (5.12) by the basis vector Λσ(x) leads to the following system of algebraic equations:

    {YT0Q+WT2Q+FT1W1QΔ1FT3Q0,YT1Q+ZT2QΔ2FT4Q0 (5.13)

    where Δi,i=1,2, and Q are (N+1)×(N+1) matrices in (4.21) and (4.23), respectively.

    Now, consider the following system of fractional Volterra integro-differential equations with variable coefficients as the second illustrative example:

    C0D0.7xy0(x)+C0D0.5xy0(x)2y0(x)x0(xC0D0.3ty0(t)+(xt)y1(t))dtf1(x)=0,C0D0.6xy1(x)+exp(x)C0D0.4xy1(x)+xy1(x)x0(x2t)C0D0.8ty0(t)dtf2(x)=0 (5.14)

    with the initial conditions y_0(0) = y_1(0) = 0 . Based on the highest orders of derivatives in Eq (5.14), the following approximations are set:

    \begin{equation} {_0^C}\mathsf{D}_x^{0.8} y_0(x) \approx \mathcal{Y}_0 ^T\, \Lambda ^{\sigma}(x), \quad {_0^C}\mathsf{D}_x^{0.6} y_1(x) \approx \mathcal{Y}_1 ^T\, \Lambda ^{\sigma}(x) \end{equation} (5.15)

    Integrating the approximations in (5.15) along with the initial conditions leads to the following approximations:

    \begin{equation} \begin{aligned} & y_0(x) \approx \mathcal{Y}_0 ^T\, \mathbf{\Theta}^{(0.8, \sigma)}\, \Lambda^{\sigma}(x)+y_0(0) = \mathcal{W}_1^T \, \Lambda ^{\sigma}(x), \, \, \mathcal{W}_1 = {\mathbf{\Theta}^{(0.8, \sigma)}}^T\, \mathcal{Y}_0, \\ & y_1(x) \approx \mathcal{Y}_1 ^T\, \mathbf{\Theta}^{(0.6, \sigma)}\, \Lambda^{\sigma}(x)+y_1(0) = \mathcal{Z}_1^T \, \Lambda ^{\sigma}(x), \, \, \mathcal{Z}_1 = {\mathbf{\Theta}^{(0.6, \sigma)}}^T\, \mathcal{Y}_1 \end{aligned} \end{equation} (5.16)

    For approximating other terms in (5.14), approximations in Eq (5.15) are written as follows:

    \begin{equation} \begin{aligned} &{_0^C}\mathsf{D}_x^{0.8} y_0(x) = {_0^C}\mathsf{D}_x^{0.1}(\mathsf{D}_x^{0.7} y_0(x)) \approx \mathcal{Y}_0 ^T\, \Lambda ^{\sigma}(x), \\ &\quad \Longrightarrow \, \, \mathsf{D}_x^{0.7} y_0(x) \approx \mathcal{Y}_0 ^T\, \mathbf{\Theta}^{(0.1, \sigma)}\, \Lambda(x) = \mathcal{W}_2^T\, \Lambda(x), \quad \mathcal{W}_2 = {\mathbf{\Theta}^{(0.1, \sigma)}}^T\, \mathcal{Y}_0, \\ &{_0^C}\mathsf{D}_x^{0.8} y_0(x) = {_0^C}\mathsf{D}_x^{0.5}(\mathsf{D}_x^{0.3} y_0(x)) \approx \mathcal{Y}_0 ^T\, \Lambda ^{\sigma}(x), \\ & \quad \Longrightarrow \, \, \mathsf{D}_x^{0.3} y_0(x) \approx \mathcal{Y}_0 ^T\, \mathbf{\Theta}^{(0.5, \sigma)}\, \Lambda(x) = \mathcal{W}_3^T\, \Lambda(x), \quad \mathcal{W}_3 = {\mathbf{\Theta}^{(0.5, \sigma)}}^T\, \mathcal{Y}_0, \\ & {_0^C}\mathsf{D}_x^{0.8} y_0(x) = {_0^C}\mathsf{D}_x^{0.3}(\mathsf{D}_x^{0.5} y_0(x)) \approx \mathcal{Y}_0 ^T\, \Lambda ^{\sigma}(x), \\ &\quad \Longrightarrow \, \, \mathsf{D}_x^{0.5} y_0(x) \approx \mathcal{Y}_0 ^T\, \mathbf{\Theta}^{(0.3, \sigma)}\, \Lambda(x) = \mathcal{W}_4^T\, \Lambda(x), \quad \mathcal{W}_4 = {\mathbf{\Theta}^{(0.3, \sigma)}}^T\, \mathcal{Y}_0, \\ & {_0^C}\mathsf{D}_x^{0.6} y_1(x) = {_0^C}\mathsf{D}_x^{0.2}(\mathsf{D}_x^{0.4} y_1(x)) \approx \mathcal{Y}_1 ^T\, \Lambda ^{\sigma}(x), \\ & \quad \Longrightarrow \, \, \mathsf{D}_x^{0.4} y_1(x) \approx \mathcal{Y}_1 ^T\, \mathbf{\Theta}^{(0.2, \sigma)}\, \Lambda(x) = \mathcal{Z}_2 ^T\, \Lambda(x), \quad \mathcal{Z}_2 = {\mathbf{\Theta}^{(0.2, \sigma)}}^T\, \mathcal{Y}_1 \end{aligned} \end{equation} (5.17)

    Using (4.12) and (5.16), the coefficients of System II can be approximated as follows:

    \begin{equation} \begin{aligned} & 2 \approx \mathcal{F}_1\, \Lambda ^{\sigma}(x), \quad \exp(x) \approx\mathcal{F}_2\, \Lambda ^{\sigma}(x), \quad x \approx\mathcal{F}_3\, \Lambda ^{\sigma}(x), \\ & 2 \, y_0(x) \approx \mathcal{F}_1 \, \Lambda ^{\sigma}(x) \, {\Lambda ^{\sigma} }^T(x)\, \mathcal{W}_1 \approx \mathcal{F}_1 \, \mathcal{W}_1^*\, \Lambda ^{\sigma}(x), \\ & \exp(x)\, \mathsf{D}_x^{0.4} y_1(x) \approx \mathcal{F}_2 \, \Lambda ^{\sigma}(x) \, {\Lambda ^{\sigma} }^T(x)\, \mathcal{Z}_2 \approx \mathcal{F}_2 \, \mathcal{Z}_2^*\, \Lambda ^{\sigma}(x), \\ & x \, y_1(x) \approx \mathcal{F}_3 \, \Lambda ^{\sigma}(x) \, {\Lambda ^{\sigma} }^T(x)\, \mathcal{Z}_1 \approx \mathcal{F}_3 \, \mathcal{Z}_1^*\, \Lambda ^{\sigma}(x) \end{aligned} \end{equation} (5.18)

    where \mathcal{W}_1^*, \mathcal{Z}_1^* , and \mathcal{Z}_2^* are (N+1) \times (N+1) operational matrices of the product corresponding to the vectors \mathcal{W}_1\mathcal{Z}_1 and \mathcal{Z}_2 , respectively. Now, the kernels and integral parts can be approximated as follows:

    \begin{equation} \begin{aligned} & x \approx {\Lambda ^{\sigma}}^T(x)\, \mathcal{K}_1\, \Lambda ^{\sigma}(t), \quad x-t \approx {\Lambda ^{\sigma}}^T(x)\, \mathcal{K}_2\, \Lambda ^{\sigma}(t), \quad x-2t \approx {\Lambda ^{\sigma}}^T(x)\, \mathcal{K}_3\, \Lambda ^{\sigma}(t), \\ & \int _0^x x\, \mathsf{D}_t^{0.3} y_0(t)\, dt \approx {\Lambda ^{\sigma}}^T(x)\, \mathcal{K}_1\, \mathcal{W}_3^*\, \int _0^x \Lambda ^{\sigma}(t) dt \approx {\Lambda ^{\sigma}}^T(x)\, \mathcal{K}_1\, \mathcal{W}_3^*\, \mathbf{\Theta }^{(\sigma)}\, \Lambda ^{\sigma}(x), \\ & \int _0^x (x-t)\, y_1(t)\, dt \approx {\Lambda ^{\sigma}}^T(x)\, \mathcal{K}_2\, \mathcal{Z}_1^*\, \int _0^x \Lambda ^{\sigma}(t) dt \approx {\Lambda ^{\sigma}}^T(x)\, \mathcal{K}_2\, \mathcal{Z}_1^*\, \mathbf{\Theta} ^{(\sigma)}\, \Lambda ^{\sigma}(x), \\ & \int _0^x (x-2t)\, \mathsf{D}_t^{0.8} y_0(t)\, dt \approx {\Lambda ^{\sigma}}^T(x)\, \mathcal{K}_3\, \mathcal{Y}_0^*\, \int _0^x \Lambda ^{\sigma}(t) dt \approx {\Lambda ^{\sigma}}^T(x)\, \mathcal{K}_3\, \mathcal{Y}_0^*\, \mathbf{\Theta }^{(\sigma)}\, \Lambda ^{\sigma}(x) \end{aligned} \end{equation} (5.19)

    where \mathbf{\Theta }^{(\sigma)} is the operational matrix of the integration in (4.14) and \mathcal{W}_3^*, \mathcal{Z}_1^* , and \mathcal{Y}_0^* are the operational matrices of the product corresponding to the vectors \mathcal{W}_3, \mathcal{Z}_1, \mathcal{Y}_0 . To use the tau method, the sources functions \mathsf{f}_1 and \mathsf{f}_2 should be approximated in terms of the Gegenbauer polynomials:

    \begin{equation} \mathsf{f}_1(x) \approx \mathcal{F}_4^T\, \Lambda ^{\sigma}(x), \quad \mathsf{f}_2(x) \approx \mathcal{F}_5^T\, \Lambda ^{\sigma}(x) \end{equation} (5.20)

    where vectors \mathcal{F}_4 and \mathcal{F}_5 are obtained using Eq (4.9). By substituting approximations (5.15)–(5.20) into system (5.14), the following algebraic system is achieved:

    \begin{equation} \begin{aligned} &\mathcal{R}_1(x) = \mathcal{W}_2 ^T\, \Lambda ^{\sigma}(x)+ \mathcal{W}_4 ^T\, \Lambda ^{\sigma}(x)-\mathcal{F}_1^T \, \mathcal{W}_1^* \, \Lambda ^{\sigma}(x)-{\Lambda ^{\sigma}}^T(x)\, \mathcal{K}_1\, \mathcal{W}_3^*\, \mathbf{\Theta}^{(\sigma)} \, \Lambda ^{\sigma}(x)\\ & \quad -{\Lambda ^{\sigma}}^T(x)\, \mathcal{K}_2\, \mathcal{Z}_1^*\, \mathbf{\Theta}^{(\sigma)} \, \Lambda ^{\sigma}(x)-\mathcal{F}_4^T\, \Lambda ^{\sigma}(x) \approx 0, \\ & \mathcal{R}_2(x) = \mathcal{Y}_1 ^T\, \Lambda ^{\sigma}(x)+ \mathcal{F}_2^T \, \mathcal{Z}_2^* \, \Lambda ^{\sigma}(x)+\mathcal{F}_3^T \, \mathcal{Z}_1^* \, \Lambda ^{\sigma}(x) -{\Lambda ^{\sigma}}^T(x)\, \mathcal{K}_3\, \mathcal{Y}_0^*\, \mathbf{\Theta}^{(\sigma)} \, \Lambda ^{\sigma}(x)-\mathcal{F}_5^T\, \Lambda ^{\sigma}(x) \approx 0 \end{aligned} \end{equation} (5.21)

    The inner product of algebraic system (5.21) by the basis vector \Lambda ^{\sigma}(x) leads to the following system of algebraic equations:

    \begin{equation} \left\{ \begin{array}{l} \mathcal{W}_2 ^T\, \mathcal{Q}+ \mathcal{W}_4 ^T\, \mathcal{Q}-\mathcal{F}_1^T \, \mathcal{W}_1^* \, \mathcal{Q}-\Delta _1-\Delta _2-\mathcal{F}_4^T\, \mathcal{Q} \approx 0, \\ \mathcal{Y}_1 ^T\, \mathcal{Q}+\mathcal{F}_2^T\, \mathcal{Z}_2 ^*\, \mathcal{Q}+\mathcal{F}_3^T\, \mathcal{Z}_1 ^*\, \mathcal{Q}-\Delta _3-\mathcal{F}_5^T\, \mathcal{Q} \approx 0 \end{array} \right. \end{equation} (5.22)

    where \Delta _i, i = 1, 2, 3 , and \mathcal{Q} are (N+1)\times (N+1) matrices in (4.21) and (4.23).

    The systems in Eqs (5.13) and (5.22) involve 2(N+1) algebraic equations in 2(N+1) unknown variables \mathsf{y}_{0, i}, \mathsf{y}_{1, i}, i = 0, 1, \cdots, N . By solving the resulted systems, the values of these variables are estimated and approximate solutions derived from Eqs (5.3) and (5.16).

    In this section, we aim to compute error bounds for the obtained approximate solutions within a Gegenbauer-weighted Sobolev space. To begin, we present the definitions of this space.

    Definition 6.1. Suppose that m \in \mathbb{N}, \mathbf{I} = [0, 1] , then the Gegenbauer-weighted Sobolev space \mathcal{A}_{\omega ^{\sigma}}^m(\mathbf{I}) is defined as below:

    \begin{equation*} \mathcal{A}_{\omega ^{\sigma}}^m(\mathbf{I}) = \{ z\, | \, z \text{ is measurable and } \Vert z \Vert _{k, \omega ^{\sigma}} < \infty, k = 0, 1, \cdots, m \} \end{equation*}

    This space is equipped with the following norm and semi-norm:

    \begin{equation*} \Vert z \Vert _{m, \omega ^{\sigma}} = \bigg ( \sum\limits_{k = 0}^{m} \bigg \Vert \frac{d^k z(x)}{dx^k}\bigg \Vert _{\omega _k^{\sigma}}^2\bigg) ^{\frac{1}{2}}, \quad \vert z \vert _{m, \omega ^{\sigma}} = \bigg \Vert \frac{d^m z(x)}{dx^m}\bigg \Vert _{\omega _m^{\sigma}} \end{equation*}

    where \Vert \, .\, \Vert _{ \omega_k ^{\sigma}} denotes the L^2_{ \omega_k ^{\sigma}} -norm and \omega_k ^{\sigma}(x) = x^{\sigma +k-\frac{1}{2}}(1-x)^{\sigma +k-\frac{1}{2}} .

    Definition 6.2. The following inequality holds for any s \in \mathbb{R} and z \in \mathcal{A}_{\omega ^{\sigma}}^m(\mathbf{I})

    \begin{equation} \Vert z \Vert _{s, \omega ^{\sigma}} \leq \Vert z \Vert _{[s]+1, \omega ^{\sigma}}^{\iota}\, \Vert z \Vert _{[s], \omega ^{\sigma}}^{1-\iota} \end{equation} (6.1)

    where s = [s]+\iota, 0 < \iota < 1 . Inequality (6.1) is known as the Gagliardo-Nirenberg inequality [35].

    Definition 6.3. If \langle., . \rangle _{m, \omega ^{\sigma}} and \Vert. \Vert _{m, \omega ^{\sigma}} are the inner and the norm in the space \mathcal{A}_{\omega ^{\sigma}}^m(\mathbf{I}) , respectively, then the following inequality holds for any two functions \phi, \varphi \in \mathcal{A}_{\omega ^{\sigma}}^m(\mathbf{I}) [36]:

    \begin{equation} \langle \phi, \varphi \rangle _{m, \omega ^{\sigma}}\leq \frac{1}{2} \bigg(\Vert \phi \Vert _{m, \omega ^{\sigma}}^2+\Vert \varphi \Vert _{m, \omega ^{\sigma}}^2\bigg) \end{equation} (6.2)

    Theorem 6.1. Assume that y \in \mathcal{A}_{\omega ^{\sigma}}^m(\mathbf{I}), m \in \mathbb{N}, 0\leq \eta \leq m, and y_N(x) is the Gegenbauer approximation to y(x) . An estimation of the error bound can be obtained as follows:

    \begin{equation} \Vert y-y_N \Vert _{\eta, \omega ^{\sigma}} \leq \mathcal{C}_0\, (N+2\sigma +1)^{\eta -m}(N+2)^{\frac{\eta -m}{2}} \vert y \vert _{m, \omega ^{\sigma}} \end{equation} (6.3)

    where \mathcal{C}_0 is a positive constant.

    Proof. Since y_N(x) is the Gegenbauer approximation to y(x) , subtracting the series in (4.13) from (4.11) and differentiating it leads to

    \begin{equation*} \frac{d^l}{dx^l}(y(x)-y_N(x)) = \frac{d^l}{dx^l} \sum\limits_{k = N+1}^{\infty} g _k^{\sigma}\, \mathcal{G}_k^{\sigma}(x) = \sum\limits_{k = N+1}^{\infty} g _k^{\sigma} \frac{d^l \mathcal{G}_k^{\sigma}(x)}{dx^l}, \quad l\leq m \end{equation*}

    According to the relation between the Gegenbauer and Jacobi polynomials, mentioned in Remark 4.1, one can get

    \begin{equation} \begin{aligned} \frac{d^l \mathcal{G}_k^{\sigma}(x)}{dx^l}& = \frac{\Gamma(k+1)\, \Gamma(\sigma +\frac{1}{2})}{\Gamma(k+\sigma +\frac{1}{2})} \frac{d^l \mathcal{J}_k^{\sigma -\frac{1}{2}, \sigma -\frac{1}{2}}(x)}{dx^l}\\ & = \frac{\Gamma(k+1)\, \Gamma(\sigma +\frac{1}{2})\, \Gamma(l+k+2\sigma)}{\Gamma(k+\sigma +\frac{1}{2})\, \Gamma(k+2\sigma)}\mathcal{J}_{k-l}^{l+\sigma -\frac{1}{2}, l+\sigma -\frac{1}{2}}(x) \end{aligned} \end{equation} (6.4)

    So, from (6.4) and (4.6), one gets

    \begin{equation} \bigg \Vert \frac{d^l}{dx^l}(y(x)-y_N(x)) \bigg \Vert _{\omega _l^{\sigma}}^2 = \sum\limits_{k = N+1}^{\infty} {g _k^{\sigma}}^2 \, \frac{\Gamma ^2(k+1)\, \Gamma ^2(\sigma +\frac{1}{2})\, \Gamma ^2 (k+l+2\sigma)}{2\, \Gamma ^2(k+2\sigma)\, \Gamma(l+2\sigma)\, \Gamma(k-l+1)\, (k+\sigma)} \end{equation} (6.5)

    Similarly, one can obtain

    \begin{equation} \bigg \Vert \frac{d^m y}{dx^m} \bigg \Vert _{\omega _m^{\sigma}}^2 = \sum\limits_{k = m+1}^{\infty} {g _k^{\sigma}}^2 \, \frac{\Gamma ^2(k+1)\, \Gamma ^2(\sigma +\frac{1}{2})\, \Gamma ^2 (k+m+2\sigma)}{2\, \Gamma ^2(k+2\sigma)\, \Gamma(m+2\sigma)\, \Gamma(k-m+1)\, (k+\sigma)} \end{equation} (6.6)

    with the aid of the Stirling formula [35], the following inequality is obtained:

    \begin{equation} \frac{\Gamma ^2(k+l+2\sigma)\, \Gamma(m+2\sigma)\, \Gamma(k-m+1)}{\Gamma ^2(k+m+2\sigma)\, \Gamma(l+2\sigma)\, \Gamma(k-l+1)}\leq \mathcal{C}_1\, (2\sigma)^{m-l}\, (k+2\sigma)^{2(l-m)}\, (k+1)^{l-m} \end{equation} (6.7)

    where \mathcal{C}_1 is a positive constant. Based on the definition of the semi-norm and using (6.5)–(6.7), one gets

    \begin{align*} \vert y-y_N \vert _{l, \omega ^{\sigma}}^2& = \bigg\Vert \frac{d^l}{dx^l}(y-y_N) \bigg \Vert _{\omega _l^{\sigma}}^2\\ & = \sum\limits _{k = N+1}^{\infty} \bigg\{\frac{2 {g_k^{\sigma}}^2\Gamma^2(k+1)\Gamma^2(\sigma +\frac{1}{2})\Gamma^2(k+l+2\sigma)\Gamma^2(k+2\sigma)\Gamma(m+2\sigma)\Gamma(k-m+1)(k+\sigma)}{2{g_k^{\sigma}}^2\Gamma(l+2\sigma)\Gamma(k-l+1)\Gamma^2(k+1)\Gamma^2(\sigma+\frac{1}{2})\Gamma^2(k+m+2\sigma)(k+\sigma)}\\ & \quad \quad \times {g_k^{\sigma}}^2\frac{\Gamma^2(k+1)\Gamma^2(\sigma+\frac{1}{2})\Gamma^2(k+m+2\sigma)}{2\Gamma^2(k+2\sigma)\Gamma(m+2\sigma)\Gamma(k-m+1)(k+\sigma)}\bigg\}\\ & \leq \sum\limits_{k = N+1}^{\infty} \mathcal{C}_1(2\sigma)^{m-l} {g_k^{\sigma}}^2 (k+2\sigma)^{2(l-m)} (k+1)^{l-m}\frac{\Gamma^2(k+1) \Gamma^2(\sigma +\frac{1}{2})\Gamma^2(k+m+2\sigma)}{2\Gamma^2(k+2\sigma)\Gamma(m+2\sigma)\Gamma(k-m+1)(k+\sigma)}\\ & \leq \sum\limits_{k = m+1}^{\infty} \mathcal{C}_1(2\sigma)^{m-l} {g_k^{\sigma}}^2 (N+2\sigma +1)^{2(l-m)} (N+2)^{l-m}\\ & \quad \quad \times \frac{\Gamma^2(k+1) \Gamma^2(\sigma +\frac{1}{2})\Gamma ^2(k+m+2\sigma)}{2\Gamma^2(k+2\sigma)\Gamma(m+2\sigma)\Gamma(k-m+1)(k+\sigma)}\\ & \leq \mathcal{C}_1(2\sigma)^{m-l} {g_k^{\sigma}}^2 (N+2\sigma +1)^{2(l-m)} (N+2)^{l-m} \vert y \vert _{m, \omega ^{\sigma}}^2 \end{align*}

    then, the following inequality is acquired

    \begin{equation} \Vert y-y_N \Vert _{m, \omega ^{\sigma}} \leq \mathcal{C}_2 (2\sigma)^{\frac{m-l}{2}}(N+2\sigma +1)^{l-m}(N+2)^{\frac{l-m}{2}} \vert y \vert _{m, \omega ^{\sigma}}, \quad l \leq m \end{equation} (6.8)

    Using the Gagliardo–Nirenberg inequality leads to the desired bound for \eta = [\eta]+\eta _0, 0 < \eta _0 < 1 :

    \begin{align*} \Vert y-y_N \Vert _{\eta, \omega ^{\sigma}} &\leq \Vert y-y_N \Vert _{[\eta]+1, \omega ^{\sigma}}^{\eta _0} \Vert y-y_N \Vert _{[\eta], \omega ^{\sigma}}^{1-\eta _0} \\ & \leq \mathcal{C}_0 (N+2\sigma +1)^{\eta -m}(N+2)^{\frac{\eta -m}{2}} \vert y \vert _{m, \omega ^{\sigma}} \end{align*}

    Corollary 6.2. Using Theorem 6.1, an error bound for \frac{dy}{dx}-\frac{dy_N}{dx} can be estimated as follows:

    \begin{align*} \frac{d^l}{dx^l}\bigg(\frac{dy(x)}{dx}-\frac{dy_N(x)}{dx}\bigg)& = \frac{d^{l+1}}{dx^{l+1}}(y(x)-y_N(x)) = \sum\limits_{k = N+1}^{\infty} g_{k}^{\sigma}\, \frac{d^{l+1}\mathcal{G}_k^{\sigma}(x)}{dx^{l+1}}\\ & = \sum\limits_{k = N+1}^{\infty} g_{k}^{\sigma} \frac{\Gamma(k+1)\, \Gamma(\sigma +\frac{1}{2})}{\Gamma(k+\sigma +\frac{1}{2})} \frac{d^{l+1}}{dx^{l+1}}\mathcal{J}_{k}^{\sigma -\frac{1}{2}, \sigma -\frac{1}{2}}(x)\\ & = \sum\limits_{k = N+1}^{\infty} g_{k}^{\sigma} \frac{\Gamma(k+1)\, \Gamma(\sigma +\frac{1}{2})\, \Gamma(l+k+2\sigma +1)}{\Gamma(k+\sigma +\frac{1}{2})\, \Gamma(k+2\sigma)} \mathcal{J}_{k-l-1}^{\sigma +l+\frac{1}{2}, \sigma +l+\frac{1}{2}}(x) \end{align*}

    Therefore, one has

    \begin{equation} \bigg \Vert \frac{dy}{dx}- \frac{dy_N}{dx}\bigg\Vert _{\eta, \omega _{1}} \leq \mathcal{C}_1\, (N+2\sigma +2)^{\frac{\eta -m}{2}}\, (N+2\sigma +4)^{2(\eta -m)}\, (N+1)^{\frac{\eta -m}{2}}\, \vert y \vert _{m, \omega ^{\sigma}} \quad l\leq m, \end{equation} (6.9)

    where \mathcal{C}_1 is a positive constant.

    Theorem 6.3. If y\in \mathcal{A}_{\omega ^{\sigma}}^m(\mathbf{I}), m \in \mathbb{N}, 0\leq \eta \leq m, 0 < \mu < 1, and y_N(x) is the Gegenbauer approximation to y(x) , then an error bound can be estimated for {_0^C}\mathsf{D}_x^{\mu} y(x)-{_0^C}\mathsf{D}_x^{\mu} y_N(x) as follows:

    \begin{equation} \bigg \Vert {_0^C}\mathsf{D}_x^{\mu} y-{_0^C}\mathsf{D}_x^{\mu} y_N \bigg \Vert _{\eta, \omega _1^{\sigma}} \leq \delta _0\, \mathcal{C}_1 \frac{\Gamma(\sigma -\mu +\frac{1}{2})\, \Gamma(\sigma +\frac{1}{2})}{\Gamma(1-\mu)\, \Gamma(2\sigma -\mu +1)}(N+2\sigma +2)^{\frac{\eta -m}{2}} (N+2\sigma +4)^{2(\eta -m)} (N+1)^{\frac{\eta -m}{2}} \vert y \vert _{m, \omega ^{\sigma}} \end{equation} (6.10)

    Proof. By noting the relation between the classical derivative and Riemann-Liouville integral operators (the second property in Definition 2.2), one has

    \begin{equation} \begin{aligned} {_0^C}\mathsf{D}_x^{\mu} y(x)-{_0^C}\mathsf{D}_x^{\mu} y_N(x)& = {_0^{\mathsf{RL}}}\mathsf{I}_x^{1-\mu}(y^{\prime}(x)-y_N^{\prime}(x))\\ & = \frac{1}{\Gamma(1-\mu)}\int _0^x (x-\xi)^{-\mu} (y^{\prime}(\xi)-y_N^{\prime}(\xi))\, d\xi \\ & = \frac{1}{\Gamma(1-\mu)} \bigg\{ x^{-\mu} * (y^{\prime}(x) -y^{\prime}_N(x)) \bigg\} \end{aligned} \end{equation} (6.11)

    where the star sign denotes the convolution of x^{-\mu} and (y^{\prime}(x) -y^{\prime}_N(x)) . Applying Young's convolution inequality \Vert u_1 *u_2 \Vert _q \leq \delta _0\, \Vert u_1 \Vert _1 \, \Vert u_2 \Vert _q and Corollary 6.2 to (6.11) leads to the desired result:

    \begin{align*} \Vert {_0^C}\mathsf{D}_x^{\mu} y-{_0^C}\mathsf{D}_x^{\mu} y_N \Vert _{\eta, \omega _1^{\sigma}}& \leq \frac{\delta _0}{\Gamma(1-\mu)} \Vert x^{-\mu} \Vert _{L^1_{\omega ^{\sigma}}(\mathbf{I})} \Vert y^{\prime}-y_N^{\prime} \Vert _{\eta, \omega _1^{\sigma}} \\ & = \frac{\delta _0}{\Gamma(1-\mu)}\frac{\Gamma(\sigma -\mu +\frac{1}{2})\, \Gamma(\sigma +\frac{1}{2})}{\Gamma(2\sigma -\mu +1)}\Vert y^{\prime} -y_N^{\prime} \Vert _{\eta, \omega _1^{\sigma}}\\ & \leq \delta _0\, \mathcal{C}_1 \frac{\Gamma(\sigma -\mu +\frac{1}{2})\, \Gamma(\sigma +\frac{1}{2})}{\Gamma(1-\mu)\, \Gamma(2\sigma -\mu +1)}(N+2\sigma +2)^{\frac{\eta -m}{2}} (N+2\sigma +4)^{2(\eta -m)}(N+1)^{\frac{\eta -m}{2}} \vert y \vert_{m, \omega ^{\sigma}} \end{align*}

    Now, suppose that y_N(x) is the obtained solutions from the suggested method, so one has the following approximate system:

    \begin{equation} \begin{aligned} {_0^C}\mathsf{D}_x^{\mu _{r, n}}y_{r, N}(x)&+\sum\limits _{k = 1}^{n-1} \xi _{r, k}(x)\, {_0^C}\mathsf{D}_x^{\mu _{r, n-k}}y_{r, N}(x)+\xi _{r, n}y_{r, N}(x)\\ &-\sum \limits _{j = 1}^{2}\nu _{r, j} \int _0^x \kappa _{r, j}(x, t)\, {_0^C}\mathsf{D}_t^{\gamma _{r, 2-j}}\, y_{j, N}(t)\, dt -\mathsf{f}_r(x)+\mathcal{R}_r(x) = 0 \end{aligned} \end{equation} (6.12)

    where \mathcal{R}_r(x), r = 1, 2 are the residual functions. By subtracting Eq (6.12) from Eq (1.1), the following error system is achieved:

    \begin{equation} \begin{aligned} \mathcal{R}_r(x) = {_0^C}\mathsf{D}_x^{\mu _{r, n}}e_{r, N}(x)&+\sum\limits _{k = 1}^{n-1} \xi _{r, k}(x)\, {_0^C}\mathsf{D}_x^{\mu _{r, n-k}}e_{r, N}(x)+\xi _{r, n}e_{r, N}(x)\\ &- \sum \limits _{j = 1}^{2}\nu _{r, j} \int _0^x \kappa _{r, j}(x, t)\, {_0^C}\mathsf{D}_t^{\gamma _{r, 2-j}}\, e_{j, N}(t)\, dt \end{aligned} \end{equation} (6.13)

    where e_{j, N} = y_{j}(x)-y_{j, N}(x) . To obtain an error bound for the residual function in (6.13), this function is multiplied by e_{r, N} ; therefore, using the inequality in (6.2) yields:

    \begin{align*} \langle \mathcal{R}_r(x), e_{r, N}(x) \rangle _{\eta, \omega ^{\sigma}} & = \langle {_0^C}\mathsf{D}_x^{\mu _{r, n}}e_{r, N}(x), e_{r, N}(x) \rangle _{\eta, \omega ^{\sigma}}+\langle\sum\limits _{k = 1}^{n-1} \xi _{r, k}(x)\, {_0^C}\mathsf{D}_x^{\mu _{r, n-k}}e_{r, N}(x), e_{r, N}(x)\rangle _{\eta, \omega ^{\sigma}}\\ &+\langle \xi _{r, n}e_{r, N}(x), e_{r, N}(x) \rangle _{\eta, \omega ^{\sigma}}-\langle \sum \limits _{j = 1}^{2}\nu _{r, j} \int _0^x \kappa _{r, j}(x, t)\, {_0^C}\mathsf{D}_t^{\gamma _{r, 2-j}}\, e_{j, N}(t)\, dt, e_{r, N}(x) \rangle _{\eta, \omega ^{\sigma}}\\ & \leq \frac{1}{2} \Vert {_0^C}\mathsf{D}_x^{\mu _{r, n}}e_{r, N} \Vert _{\eta, \omega _1^{\sigma}}^2 +\frac{1}{2} \Vert e_{r, N} \Vert _{\eta, \omega ^{\sigma}}^2+\frac{1}{2} \sum\limits_{k = 1}^{n-1}\Vert \xi _{r, k}\Vert _{L^2_{\omega^{\sigma}}(\mathbf{I})}^2 \, \Vert {_0^C}\mathsf{D}_x^{\mu _{r, n-k}}e_{r, N} \Vert _{\eta, \omega _1^{\sigma}}^2\\ &\, \, +\frac{1}{2} \Vert e_{r, N} \Vert _{\eta, \omega ^{\sigma}}^2+\frac{1}{2} \Vert \xi _{r, N}\Vert _{L^2_{\omega^{\sigma}}(\mathbf{I})}^2 \, \Vert e_{r, N} \Vert _{\eta, \omega ^{\sigma}}^2+\frac{1}{2} \Vert e_{r, N} \Vert _{\eta, \omega ^{\sigma}}^2 \\ & \, \, +\frac{1}{2} \sum\limits_{j = 1}^2 \vert \nu _{r, j} \vert ^2 \, \Vert \kappa_{r, j} \Vert _{L^2_{ \omega_0^{\sigma} }(\mathbf{I}^2)}^2\, \Vert {_0^C}\mathsf{D}_x^{\gamma _{r, 2-j}}e_{j, N} \Vert _{\eta, \omega _1^{\sigma}}^2+\frac{1}{2} \Vert e_{r, N} \Vert _{\eta, \omega ^{\sigma}}^2 \end{align*}

    where \omega_0^{\sigma}(x, t) = \omega ^{\sigma}(x)\, \omega ^{\sigma}(t) and \mathbf{I}^2 = \mathbf{I} \times \mathbf{I} . So, one has

    \begin{equation} \begin{aligned} \Vert \mathcal{R}_r \Vert _{\eta, \omega ^{\sigma}}^2 &\leq \Vert {_0^C}\mathsf{D}_x^{\mu _{r, n}}e_{r, N} \Vert _{\eta, \omega _1^{\sigma}}^2+\sum\limits_{k = 1}^{n-1}\Vert \xi _{r, k}\Vert _{L^2_{\omega^{\sigma}}(\mathbf{I})}^2 \, \Vert {_0^C}\mathsf{D}_x^{\mu _{r, n-k}}e_{r, N} \Vert _{\eta, \omega _1^{\sigma}}^2+3 \Vert e_{r, N} \Vert _{\eta, \omega ^{\sigma}}^2\\ & \, \, +\sum\limits_{j = 1}^2 \vert \nu _{r, j} \vert ^2 \, \Vert \kappa_{r, j} \Vert _{L^2_{ \omega_0^{\sigma} }(\mathbf{I}^2)}^2\, \Vert {_0^C}\mathsf{D}_x^{\gamma _{r, 2-j}}e_{j, N} \Vert _{\eta, \omega _1^{\sigma}}^2+\Vert \xi _{r, N}\Vert _{L^2_{\omega^{\sigma}}(\mathbf{I})}^2 \, \Vert e_{r, N} \Vert _{\eta, \omega ^{\sigma}}^2\\ & \leq \delta _0^2\, \mathcal{C}_1^2\, \frac{\Gamma ^2(\sigma -\mu _{r, n}+\frac{1}{2})\, \Gamma ^2(\sigma +\frac{1}{2})}{\Gamma ^2(1-\mu _{r, n})\, \Gamma ^2(2\sigma -\mu _{r, n}+1)}\\ & \quad \quad \times (N+2\sigma +2)^{\eta -m}(N+2\sigma +4)^{4(\eta -m)}(N+1)^{\eta -m}\vert y \vert _{m, \omega ^{\sigma}}^2\\ & + \delta _0^2\, \mathcal{C}_1^2\, \sum\limits _{k = 1}^{n-1}\Vert \xi _{r, k} \Vert _{L^2_{\omega ^{\sigma}}(\mathbf{I})} ^2\frac{\Gamma^2(\sigma -\mu _{r, n-k}+\frac{1}{2})\, \Gamma^2(\sigma +\frac{1}{2})}{\Gamma^2(1-\mu _{r, n-k})\, \Gamma^2(2\sigma -\mu _{r, n}+1)}\\ & \quad \quad \times (N+2\sigma +2)^{\eta -m}\, (N+2\sigma +4)^{4(\eta -m)} (N+1)^{\eta -m}\, \vert y \vert_{m, \omega ^{\sigma}}^2\\ & +3 \mathcal{C}_1^2\, (N+2\sigma +1)^{2(\eta-m)} (N+2)^{\eta -m}\vert y \vert _{m, \omega ^{\sigma}}^2\\ & +\delta _0^2\, \mathcal{C}_1^2\, \sum\limits _{j = 1}^{2} \vert \nu _{r, j} \vert ^2 \, \Vert \kappa_{r, j} \Vert _{L^2_{\omega _0^{\sigma}}(\mathbf{I}^2)}^2 \frac{\Gamma ^2(\sigma -\gamma_{r, 2-j}+\frac{1}{2})\, \Gamma^2(\sigma +\frac{1}{2})}{\Gamma^2(1-\gamma_{r, 2-j})\, \Gamma^2(2\sigma -\gamma _{r, 2-j})}\\ & \quad \quad \times (N+2\sigma +2)^{\eta -m}(N+2\sigma +4)^{4(\eta -m)} (N+1)^{\eta -m} \vert y \vert _{m, \omega^{\sigma}}\\ & +\mathcal{C}_1^2\, \Vert \mathcal{\xi}_{r, N} \Vert _{L^2_{\omega ^{\sigma}}(\mathbf{I})}^2(N+2\sigma +1)^{2(\eta -m)}(N+2)^{\eta -m} \vert y \vert _{m, \omega ^{\sigma}}^2, \quad r = 1, 2 \end{aligned} \end{equation} (6.14)

    From the righthand side of (6.14), increasing the value of N leads to a smaller error bound for the residual function.

    In Section 5, we implemented the proposed approach on two systems of Volterra-type FIDEs. Systems I and II were transformed into corresponding systems of algebraic equations, as shown in Eqs (5.13) and (5.22). We solved these systems for various values of N and \sigma and calculated the maximum absolute errors (MAEs) and least square errors (LSEs). The obtained results are presented in the forms of figures and tables. All computations were performed using Maple 16 software. Also, the convergence rate C_R(y_j), j = 0, 1 is computed by the following formula:

    \begin{equation*} C_R(y_j) = \frac{|\ln(e_{i+1}^j/e_i^j)|}{\ln(i+1/i)}, \quad e_i^j = |y_j-y_{j, N}|, \quad j = 0, 1, \quad i = 1, 2, \ldots, 8 \end{equation*}

    Example 1. Consider system (5.1) with the exact solutions (y_0(x), y_1(x)) = (-x^2, 1-x^3) , initial conditions (y_0(0), y_1(0)) = (0, 1) , and source functions

    \begin{align*} & \mathsf{f}_1(x) = -\frac{2}{\Gamma(2.1)}x^{1.1}-\frac{2}{\Gamma(2.4)}x^{1.4}-2x^3+\frac{10}{\Gamma(3.7)} \exp(x)\, x^{2.7}, \\ & \mathsf{f}_2(x) = -\frac{6}{\Gamma(3.2)}x^{2.2}-\frac{6}{\Gamma(3.5)}x^{3.5}+\frac{2}{\Gamma(3.2)}\sin(x) \, x^{2.2}-\frac{2}{3.2 \Gamma(2.2)} x^{3.2} \end{align*}

    Table 1 displays the values of approximate solutions at equally spaced points x_i = 0.2 i, i = 0, 1, \cdots, 5 as well as the corresponding least square errors obtained from the proposed tau-Gegenbauer method (for N = 7 , \sigma = 1 ) and the block-by-block approach combined with a finite difference approximation [37] (for N = 10 , h = 0.1 ). It is evident that the tau-Gegenbauer method yields results with significantly smaller errors compared to those reported in [37]. MAEs for various values of N (ranging from 2 to 7) and \sigma = 1 are presented in Table 2. As expected, increasing the number of Gegenbauer polynomials in the series solutions (represented by N ) leads to a reduction in errors. Table 4 reports the MAEs of the approximate solutions for N = 5 and different values of \sigma .

    Table 1.  Values of exact and approximate solutions at selected points for N = 7 and \sigma = 1 for Example 1.
    Exact values Proposed scheme Method in [37]
    x_i {y_0}_{Exact} {y_1}_{Exact} y_{0, N} y_{1, N} y_{0, N} y_{1, N}
    0.0 0.0000 1.0000 3.2144\times10^{-5} 0.9999946673 0.0000 1.0000
    0.2 -0.0400 0.9920 -0.0399856254 0.9919993594 -0.04387683293 0.99071904024
    0.4 -0.1600 0.9360 -0.1599851160 0.9360018667 -0.1667659284 0.93075651777
    0.6 -0.3600 0.7840 -0.3599697321 0.7840028442 -0.36880182427 0.77152149574
    0.8 -0.6400 0.4880 -0.6399544939 0.4880059282 -0.64932899451 0.46536622778
    1.0 -1.0000 0.0000 -0.9999322792 1.2027 \times 10^{-5} -1.00653583547 -0.0349831482
    LSE 3.7817\times 10^{-5} 5.0342\times 10^{-6} 3.3483\times 10^{-4} 3.1115\times 10^{-3}

     | Show Table
    DownLoad: CSV
    Table 2.  MAEs and convergence rate of approximate solutions for various values of N and \sigma = 1 of Example 1.
    N MAE (y_0) C_R(y_0) MAE (y_1) C_R(y_1)
    2 2.2605 \times 10^{-2} -- 6.6949\times 10^{-2} --
    3 2.3660 \times 10^{-3} 5.5664 4.3205\times 10^{-4} 12.4379
    4 9.4636 \times 10^{-4} 3.1852 9.8083\times 10^{-5} 5.1540
    5 2.8943 \times 10^{-4} 5.3092 4.9989\times 10^{-5} 3.0205
    6 1.9745 \times 10^{-4} 2.0975 1.3829\times 10^{-5} 7.0482
    7 7.7834 \times 10^{-5} 6.0389 1.2027\times 10^{-5} 0.9057
    8 4.2115 \times 10^{-5} 4.5995 8.6863\times 10^{-6} 2.4369

     | Show Table
    DownLoad: CSV
    Table 3.  Converegence rates at x = 0.5, 0.7 for various values of N and \sigma = 1 of Example 1.
    x=0.5 x=0.7
    N Error (y_0) C_R(y_0) Error (y_1) C_R(y_1) Error (y_0) C_R(y_0) Error (y_1) C_R(y_1)
    2 4.9384 \times 10^{-3} -- 1.8672\times 10^{-4} -- 9.6234 \times 10^{-3} -- 1.6530 \times 10^{-2} --
    4 2.2761 \times 10^{-4} 4.4394 6.3837\times 10^{-7} 8.1923 3.3882 \times 10^{-4} 4.8280 3.8006 \times 10^{-5} 8.7646
    6 3.5898 \times 10^{-5} 4.5551 5.0491\times 10^{-6} 5.1004 7.2318 \times 10^{-5} 3.8089 6.4580 \times 10^{-6} 4.3713
    8 4.6590 \times 10^{-5} 0.9062 4.3682\times 10^{-8} 16.5114 5.4926 \times 10^{-5} 0.9562 1.4408 \times 10^{-6} 5.2145
    10 3.7557 \times 10^{-4} 9.3530 1.9654\times 10^{-5} 27.3774 7.2617 \times 10^{-5} 1.2513 2.9547 \times 10^{-5} 13.5374

     | Show Table
    DownLoad: CSV
    Table 4.  LSEs of approximate solutions for various values of \sigma and N = 5 in Example 1.
    \sigma 0.5 1 1.25 1.5 2 2.5
    LSE (y_0) 2.6457 \times 10^{-5} 1.4897 \times 10^{-4} 2.3482\times 10^{-4} 3.2302 \times 10^{-4} 4.9850\times 10^{-4} 6.6611\times 10^{-4}
    LSE (y_1) 8.6608 \times 10^{-5} 2.0616 \times 10^{-5} 2.8292\times 10^{-5} 3.6091 \times 10^{-5} 5.1284\times 10^{-5} 6.5460\times 10^{-5}

     | Show Table
    DownLoad: CSV

    In Figure 1, we present the exact and approximate solutions, as well as the corresponding absolute error functions, for N = 7 and \sigma = 1 . The plots demonstrate a good agreement between the approximate solutions and the exact ones. Furthermore, Figure 2 illustrates the absolute error functions for various values of \sigma while keeping N = 5 constant. These plots further highlight the accuracy of the proposed method. The results substantiate the effectiveness and reliability of the tau-Gegenbauer method in providing accurate approximate solutions. The convergence rate of approximate solutions have been computed and listed in Table 2 for different values of N . In adition, the convergence rates at x = 0.5, 0.7 are calculated for different values of N and \sigma = 1 , which can be seen in Table 3.

    Figure 1.  Plots of (a) Exact solution y_0(x) , (b) Approximate solution y_{0, N}(x) , (c) Exact solution y_1(x) , (d) Approximate solution y_{1, N}(x) , (e) Absolute error function \vert y_0(x)-y_{0, N}(x) \vert , (f) Absolute error function \vert y_1(x)-y_{1, N}(x) \vert , for N = 7 and \sigma = 1 for Example 1.
    Figure 2.  Plots of absolute error functions for N = 5 and \sigma = 0.5, 1.5, 2.5 for Example 1.

    Example 2. Consider system (5.14) as the second example with the exact solutions (y_0(x), y_1(x)) = (x^4, x(1-x)) , initial conditions (y_0(0), y_1(0)) = (0, 0) , and source functions

    \begin{align*} & \mathsf{f}_1(x) = \frac{\Gamma(5)}{\Gamma(4.3)}x^{3.3}+\frac{\Gamma(5)}{\Gamma(3.5)}x^{4.5}-\frac{23}{12}x^4-\frac{1}{6} x^{3}-\frac{\Gamma(5)}{\Gamma(5.7)}x^{5.7}, \\ & \mathsf{f}_2(x) = \frac{1}{\Gamma(1.4)}x^{0.4}-\frac{2}{\Gamma(2.4)}x^{1.4}+\frac{1}{\Gamma(1.6)}\, x^{0.6}\exp(x)-\frac{2}{ \Gamma(2.6)} x^{1.6}\exp(x)\\ & \quad +x^2-x^3+\bigg ( \frac{2\Gamma(5)}{5.2 \Gamma(4.2)}- \frac{5}{\Gamma(5.2)}\bigg) x^{5.2} \end{align*}

    A comparison between exact and approximate solutions at equally spaced points x_i = 0.2 i, i = 0, 1, \cdots, 5 is seen in Table 5 for N = 7 , \sigma = 1 . The LSEs of the results obtained from the proposed method are smaller than those obtained from the block-by-block approach in [37] (for N = 10 , h = 0.1 ). MAEs have been computed for N = 2, 3, \cdots, 8 and \sigma = 1 and listed in Table 6. In Table 7, MAEs of the approximate solutions are computed for N = 5 and diverse values of \sigma . Figures of exact and approximate solutions and absolute error functions are plotted in Figure 3 for N = 7 and \sigma = 1 . Plots of absolute error functions are seen in Figure 4 for various values of \sigma and N = 5 . The convergence rate of approximate solutions have been computed and listed in Table 6 for different values of N . Error bounds of y_{0, N}(x) and y_{1, N}(x) in (6.3) are 8.2270\times 10^{-3} and 2.0428\times 10^{-3} for N = 7, m = 2, \eta = 1, \sigma = 1 , respectively. Based on the convergenc rate in Table 6, the absolute error of y_{0, 7}(x) is proportional to \frac{1}{N^{2.1527}} = 1.5162\times 10^{-2} and the absolute error of y_{1, 7}(x) is proportional to \frac{1}{N^{1.7767}} = 3.1515\times 10^{-2} .

    Table 5.  Values of exact and approximate solutions at selected points for N = 7 and \sigma = 1 for Example 2.
    Exact Proposed scheme Method in [37]
    x_i {y_0}_{Exact} {y_1}_{Exact} y_{0, N} y_{1, N} y_{0, N} y_{1, N}
    0.0 0.0000 0.0000 -2.3182\times10^{-6} 0.0018380780 0.0000 0.0000
    0.2 0.0016 0.1600 0.0016007312 0.1603567253 0.0017309978 0.1600362527
    0.4 0.0256 0.2400 0.0256076988 0.2400674525 0.0271385452 0.2390378788
    0.6 0.1296 0.2400 0.1296217631 0.2402228353 0.1359682357 0.2321624668
    0.8 0.4096 0.1600 0.4096477585 0.1599868537 0.4262711145 0.1302339480
    1.0 1.0000 0.0000 1.0000904470 -0.0009318649 1.0325094318 -0.0805029267
    LSE 3.8913\times 10^{-5} 6.4800\times 10^{-4} 7.1677\times 10^{-3} 1.0402\times 10^{-2}

     | Show Table
    DownLoad: CSV
    Table 6.  MAEs of approximate solutions for various values of N and \sigma = 1 of Example 2.
    N MAE (y_0) C_R(y_0) MAE (y_1) C_R(y_1)
    2 1.3064 \times 10^{-1} -- 4.6931\times 10^{-3} --
    3 1.9922 \times 10^{-2} 4.6382 9.3530\times 10^{-3} 1.7008
    4 2.0981 \times 10^{-4} 15.8278 4.8217\times 10^{-3} 2.3031
    5 1.9955 \times 10^{-4} 0.2247 3.3191\times 10^{-3} 1.6735
    6 1.2604 \times 10^{-4} 2.5201 2.4172\times 10^{-3} 1.7391
    7 9.0447 \times 10^{-5} 2.1527 1.8381\times 10^{-3} 1.7767
    8 6.5017 \times 10^{-5} 2.4742 8.4361\times 10^{-4} 5.8323

     | Show Table
    DownLoad: CSV
    Table 7.  LSEs of approximate solutions for various values of \sigma and N = 5 in Example 2.
    \sigma 0.5 1 1.25 1.5 2 2.5
    LSE (y_0) 4.5374 \times 10^{-6} 8.3138 \times 10^{-4} 1.2031\times 10^{-4} 1.5449 \times 10^{-4} 2.1416\times 10^{-4} 2.6366\times 10^{-4}
    LSE (y_1) 8.1582 \times 10^{-4} 1.1886 \times 10^{-3} 1.4490\times 10^{-3} 1.7085 \times 10^{-3} 2.1904\times 10^{-3} 2.6115\times 10^{-3}

     | Show Table
    DownLoad: CSV
    Figure 3.  Plots of (a) Exact solution y_0(x) , (b) Approximate solution y_{0, N}(x) , (c) Exact solution y_1(x) , (d) Approximate solution y_{1, N}(x) , (e) Absolute error function y_{0, N}(x) , (f) Absolute error function y_{1, N}(x) , for N = 7 and \sigma = 1 for Example 2.
    Figure 4.  Plots of absolute error functions for N = 5 and \sigma = 0.5, 1.25, 1.5 for Example 2.

    A matrix-based tau spectral method utilizing Gegenbauer polynomials was developed to numerically solve a specific class of FIDE systems. A fractional-order integral operational matrix was constructed for this purpose. In [22], differential equations with time-fractional delays were analyzed using spectral operational matrices. Operational matrices of the fractional-order derivative related to shifted Gegenbauer polynomials were derived. To solve fractional differential equations based on the Gegenbauer-Humbert polynomials, a collocation wavelet method was proposed in [24]. The operational matrices of the fractional derivatives were derived. Additional to the main equation, derivative operational matrices were also used to approximate initial and boundary conditions. In contrast, the authors of the current paper used integral operational matrices without requiring conditions to be approximated. The authors have observed in previous works that operational matrices reduce error. The derived error bound for the residual function, evaluated within a Gegenbauer-weighted Sobolev space, demonstrates that selecting a sufficiently large value for the parameter N leads to a sufficiently small error. Notably, the Gegenbauer polynomials were dependent on the parameter \sigma , and altering its value yields different versions of these polynomials. The impact of varying \sigma can be observed in Tables 4 and 7, as well as Figures 2 and 4. The most favorable outcomes were obtained for \sigma = 0.5 and 1 . A comparison between the results obtained from the proposed method and those reported in [37] (employing a block-by-block approach combined with a finite difference method) was presented in Tables 1 and 5. The numerical results obtained through the tau-Gegenbauer scheme exhibited better agreement with the exact solutions. The authors intend to apply the proposed approach to systems of fractional integral equations with variable orders and generalize it to solve two-dimensional integro-partial differential equations of the fractional order.

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

    The authors are thankful to dear referees for the valuable comments which have improved the final version of this work. The authors extend their appreciation to the Deputyship for Research and Innovation, Ministry of Education in Saudi Arabia for funding this research work through project number: 445-9-676. Furthermore, the authors would like to extend their appreciation to Taibah University for its supervision support.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.



    [1] L. Bieberbach, Über dié koeffizienten derjenigen Potenzreihen welche eine schlichte Abbildung des Einheitskreises vermitteln, S.-B. Preuss. Akad. Wiss., 38 (1916), 940–955.
    [2] K. Löwner, Untersuchungen über schlichte konforme Abbildungen des Einheitskreises. I, Math. Ann., 89 (1923), 103–121. https://doi.org/10.1007/BF01448091 doi: 10.1007/BF01448091
    [3] P. R. Garabedian, M. Schiffer, A proof of the Bieberbach conjecture for the fourth coefficient, Journal of Rational Mechanics and Analysis, 4 (1955), 427–465.
    [4] R. Pederson, M. Schiffer, A proof of the Bieberbach conjecture for the fifth coefficient, Arch. Rational Mech. Anal., 45 (1972), 161–193. https://doi.org/10.1007/BF00281531 doi: 10.1007/BF00281531
    [5] R. N. Pederson, A proof of the Bieberbach conjecture for the sixth coefficient, Arch. Rational Mech. Anal., 31 (1968), 331–351. https://doi.org/10.1007/BF00251415 doi: 10.1007/BF00251415
    [6] L. Branges, A proof of the Bieberbach conjecture, Acta Math., 154 (1985), 137–152. https://doi.org/10.1007/BF02392821 doi: 10.1007/BF02392821
    [7] J. E. Brown, A. Tsao, On the Zalcman conjecture for starlike and typically real functions, Math. Z., 191 (1986), 467–474. https://doi.org/10.1007/BF01162720 doi: 10.1007/BF01162720
    [8] L. Li, S. Ponnusamy, J. Qiao, Generalized Zalcman conjecture for convex functions of order \alpha, Acta Math. Hungar., 150 (2016), 234–246. https://doi.org/10.1007/s10474-016-0639-5 doi: 10.1007/s10474-016-0639-5
    [9] W. C. Ma, The Zalcman conjecture for close-to-convex functions, Proc. Amer. Math. Soc., 104 (1988), 741–744. https://doi.org/10.1090/S0002-9939-1988-0964850-X doi: 10.1090/S0002-9939-1988-0964850-X
    [10] S. L. Krushkal, Proof of the Zalcman conjecture for initial coefficients, Georgian Math. J., 17 (2010), 663–681. https://doi.org/10.1515/gmj.2010.043 doi: 10.1515/gmj.2010.043
    [11] S. L. Krushkal, A short geometric proof of the Zalcman and Bieberbach conjectures, arXiv: 1408.1948.
    [12] W. Ma, Generalized Zalcman conjecture for starlike and typically real functions, J. Math. Anal. Appl., 234 (1999), 328–339.
    [13] R. Mendiratta, S. Nagpal, V. Ravichandran, On a subclass of strongly starlike functions associated with exponential function, Bull. Malays. Math. Sci. Soc., 38 (2015), 365–386. https://doi.org/10.1007/s40840-014-0026-8 doi: 10.1007/s40840-014-0026-8
    [14] P. Goel, S. S. Kumar, Certain class of starlike functions associated with modified sigmoid function, Bull. Malays. Math. Sci. Soc., 43 (2020), 957–991. https://doi.org/10.1007/s40840-019-00784-y doi: 10.1007/s40840-019-00784-y
    [15] R. K. Raina, P. Sharma, J. Sokół, Certain classes of analytic functions related to the crescent-shaped regions, J. Contemp. Math. Anal., 53 (2018), 355–362. https://doi.org/10.3103/S1068362318060067 doi: 10.3103/S1068362318060067
    [16] L. A. Wani, A. Swaminathan, Radius problems for functions associated with a nephroid domain, RACSAM, 114, (2020), 178. https://doi.org/10.1007/s13398-020-00913-4
    [17] S. Gandhi, P. Gupta, S. Nagpal, V. Ravichandran, Starlike functions associated with an Epicycloid, Hacet. J. Math. Stat., 51 (2022), 1637–1660. https://doi.org/10.15672/hujms.1019973 doi: 10.15672/hujms.1019973
    [18] B. Gul, M. Arif, R. K. Alhefthi, D. Breaz, L. I. Cotȋrlă, E. Rapeanu, On the study of starlike functions associated with the generalized sine hyperbolic function, Mathematics, 11 (2023), 4848. https://doi.org/10.3390/math11234848 doi: 10.3390/math11234848
    [19] L. Shi, M. Arif, M. Abbas, M. Ihsan, Sharp bounds of Hankel determinant for the inverse functions on a subclass of bounded turning functions, Mediterr. J. Math., 20 (2023), 156. https://doi.org/10.1007/s00009-023-02371-9 doi: 10.1007/s00009-023-02371-9
    [20] R. J. Libera, E. J. Zlotkiewicz, Coefficient bounds for the inverse of a function with derivative in \mathcal{P}, Proc. Amer. Math. Soc., 87 (1983), 251–257. https://doi.org/10.1090/S0002-9939-1983-0681830-8 doi: 10.1090/S0002-9939-1983-0681830-8
    [21] L. Shi, H. M. Srivastava, A. Rafiq, M. Arif, M. Ihsan, Results on Hankel determinants for the inverse of certain analytic functions subordinated to the exponential function, Mathematics, 10 (2022), 3429. https://doi.org/10.3390/math10193429 doi: 10.3390/math10193429
    [22] M. Raza, A. Riaz, D. K. Thomas, The third Hankel determinant for inverse coefficients of convex functions, Bull. Aust. Math. Soc., 109 (2024), 94–100. https://doi.org/10.1017/S0004972723000357 doi: 10.1017/S0004972723000357
    [23] L. Shi, M. Arif, H. M. Srivastava, M. Ihsan, Sharp bounds on the Hankel determinant of the inverse functions for certain analytic functions, J. Math. Inequal., 17 (2023), 1129–1143. https://doi.org/10.7153/jmi-2023-17-73 doi: 10.7153/jmi-2023-17-73
    [24] C. Pommerenke, On the coefficients and Hankel determinants of univalent functions, J. Lond. Math. Soc., s1-41 (1966), 111–122. https://doi.org/10.1112/jlms/s1-41.1.111 doi: 10.1112/jlms/s1-41.1.111
    [25] C. Pommerenke, On the Hankel determinants of univalent functions, Mathematika, 14 (1967), 108–112. https://doi.org/10.1112/S002557930000807X doi: 10.1112/S002557930000807X
    [26] W. K. Hayman, On second Hankel determinant of mean univalent functions, Proc. Lond. Math. Soc., s3-18 (1968), 77–94. https://doi.org/10.1112/plms/s3-18.1.77 doi: 10.1112/plms/s3-18.1.77
    [27] M. Obradović, N. Tuneski, Hankel determinants of second and third-order for the class \mathcal{S} of univalent functions, Math. Slovaca, 71 (2021), 649–654. https://doi.org/10.1515/ms-2021-0010 doi: 10.1515/ms-2021-0010
    [28] A. Janteng, S. A. Halim, M. Darus, Hankel determinant for starlike and convex functions, Int. J. Math. Anal., 1 (2007), 619–625.
    [29] S. K. Lee, V. Ravichandran, S. Supramaniam, Bounds for the second Hankel determinant of certain univalent functions, J. Inequal. Appl., 2013 (2013), 281. https://doi.org/10.1186/1029-242X-2013-281 doi: 10.1186/1029-242X-2013-281
    [30] A. Ebadian, T. Bulboaca, N. E. Cho, E. A. Adegani, Coefficient bounds and differential subordinations for analytic functions associated with starlike functions, RACSAM, 114 (2020), 128. https://doi.org/10.1007/s13398-020-00871-x doi: 10.1007/s13398-020-00871-x
    [31] N. E. Cho, B. Kowalczyk, O. S. Kwon, A. Lecko, Y. J. Sim, Some coefficient inequalities related to the Hankel determinant for strongly starlike functions of order alpha, J. Math. Inequal., 11 (2017), 429–439. https://doi.org/10.7153/jmi-11-36 doi: 10.7153/jmi-11-36
    [32] B. Kowalczyk, A. Lecko, Y. J. Sim, The sharp bound of the Hankel determinant of the third kind for convex functions, Bull. Aust. Math. Soc., 97 (2018), 435–445. https://doi.org/10.1017/S0004972717001125 doi: 10.1017/S0004972717001125
    [33] B. Kowalczyk, A. Lecko, D. K. Thomas, The sharp bound of the third Hankel determinant for starlike functions, Forum Math., 34 (2022), 1249–1254. https://doi.org/10.1515/forum-2021-0308 doi: 10.1515/forum-2021-0308
    [34] B. Kowalczyk, A. Lecko, The sharp bound of the third Hankel determinant for functions of bounded turning, Bol. Soc. Mat. Mex., 27 (2021), 69. https://doi.org/10.1007/s40590-021-00383-7 doi: 10.1007/s40590-021-00383-7
    [35] A. Lecko, Y. J. Sim, B. Śmiarowska, The sharp bound of the Hankel determinant of the third kind for starlike functions of order 1/2, Complex Anal. Oper. Theory, 13 (2019), 2231–2238. https://doi.org/10.1007/s11785-018-0819-0 doi: 10.1007/s11785-018-0819-0
    [36] A. Riaz, M. Raza, The third Hankel determinant for starlike and convex functions associated with lune, Bull. Sci. Math., 187 (2023), 103289. https://doi.org/10.1016/j.bulsci.2023.103289 doi: 10.1016/j.bulsci.2023.103289
    [37] B. Kowalczyk, A. Lecko, D. K. Thomas, The sharp bound of the third Hankel determinant of convex functions of order -1/2, J. Math. Inequal., 17 (2023), 191–204. https://doi.org/10.7153/jmi-2023-17-14 doi: 10.7153/jmi-2023-17-14
    [38] S. Banga, S. S. Kumar, The sharp bounds of the second and third Hankel determinants for the class \mathcal{SL}, Math. Slovaca, 70 (2020), 849–862. https://doi.org/10.1515/ms-2017-0398 doi: 10.1515/ms-2017-0398
    [39] M. I. Faisal, I. Al-Shbeil, M. Abbas, M. Arif, R. K. Alhefthi, Problems concerning coefficients of symmetric starlike functions connected with the sigmoid function, Symmetry, 15 (2023), 1292. https://doi.org/10.3390/sym15071292 doi: 10.3390/sym15071292
    [40] H. Tang, M. Arif, M. Abbas, F. M. O. Tawfiq, S. N. Malik, Analysis of coefficient-related problems for starlike functions with symmetric points connected with a three-leaf-shaped domain, Symmetry, 15 (2023), 1837. https://doi.org/10.3390/sym15101837 doi: 10.3390/sym15101837
    [41] W. Hu, J. Deng, Hankel determinants, Fekete-Szegö inequality, and estimates of initial coefficients for certain subclasses of analytic functions, AIMS Math., 9 (2024), 6445–6467. https://doi.org/10.3934/math.2024314 doi: 10.3934/math.2024314
    [42] D. V. Prokhorov, J. Szynal, Inverse coefficients for (\alpha, \beta )-convex functions, Ann. Univ. Mariae Curie-Sk lodowska Sect. A, 35 (1981), 125–143.
    [43] P. Zaprawa, M. Obradović, N. Tuneski, Third Hankel determinant for univalent starlike functions, RACSAM, 115 (2021), 49. https://doi.org/10.1007/s13398-020-00977-2 doi: 10.1007/s13398-020-00977-2
    [44] F. Carlson, Sur les coefficients d'une fonction bornée dans le cercle unité, Arkiv för Matematik, Astronomi och Fysik, 27A (1940), 8.
    [45] P. Zaprawa, Initial logarithmic coefficients for functions starlike with respect to symmetric points, Bol. Soc. Mat. Mex., 27 (2021), 62. https://doi.org/10.1007/s40590-021-00370-y doi: 10.1007/s40590-021-00370-y
    [46] M. Arif, L. Rani, M. Raza, P. Zaprawa, Fourth Hankel determinant for the family of functions with bounded turning, Bull. Korean Math. Soci., 55 (2018), 1703–1711. https://doi.org/10.4134/BKMS.b170994 doi: 10.4134/BKMS.b170994
    [47] H. M. Srivastava, M. Arif, M. Raza, Convolution properties of meromorphically harmonic functions defined by a generalized convolution q -derivative operator, AIMS Math., 6 (2021), 5869–5885. https://doi.org/10.3934/math.2021347 doi: 10.3934/math.2021347
    [48] Q. Khan, M. Arif, M. Raza, G. Srivastava, H. Tang, S. U. Rehman, Some applications of a new integral operator in q-analog for multivalent functions, Mathematics, 7 (2019), 1178. https://doi.org/10.3390/math7121178 doi: 10.3390/math7121178
  • This article has been cited by:

    1. Ravikiran A. Mundewadi, Raju B. Jummannaver, Hosoya polynomial method for the numerical solution of Volterra integral equations, 2024, 2731-6734, 10.1007/s43994-024-00191-5
    2. Khadijeh Sadri, David Amilo, Muhammad Farman, Evren Hinçal, Bivariate Jacobi polynomials depending on four parameters and their effect on solutions of time-fractional Burgers’ equations, 2024, 83, 18777503, 102450, 10.1016/j.jocs.2024.102450
    3. Aly R. Seadawy, Bayan A. Alsaedi, Soliton solutions of nonlinear Schrödinger dynamical equation with exotic law nonlinearity by variational principle method, 2024, 56, 1572-817X, 10.1007/s11082-024-06367-x
    4. Syed T. R. Rizvi, Aly R. Seadawy, Bazgha Mustafa, Discussion on chirped periodic and optical soliton solutions for chiral nonlinear Schrödinger dynamical equation with Bohm potential, 2024, 56, 1572-817X, 10.1007/s11082-024-06658-3
    5. Hoorieh Fakhari, Akbar Mohebbi, Galerkin spectral and finite difference methods for the solution of fourth-order time fractional partial integro-differential equation with a weakly singular kernel, 2024, 70, 1598-5865, 5063, 10.1007/s12190-024-02173-6
    6. Waleed Mohamed Abd-Elhameed, Omar Mazen Alqubori, Ahmed Gamal Atta, A collocation procedure for the numerical treatment of FitzHugh–Nagumo equation using a kind of Chebyshev polynomials, 2025, 10, 2473-6988, 1201, 10.3934/math.2025057
    7. Saman Bagherbana, Jafar Biazar, Hossein Aminikhah, Convergence analysis of a novel fractional product integration–spectral collocation scheme for nonlinear weakly singular Volterra integro–differential systems, 2025, 0020-7160, 1, 10.1080/00207160.2025.2465770
    8. Ramy Mahmoud Hafez, Hany Mostafa Ahmed, Omar Mazen Alqubori, Amr Kamel Amin, Waleed Mohamed Abd-Elhameed, Efficient Spectral Galerkin and Collocation Approaches Using Telephone Polynomials for Solving Some Models of Differential Equations with Convergence Analysis, 2025, 13, 2227-7390, 918, 10.3390/math13060918
    9. Waleed Mohamed Abd-Elhameed, Omar Mazen Alqubori, Naher Mohammed A. Alsafri, Amr Kamel Amin, Ahmed Gamal Atta, A Matrix Approach by Convolved Fermat Polynomials for Solving the Fractional Burgers’ Equation, 2025, 13, 2227-7390, 1135, 10.3390/math13071135
    10. David Amilo, Khadijeh Sadri, Evren Hincal, Comparative Analysis of Caputo and Variable-Order Fractional Derivative Algorithms Across Various Applications, 2025, 11, 2349-5103, 10.1007/s40819-025-01887-w
  • 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(737) PDF downloads(46) Cited by(1)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog