Loading [MathJax]/jax/element/mml/optable/BasicLatin.js
Research article

Sum of the triple divisor function of mixed powers

  • Let d3(n) denote the 3-th divisor function. In this paper, we study the asymptotic formula of the sum

    1n1,n2X121n3X1kd3(n21+n22+nk3)

    with n1,n2,n3Z+ and k3 be an integer. Previously only the case of k=2 is studied.

    Citation: Li Zhou, Liqun Hu. Sum of the triple divisor function of mixed powers[J]. AIMS Mathematics, 2022, 7(7): 12885-12896. doi: 10.3934/math.2022713

    Related Papers:

    [1] R. T. Matoog, M. A. Abdou, M. A. Abdel-Aty . New algorithms for solving nonlinear mixed integral equations. AIMS Mathematics, 2023, 8(11): 27488-27512. doi: 10.3934/math.20231406
    [2] Hawsar Ali Hama Rashid, Mudhafar Fattah Hama . Approximate solutions for a class of nonlinear Volterra-Fredholm integro-differential equations under Dirichlet boundary conditions. AIMS Mathematics, 2023, 8(1): 463-483. doi: 10.3934/math.2023022
    [3] Afrah Ahmad Noman Abdou . A fixed point approach to predator-prey dynamics via nonlinear mixed Volterra–Fredholm integral equations in complex-valued suprametric spaces. AIMS Mathematics, 2025, 10(3): 6002-6024. doi: 10.3934/math.2025274
    [4] Mian Bahadur Zada, Muhammad Sarwar, Reny George, Zoran D. Mitrović . Darbo-Type Zm and Lm contractions and its applications to Caputo fractional integro-differential equations. AIMS Mathematics, 2021, 6(6): 6340-6355. doi: 10.3934/math.2021372
    [5] Mohammed A. Almalahi, Satish K. Panchal, Fahd Jarad, Mohammed S. Abdo, Kamal Shah, Thabet Abdeljawad . Qualitative analysis of a fuzzy Volterra-Fredholm integrodifferential equation with an Atangana-Baleanu fractional derivative. AIMS Mathematics, 2022, 7(9): 15994-16016. doi: 10.3934/math.2022876
    [6] Gamal A. Mosa, Mohamed A. Abdou, Ahmed S. Rahby . Numerical solutions for nonlinear Volterra-Fredholm integral equations of the second kind with a phase lag. AIMS Mathematics, 2021, 6(8): 8525-8543. doi: 10.3934/math.2021495
    [7] Marimuthu Mohan Raja, Velusamy Vijayakumar, Anurag Shukla, Kottakkaran Sooppy Nisar, Wedad Albalawi, Abdel-Haleem Abdel-Aty . A new discussion concerning to exact controllability for fractional mixed Volterra-Fredholm integrodifferential equations of order r(1,2) with impulses. AIMS Mathematics, 2023, 8(5): 10802-10821. doi: 10.3934/math.2023548
    [8] Hawsar HamaRashid, Hari Mohan Srivastava, Mudhafar Hama, Pshtiwan Othman Mohammed, Musawa Yahya Almusawa, Dumitru Baleanu . Novel algorithms to approximate the solution of nonlinear integro-differential equations of Volterra-Fredholm integro type. AIMS Mathematics, 2023, 8(6): 14572-14591. doi: 10.3934/math.2023745
    [9] Muhammad Sarwar, Muhammad Fawad, Muhammad Rashid, Zoran D. Mitrović, Qian-Qian Zhang, Nabil Mlaiki . Rational interpolative contractions with applications in extended b-metric spaces. AIMS Mathematics, 2024, 9(6): 14043-14061. doi: 10.3934/math.2024683
    [10] Shahid Bashir, Naeem Saleem, Syed Muhammad Husnine . Fixed point results of a generalized reversed F-contraction mapping and its application. AIMS Mathematics, 2021, 6(8): 8728-8741. doi: 10.3934/math.2021507
  • Let d3(n) denote the 3-th divisor function. In this paper, we study the asymptotic formula of the sum

    1n1,n2X121n3X1kd3(n21+n22+nk3)

    with n1,n2,n3Z+ and k3 be an integer. Previously only the case of k=2 is studied.



    The solution of the mixed Volterra-Fredholm integral equations has been a subject of considerable interest. Studies on population dynamics, parabolic boundary value problems, the mathematical modeling of the spatio-temporal development of an epidemic and various physical and biological models lead to the mixed Volterra-Fredholm integral equations. A discussion of the formulation of such models is given in [1].

    The linear mixed Volterra-Fredholm integral equation is given by

    u(x)=f(x)+λxabak(r,t)u(t)dtdr,axb, (1)

    where f(x) is a known continuous function on the interval [a, b], the kernel k(r, t), is known and continuous on the region D = {r, t: atb & arxb}, and λ, ∈R{0}, while u(x) is the unknown continuous function in [a, b] that must be determined. The solution of this type of equation using Fibonacci collocation method has been discussed by Mirzaee and Hoseini in [2], while the numerical solution via modification of the hat function was presented in [3]. The existence and uniqueness results of equation (1) can be found in [4]. The reproducing kernel method and a study over the role of iterative methods for solving linear and mixed integral equations with variable coefficients were recently investigated in [5,6]. Several other fixed point type and numerical methods were discussed in the literature, see for more [7,8,9,10] and the references therein.

    A linear system of second kind mixed Volterra-Fredholm integral equations (LSMVFIE2nd) can be written as:

    U(x)=F(x)+λxabaK(r,t)U(t)dtdr,axb, (2)

    where

    U(x)=[u1(x),u2(x),,un(x)]t,
    F(x)=[f1(x),f2(x),,fn(x)]t,
    λK(r,t)=[λijkij(r,t)],i,j=1,2,,n.

    The functions fi(x), i = 1, 2, …, n are continuous on the interval [a, b] and all kernels kij(r, t), for i, j = 1, 2, …, n are continuous on D = {r, t: atb & arxb}, while ui(x), i = 1, 2, …, n are the unknown continuous functions to be determined.

    During the last 20 years, significant progress has been made in the numerical analysis for the linear and nonlinear versions of (2), since finding the exact solution in most cases is challenging or even not possible. To recall some of such works, Rabbani and Jamali used variational iterative method to find the solution of nonlinear system of Volterra-Fredholm [11], while Chasemi et al. presented an analytical method to solve this problem using the so called h-curves [12], Wazwaz also approximated the exact solution by Adomian decomposition method and discussed the procedure in the book [13].

    Some other well known techniques including collocation methods which are based on the discretization of the spatial domain are discussed in [14,15,16]. Kakde et al. used fixed point (FP) approach to solve differential and integral equations in [17]. In addition, a fixed oint type scheme was used for solving non-linear quadratic Volterra integral equation in [18], while in [19], the appropriate condition and performance of fixed-point method for Volterra-Hammerstein equation are studied. Finally, in the book [20], Jerri used a fixed-point approach to solve linear Volterra integral equations.

    In this paper, we have made an attempt to propose a numerical scheme to approximate the solution of (2) based on the contractive mapping and a fixed-point method. We study the appropriate conditions and performance of this method for (1). This method has two advantages that encourage us to use it. Firstly, there is not any system of linear equations with its relevant difficulties. Furthermore, it is simple to be applied when programming.

    The main goal of this work is in studying the existence and uniqueness of a continuous solution of the presented system in the Banach space based on the FP theory. In fact, the approximation of the solution is also discussed using an iterative method of fixed point. Finally several experiments are presented to show the efficiency and accuracy of the proposed methods.

    Before going to the next section, several definitions and theorems are reminded as follows [13,15].

    Definition 1. For a metric space (M, d), let M°⊆M with a map f:M°→M, a point pM° is said to be a fixed point of f if f(p) = p.

    Definition 2. Let (M, d) be a complete metric space, a mapping f:MM is said to be contraction if αR for 0 ≤ α < 1 such that

    d(f(u),f(v))αd(u,v),u,vM.

    Theorem 1.1. (Fixed-point Theorem; FPT) If the mapping f:MM is contraction on a complete metric space (M, d), then f has a unique fixed point xM.

    The FPM provides a scheme which is used to solve LSMVFIE2nd by starting with an initial approximation that will be used in a recurrence relation to find the other approximate solutions. First, consider the system (2) and define the operator T as follows:

    U(x)=T(u)=F(x)+λxabaK(r,t)U(t)dtdr,whereu=u1,u2,,un, (3)

    where F, U and K are defined in Section 1, while T=[T1,T2,,Tn]t. The solution of the system (2) is fixed-point of T. Choose ui0(x)C[a,b] as an initial function and the following fixed-point iteration will be introduced [4]:

    uir(x)=Ti(ur1)=fi(x)+nj=1λijxabakij(r,t)ujr1(t)dtdr. (4)

    Applying (4) will determine multiple approximations uir(x), for i = 1, 2, …, n, r≥1 and the sequence {uin} converges to U(x) as n→∞. It is just the contractive property which is responsible for clustering the sequence {uin} towards a limit point. Then the major concepts that required to the FPT are contraction mapping and a complete metric space.

    The theorem below shows that T becomes a contractive mapping under some assumptions.

    Theorem 2.1. For a complete metric space (C[a, b], ), and the continuous functions FC[a, b] and KC([a, b]×[a, b]), if the following conditions satisfied

    λi<1n(ba)2Mi,i=1,2,,n.

    Then, the mapping T that defined in (3) becomes a contractive mapping.

    Proof: Since the kernels are continuous on abounded region, then |Kij(r,t)|Mij,i,j=1,2,,n for some positive real numbers Mij. For such functions we work with the complete metric space (C[a, b], ).

    Now, to find a sufficient condition for the mapping TU in (3) to be a contractive mapping, consider the ith equation of the equation system (2) as follows:

    ui(x)=Ti(u)=fi(x)+nj=1λijxabakij(r,t)uj(t)dtdr,whereu=u1,u2,,un. (5)

    Now for some v={v1,v2,,vn} and for each i = 1, 2, …, n; we have

    u|Ti(u(x))Ti(v(x))|=|fi(x)+nj=1λijxabakij(r,t)uj(t)dtdr[fi(x)+nj=1λijxabakij(r,t)vj(t)dtdr]|=|nj=1λijxabakij(r,t)(uj(t)vj(t))dtdr|nj=1|λij|xaba|kij(r,t)||(uj(t)vj(t))|dtdr.

    Thus, it is possible to find

    |Ti(u(x))Ti(v(x))|nj=1|λij|Mijxaba(ujvj)dtdr. (6)

    Now, we have

    maxx[a,b]|Ti(u(x))Ti(v(x))|maxx[a,b]nj=1|λij|Mijxaba(ujvj)dtdr, (7)

    and

    (Ti(u)Ti(v))(ba)2nj=1|λij|Mij(ujvj). (8)

    Let

    λi=maxj=1,2,,n|λij|,andMi=maxj=1,2,,nMij (9)
    (Ti(u)Ti(v))(ba)2λiMinj=1(ujvj)n(ba)2λiMimaxj=1,2,,n(ujvj). (10)

    Hence, if

    0αi=n(ba)2λiMi<1,

    or similarly as long as we have

    λi<1n(ba)2Mi,i=1,2,,n, (11)

    Then, the mapping T of the LSMVFIE is a contractive mapping. The proof is complete now.

    To show that there exists only one solution for LSMVFIE2nd, we have to prove that T has a unique FP and the generated sequence {uin}n=0 in (4) converges to this FP. The following theorem justifies the convergence of uir(x).

    Theorem 3.1. Let (C[a, b], ) be a complete metric space and Ti be n contraction mapping on the LSMVFIE2nd as defined in equation (3) then for each i = 1, 2, …, n, we obtain:

    . Ti has a unique fixed-point uiC[a,b] such that ui=Ti(u)

    . For any ui0C[a,b], the sequence {uir(x)}C[a,b] defined by uir(x)=Ti(ur1), for r = 0, 1, …, converges to ui.

    Proof: ⅰ. Taking limits of both sides of uir(x)=Ti(ur1), we have:

    lim

    where Ti is the contraction mapping for each i, and Ti is continuous. So, we obtain:

    \mathop {\lim }\limits_{n \to \infty } {u_i}^r = \mathop {{T_i}{\rm{lim}}}\limits_{n \to \infty } \left( {{\mathit{\boldsymbol{u}}^{r - 1}}} \right).

    Thus, {u_i}^{\rm{*}} = {T_i}\left({{\mathit{\boldsymbol{u}}^*}} \right), for {\mathit{\boldsymbol{u}}^*} = {u_1}^*, {u_2}^*, \ldots, {u_n}^*. Hence Ti has a fixed-point i = 1, 2, …, n. Now by reductio ad absurdum, suppose that (if possible) vi* is also a fixed-point of Ti, this means that {v_i}^{\rm{*}} = {T_i}\left({{\mathit{\boldsymbol{v}}^*}} \right) for {\mathit{\boldsymbol{v}}^*} = {v_1}^*, {v_2}^*, \ldots, {v_n}^*, then

    {\left\| {{u_i}^{\rm{*}} - {v_i}^{\rm{*}}} \right\|_\infty } \le {\alpha _i}\mathop {max}\limits_{j = 1,2, \ldots ,n} {\left\| {{u_j}^{\rm{*}} - {v_j}^{\rm{*}}} \right\|_\infty }.

    This means αi≤1, this is a contradiction. Thus {u_i}^{\rm{*}} = {v_i}^{\rm{*}}.

    ⅱ. From part (ⅰ), we have:

    {\left\| {{u_i}^r\left( x \right) - {u_i}^*\left( x \right)} \right\|_\infty } = {\left\| {{T_i}\left( {{{\boldsymbol{u}}^{r - 1}}} \right) - {T_i}\left( {{{\boldsymbol{u}}^*}} \right)} \right\|_\infty } \le {\alpha _i}\mathop {max}\limits_{j = 1,2, \ldots ,n} {\left\| {{u_j}^{r - 1} - {u_j}^*} \right\|_\infty }.

    Let \mathop {max}\limits_{j = 1,2, \ldots ,n} {\left\| {{u_j}^{r - 1} - {u_j}^*} \right\|_\infty } = {\left\| {{u_{l1}}^{r - 1} - {u_{l1}}^*} \right\|_\infty } and then we have

    {\left\| {{u_{l1}}^{r - 1} - {u_{l1}}} \right\|^*}_\infty = {\left\| {{T_{l1}}\left( {{{\boldsymbol{u}}^{r - 2}}} \right) - {T_{l1}}\left( {{{\boldsymbol{u}}^*}} \right)} \right\|_\infty } \le {\alpha _{l1}}\mathop {max}\limits_{l2 = 1,2, \ldots ,n} {\left\| {{u_{l2}}^{r - 2} - {u_{l2}}^*} \right\|_\infty }.

    Thus, we get the following inequality to complete this part:

    \begin{array}{l} {\left\| {{u_i}^r\left( x \right) - {u_i}^*\left( x \right)} \right\|_\infty } \le {\alpha _i}\mathop {max}\limits_{j = 1,2, \ldots ,n} {\left\| {{u_j}^{r - 1} - {u_j}^*} \right\|_\infty } = {\alpha _i}{\left\| {{u_{l1}}^{r - 1} - {u_{l1}}^*} \right\|_\infty }\\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\; \le {\alpha _i}{\alpha _{l1}}\mathop {max}\limits_{l2 = 1,2, \ldots ,n} {\left\| {{u_{l2}}^{r - 2} - {u_{l2}}^*} \right\|_\infty }\\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\; \le \ldots \le ({\alpha _i}{\alpha _{l1}}{\alpha _{l2}} \ldots {\alpha _{l\left( {r - 1} \right)}})\mathop {max}\limits_{lr = 1,2, \ldots ,n} {\left\| {{u_{lr}}^0 - {u_{lr}}^*} \right\|_\infty }\\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\; \le {\alpha ^r}\mathop {max}\limits_{lr = 1,2, \ldots ,n} {\left\| {{u_{lr}}^0 - {u_{lr}}^*} \right\|_\infty }, \end{array}

    where {\alpha ^r} = max\left\{ {{\alpha _i}, {\alpha _{l1}}, {\alpha _{l2}}, \ldots, {\alpha _{l\left({r - 1} \right)}}} \right\}. Thus \mathop {\lim }_{r \to \infty } {\left\| {{u_i}^r\left( x \right) - {u_i}^*\left( x \right)} \right\|_\infty } \le \mathop {\lim }_{r \to \infty } {\alpha ^r}{\left\| {{u_{lr}}^0 - {u_{lr}}^*} \right\|_\infty }. Since 0≤α < 1, αr→0 as r→∞. This means that {u_i}^r \to {u_i}^*, for i = 1, 2, …, n. The proof is ended now.

    As discussed in section 1 and since the exact solution of the LSMVFIE2nd could be found in all cases, here an approximate-analytic solution is presented to find the solution specially for cases at which there is no special peak or oscillation in the solution of the problem.

    Perform the steps below to get the approximation for LSMVFIE2nd by using FPM:

    Step 1: Choose a, b as the integration bound, n as the number of unknown functions, and m as the number of points in the interval [a, b].

    Step 2: Let {u_i}^0\left(x \right) = {f_i}\left(x \right) be an initial solution.

    Step 3: Calculate {u_i}^r\left(x \right) in (3) for all i = 1, 2, …, n.

    Step 4: Repeat Step 3 until desired level of accuracy is reached.

    Step 5: Find {u_i}^r\left({{x_j}} \right), {\rm{}}{x_j} \in \left[{a, b} \right], where {x_j} = {\rm{}}{x_0} + jh, for j = 1, 2, …, m where h = \frac{{b - a}}{m}.

    Step 6: Compute \left| {{u_i}\left({{x_j}} \right) - {\rm{}}{u_i}^r\left({{x_j}} \right)} \right|, for any i and j.

    To show the implementation of the method and the accuracy of the approach two tests will be solved in this section.

    Example 1. Consider the following LSMVFIE2nd

    \begin{array}{l} {u_1}\left( x \right) = {f_1} + \frac{1}{2}\mathop \smallint \limits_0^x \mathop \smallint \limits_0^1 {k_{11}}{u_1}\left( t \right)dtdr + \frac{1}{4}\mathop \smallint \limits_0^x \mathop \smallint \limits_0^1 {k_{12}}{u_2}\left( t \right)dtdr, \;\;\;\;0 \le x \le 1, \\ {u_2}\left( x \right) = {f_2} + \frac{1}{3}\mathop \smallint \limits_0^x \mathop \smallint \limits_0^1 {k_{21}}{u_1}\left( t \right)dtdr + \frac{1}{4}\mathop \smallint \limits_0^x \mathop \smallint \limits_0^1 {k_{22}}{u_2}\left( t \right)dtdr, \;\;\;\;\;0 \le x \le 1, \end{array}

    where

    \begin{array}{l} {k_{11}}\left( {r, t} \right) = \left( {r - t} \right), {k_{12}}\left( {r, t} \right) = rt, \\ \;\;\;\;{k_{21}}\left( {r, t} \right) = rs + 1, {k_{22}}\left( {r, t} \right) = r, \\ {f_1}\left( x \right) = \frac{x}{2} + {e^x} - \frac{{\left( {{x^2}\left( {\frac{e}{2} - \frac{1}{2}} \right)} \right)}}{2} - \frac{{{x^2}}}{{32}}, \\ \;\;\;\;\;\;\;\;\;\;\;\;{f_2}\left( x \right) = \frac{{19{x^2}}}{{24}} - \frac{{\left( {x\left( {e - 1} \right)} \right)}}{3}, \end{array}

    and the exact solution is given by {u_1}\left(x \right) = {e^x}, \; {\rm{and}}\; {u_2}\left(x \right) = {x^2}. Recalling that the solution does not have sharp behaviors including oscillations or peaks and thus the system is a good candidate for taking into account the approximate-analytic methods, such as the one discussed in section 4.

    First, we let

    {u_1}^0\left( x \right) = \frac{x}{2} + {e^x} - \frac{{\left( {{x^2}\left( {\frac{e}{2} - \frac{1}{2}} \right)} \right)}}{2} - \frac{{{x^2}}}{{32}}, \;\;{\rm{and}}\;\;\;{u_2}^0\left( x \right) = \frac{{19{x^2}}}{{24}} - \frac{{\left( {x\left( {e - 1} \right)} \right)}}{3}.

    Now by using (4) the values of {u_1}^i\left(x \right) and {u_2}^i\left(x \right) will be determined for all i = 1, …, 10. The absolute errors for the approximate solutions {u_1}^{10}\left(x \right)\; {\rm{and}}\; {u_2}^{10}\left(x \right) are listed in Tables 1 and 2 respectively.

    Table 1.  The results for {u_1}^{10}\left(x \right) and the absolute errors of Example 1.
    {x_j} Exact solution {u_1}\left({{x_j}} \right) Approximate Values of u_1^{10}({x_j}) Absolute error \left| {{u_1}\left({{x_j}} \right) - {\rm{}}u_1^{10}\left({{x_j}} \right)} \right|
    0.1 1.10517091807 1.10517091808 9.034 e-12
    0.2 1.22140275816 1.22140275817 1.7206 e-11
    0.3 1.34985880757 1.34985880760 2.4518 e-11
    0.4 1.49182469764 1.49182469767 3.0969 e-11
    0.5 1.64872127070 1.64872127073 3.6558 e-11
    0.6 1.82211880039 1.82211880043 4.1287 e-11
    0.7 2.01375270747 2.01375270751 4.5155 e-11
    0.8 2.22554092849 2.22554092854 4.8161 e-11
    0.9 2.45960311115 2.45960311120 5.0306 e-11
    1.0 2.71828182845 2.71828182851 5.1592 e-11

     | Show Table
    DownLoad: CSV
    Table 2.  The results u_2^{10}\left(x \right) and the absolute errors of Example 1.
    {x_{\rm{j}}} Exact solution {u_2}\left({{x_{\rm{j}}}} \right) Approximate Values of u_2^{10}({x_j}) Absolute error \left| {{u_2}\left({{x_{\rm{j}}}} \right) - {\rm{}}u_2^{10}\left({{x_{\rm{j}}}} \right)} \right|
    0 0 0 0
    0.1 0.01 0.00999999999 9.808 e-12
    0.2 0.04 0.03999999998 1.927 e-11
    0.3 0.09 0.08999999997 2.840 e-11
    0.4 0.16 0.15999999996 3.718 e-11
    0.5 0.25 0.24999999995 4.563 e-11
    0.6 0.36 0.35999999995 5.373 e-11
    0.7 0.49 0.48999999994 6.149 e-11
    0.8 0.64 0.63999999993 6.891 e-11
    0.9 0.81 0.80999999992 7.599 e-11
    1 1 0.99999999992 8.273 e-11

     | Show Table
    DownLoad: CSV

    Comparison between the exact and approximate solutions have been made and illustrated depending on least square error (L.S.E.) and the time required for running the program (R.T.) together in Table 3. While, Figure 1 shows the convergence of the method for {u_1}^i\left(x \right)\; {\rm{and}}\; {u_2}^i\left(x \right) with i = 1, 2, 3.

    Table 3.  The exact solution is compared with the numerical method for different iterates.
    n 3-iterations 6-iterations 9-iterations 12-terations
    L.S.E u1 1.17e-03 8.35 e-11 3.73 e-17 5.45 e-24
    L.S.E u2 2.73 e-04 5.07 e-10 7.43 e-18 1.18 e-23
    R.T. 2.4s 5.6s 9.1s 14.5s

     | Show Table
    DownLoad: CSV
    Figure 1.  The exact and numerical solution with n = 1, 2, 3 for Example 1.

    Example 2. Consider the following system of integral equations:

    {u_1}\left( x \right) = cos\left( x \right) + \frac{\pi }{{16}}x - \left( {\frac{{{\pi ^2} + 4}}{{128}}} \right){x^2} + \frac{1}{4}\mathop \smallint \limits_0^x \mathop \smallint \limits_0^{\frac{\pi }{2}} \left( {rt - 1} \right){u_2}\left( t \right)dtdr, \;\;\;\;\;\;0 \le x \le \frac{\pi }{2},
    {u_2}\left( x \right) = si{n^2}\left( x \right) + \left( {\frac{{\pi - 2}}{8}} \right)x - \frac{{{x^2}}}{8} + \frac{1}{4}\mathop \smallint \limits_0^x \mathop \smallint \limits_0^{\frac{\pi }{2}} \left( {r - t} \right){u_2}\left( t \right)dtdr, \;\;\;\;\;\;\;\;\;\;0 \le x \le \frac{\pi }{2},

    where the exact solution is given by {u_1}\left(x \right) = \cos \left(x \right), and {u_2}\left(x \right) = {\sin ^2}\left(x \right).

    Applying the algorithm of FPM with n = 10, gives the results in Tables 4 and 5. While, Figure 2 shows the convergence of the method for u_1^i\left(x \right)\; {\rm{and}}\; u_2^i\left(x \right) for i = 1, 2, 3.

    Table 4.  The results u_1^{10}\left(x \right) of Example 2, compared with exact solution {u_1}\left(x \right).
    {x_j} Exact solution {u_1}\left({{x_j}} \right) Approximate Values of u_1^{10}({x_j}) Absolute error \left| {{u_1}\left({{x_j}} \right) - {\rm{}}u_1^{10}\left({{x_j}} \right)} \right|
    0.1(π) 0.95105651630 0.95105651644 1.52318 e-10
    0.2(π) 0.80901699437 0.80901699463 2.55728 e-10
    0.3(π) 0.58778525229 0.58778525260 3.10231 e-10
    0.4(π) 0.30901699437 0.30901699469 3.15826 e-10
    0.5(π) 0 2.72513 e-10 2.72513 e-10

     | Show Table
    DownLoad: CSV
    Table 5.  The results u_2^{10}(x) of Example 2, compared with exact solution {u_2}\left(x \right).
    {x_j} Exact solution {u_2}\left({{x_j}} \right) Approximate Values of u_2^{10}({x_j}) Absolute error \left| {{u_2}\left({{x_j}} \right) - {\rm{}}u_2^{10}\left({{x_j}} \right)} \right|
    0.1(π) 0.09549150281 0.09549150289 8.2427 e-11
    0.2(π) 0.34549150281 0.34549150294 1.3157 e-10
    0.3(π) 0.65450849719 0.65450849734 1.4742 e-10
    0.4(π) 0.90450849719 0.90450849732 1.3003 e-10
    0.5(π) 1 1.00000000008 7.9341 e-11

     | Show Table
    DownLoad: CSV
    Figure 2.  The exact solution and the numerical solution with n = 1, 2, 3, 4 for Example 2.

    In summary, the existence of a unique solution was verified and proved for linear system of mixed Volterra-Fredholm integral equations of the second kind and the result was obtained by using some principles of Banach's contraction in complete metric space. Also, an iterative technique based on fixed point method was discussed for solving the system, and an algorithm is constructed. Here our programs are written by Matlab software 2015Ra, and two experiments were presented for illustration. Good approximations were obtained, while better results have been found by increasing the number of iterations (n). Moreover, comparison between the exact and approximate solutions was made to demonstrate the application method. It is worth mentioning that the technique can be used as a very accurate algorithm for solving linear system of mixed Volterra-Fredholm integral equations of the second kind. These claims are supported by the results of the given numerical examples in Tables 15 and Figures 12.

    The authors declare no conflict of interest in this paper.



    [1] C. E. Chace, The divisor problem for arithmetic progressions with small modulus, Acta Arith., 61 (1992), 35–50.
    [2] C. E. Chace, Writing integers as sums of products, Trans. Am. Math. Soc., 345 (1994), 367–379. https://doi.org/10.1090/S0002-9947-1994-1257641-3 doi: 10.1090/S0002-9947-1994-1257641-3
    [3] R. T. Guo, W. G. Zhai, Some problems about the ternary quadratic form m_1^2 +m_2^2 +m_3^2, Acta Arith., bf 156 (2012), 101–121.
    [4] L. Q. Hu, An asymptotic formula related to the divisors of the quaternary quadratic form, Acta Arith., 166 (2014), 129–140. https://doi.org/10.4064/aa166-2-2 doi: 10.4064/aa166-2-2
    [5] L. Q. Hu and L. Yang, Sums of the triple divisor function over values of a quaternary quadratic form, Acta Arith., 183 (2018), 63–85. https://doi.org/10.4064/aa170120-20-10 doi: 10.4064/aa170120-20-10
    [6] L. K. Hua, Introduction to Number Theory, Science Press, Beijing, 1957 (in Chinese).
    [7] X. D. Lü, Q. W. Mu, The Sum of Divisors of Mixed Powers, Advances in Mathematics (in China), 45 (2016), 357–364.
    [8] Q. F. Sun, D. Y. Zhang, Sums of the triple divisor function over values of a ternary quadratic form, J. Number Theory, 168 (2016), 215–246.
    [9] R. C. Vaughan, The Hardy-Littlewood Method, 2nd ed., Cambridge Tracts in Math., vol. 125, Cambridge University, Cambridge, 1997.
    [10] L. L. Zhao, The sum of divisors of a quadratic form, Acta Arith., 163 (2014), 161–177. https://doi.org/10.4064/aa163-2-6 doi: 10.4064/aa163-2-6
  • This article has been cited by:

    1. Amira Zaki, Ahmed Métwalli, Moustafa H. Aly, Waleed K. Badawi, Enhanced feature selection method based on regularization and kernel trick for 5G applications and beyond, 2022, 61, 11100168, 11589, 10.1016/j.aej.2022.05.024
    2. Yan Tang, Yeyu Zhang, Aviv Gibali, New Self-Adaptive Inertial-like Proximal Point Methods for the Split Common Null Point Problem, 2021, 13, 2073-8994, 2316, 10.3390/sym13122316
    3. Chinedu Nwaigwe, Deborah Ngochinma Benedict, Generalized Banach fixed-point theorem and numerical discretization for nonlinear Volterra–Fredholm equations, 2023, 425, 03770427, 115019, 10.1016/j.cam.2022.115019
    4. Chinedu Nwaigwe, Sanda Micula, Fast and Accurate Numerical Algorithm with Performance Assessment for Nonlinear Functional Volterra Equations, 2023, 7, 2504-3110, 333, 10.3390/fractalfract7040333
    5. Chinedu Nwaigwe, Azubuike Weli, Dang Ngoc Hoang Thanh, Fourth-Order Trapezoid Algorithm with Four Iterative Schemes for Nonlinear Integral Equations, 2023, 44, 1995-0802, 2822, 10.1134/S1995080223070314
  • Reader Comments
  • © 2022 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(1981) PDF downloads(80) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog