Abstract

This paper investigates the numerical solution of nonlinear Fredholm-Volterra integro-differential equations using reproducing kernel Hilbert space method. The solution 𝑢(𝑥) is represented in the form of series in the reproducing kernel space. In the mean time, the n-term approximate solution 𝑢𝑛(𝑥) is obtained and it is proved to converge to the exact solution 𝑢(𝑥). Furthermore, the proposed method has an advantage that it is possible to pick any point in the interval of integration and as well the approximate solution and its derivative will be applicable. Numerical examples are included to demonstrate the accuracy and applicability of the presented technique. The results reveal that the method is very effective and simple.

1. Introduction

In recent years, there has been a growing interest in the integrodifferential equations (IDEs) which are a combination of differential and Fredholm-Volterra integral equations. IDEs are often involved in the mathematical formulation of physical phenomena. IDEs can be encountered in various fields of science such as physics, biology, and engineering. These kinds of equations can also be found in numerous applications, such as biomechanics, electromagnetic, elasticity, electrodynamics, fluid dynamics, heat and mass transfer, and oscillation theory [14].

The purpose of this paper is to extend the application of the reproducing kernel Hilbert space (RKHS) method to solve the nonlinear Fredholm-Volterra IDE which is as follows: 𝑢(𝑥)+𝑓(𝑥)𝑁(𝑥,𝑢(𝑥))+𝑏𝑎𝑘1(𝑥,𝑡)𝐺1(𝑢(𝑡))𝑑𝑡+𝑥𝑎𝑘2(𝑥,𝑡)𝐺2(𝑢(𝑡))𝑑𝑡=𝑔(𝑥),𝑎𝑥,𝑡𝑏,(1.1) subject to the initial condition 𝑢(𝑎)=𝛼,(1.2) where 𝑎,𝑏,𝛼 are real finite constants, 𝑢𝑊22[𝑎,𝑏] is an unknown function to be determined, 𝑓,𝑔𝑊12[𝑎,𝑏], 𝑘1(𝑥,𝑡), 𝑘2(𝑥,𝑡) are continuous functions on [𝑎,𝑏]×[𝑎,𝑏], 𝐺1(𝑤), 𝐺2(𝑦), 𝑁(𝑥,𝑧) are continuous terms in 𝑊12[𝑎,𝑏] as 𝑤=𝑤(𝑥), 𝑦=𝑦(𝑥), 𝑧=𝑧(𝑥)𝑊22[𝑎,𝑏], 𝑎𝑥𝑏, <𝑤,𝑦,𝑧< and are depending on the problem discussed, and 𝑊12[𝑎,𝑏],𝑊22[𝑎,𝑏] are reproducing kernel spaces.

In general, nonlinear Fredholm-Volterra IDEs do not always have solutions which we can obtain using analytical methods. In fact, many of real physical phenomena encountered are almost impossible to solve by this technique. Due to this, some authors have proposed numerical methods to approximate the solutions of nonlinear Fredholm-Volterra IDEs. To mention a few, in [5] the authors have discussed the Taylor polynomial method for solving IDEs (1.1) and (1.2) when 𝑁(𝑥,𝑢(𝑥))=1, 𝐺1(𝑢(𝑡))=𝑢(𝑡), and 𝐺2(𝑢(𝑡))=[𝑢(𝑡)]𝑞, where 𝑞. The triangular functions method has been applied to solve the same equations when 𝑁(𝑥,𝑢(𝑥))=𝑢(𝑥), 𝐺1(𝑢(𝑡))=[𝑢(𝑡)]𝑝, and 𝐺2(𝑢(𝑡))=[𝑢(𝑡)]𝑞, where 𝑝, 𝑞 as described in [6]. Furthermore, the operational matrix with block-pulse functions method is carried out in [7] for the aforementioned IDEs in the case 𝑁(𝑥,𝑢(𝑥))=1, 𝐺1(𝑢(𝑡))=[𝑢(𝑡)]𝑝, and 𝐺2(𝑢(𝑡))=[𝑢(𝑡)]𝑞, where 𝑝,𝑞. Recently, the Hybrid Legendre polynomials and Block-Pulse functions approach for solving IDEs (1.1) and (1.2) when 𝑁(𝑥,𝑢(𝑥))=1, 𝐺1(𝑢(𝑡))=[𝑢(𝑡)]𝑝, and 𝐺2(𝑢(𝑡))=[𝑢(𝑡)]𝑞, where 𝑝,𝑞 are proposed in [8]. The numerical solvability of Fredholm and Volterra IDEs and other related equations can be found in [911] and references therein. However, none of previous studies propose a methodical way to solve these equations. Moreover, previous studies require more effort to achieve the results, they are not accurate, and usually they are developed for special types of IDEs (1.1) and (1.2).

Reproducing kernel theory has important application in numerical analysis, differential equations, integral equations, probability and statistics, and so on [1214]. Recently, using the RKHS method, the authors in [1529] have discussed singular linear two-point boundary value problems, singular nonlinear two-point periodic boundary value problems, nonlinear system of boundary value problems, initial value problems, singular integral equations, nonlinear partial differential equations, operator equations, and fourth-order IDEs.

The outline of the paper is as follows: several reproducing kernel spaces are described in Section 2. In Section 3, a linear operator, a complete normal orthogonal system, and some essential results are introduced. Also, a method for the existence of solutions for (1.1) and (1.2) based on reproducing kernel space is described. In Section 4, we give an iterative method to solve (1.1) and (1.2) numerically in RKHS. Various numerical examples are presented in Section 5. This paper ends in Section 6 with some concluding remarks.

2. Several Reproducing Kernel Spaces

In this section, several reproducing kernels needed are constructed in order to solve (1.1) and (1.2) using RKHS method. Before the construction, we utilize the reproducing kernel concept. Throughout this paper is the set of complex numbers, 𝐿2[𝑎,𝑏]={𝑢𝑏𝑎𝑢2(𝑥)𝑑𝑥<}, 𝑙2={𝐴𝑖=1(𝐴𝑖)2<}, and the superscript (𝑛) in 𝑢(𝑛)(𝑡) denotes the 𝑛-th derivative of 𝑢(𝑡).

Definition 2.1 (see [18]). Let 𝐸 be a nonempty abstract set. A function 𝐾𝐸×𝐸 is a reproducing kernel of the Hilbert space 𝐻 if(1)for each 𝑡𝐸, 𝐾(,𝑡)𝐻,(2)for each 𝑡𝐸 and 𝜑𝐻, 𝜑(),𝐾(,𝑡)=𝜑(𝑡).

The last condition is called “the reproducing property”: the value of the function 𝜑 at the point 𝑡 is reproducing by the inner product of 𝜑 with 𝐾(,𝑡). A Hilbert space which possesses a reproducing kernel is called a RKHS [18].

Next, we first construct the space 𝑊22[𝑎,𝑏] in which every function satisfies the initial condition (1.2) and then utilize the space 𝑊12[𝑎,𝑏].

Definition 2.2 (see [30]). 𝑊22[𝑎,𝑏]={𝑢𝑢,𝑢 are absolutely continuous on [𝑎,𝑏], 𝑢,𝑢,𝑢𝐿2[𝑎,𝑏], and 𝑢(𝑎)=0}. The inner product and the norm in 𝑊22[𝑎,𝑏] are defined, respectively, by 𝑢,𝑣𝑊22=𝑢(𝑎)𝑣(𝑎)+𝑢(𝑎)𝑣(𝑎)+𝑏𝑎𝑢(𝑦)𝑣(𝑦)𝑑𝑦(2.1) and 𝑢𝑊22=𝑢,𝑢𝑊22, where 𝑢,𝑣𝑊22[𝑎,𝑏].

Definition 2.3 (see [23]). 𝑊12[𝑎,𝑏]={𝑢𝑢 is absolutely continuous on [𝑎,𝑏] and 𝑢,𝑢𝐿2[𝑎,𝑏]}. The inner product and the norm in 𝑊12[𝑎,𝑏] are defined, respectively, by 𝑢,𝑣𝑊12=𝑏𝑎𝑢(𝑡)𝑣(𝑡)+𝑢(𝑡)𝑣(𝑡)𝑑𝑡 and 𝑢𝑊12=𝑢,𝑢𝑊12, where 𝑢,𝑣𝑊12[𝑎,𝑏].

In [23], the authors have proved that the space𝑊12[𝑎,𝑏] is a complete reproducing kernel space and its reproducing kernel function is given by 𝑇𝑥1(𝑦)=||||.2sinh(𝑏𝑎)cosh(𝑥+𝑦𝑏𝑎)+cosh𝑥𝑦𝑏+𝑎(2.2) From the definition of the reproducing kernel spaces 𝑊12[𝑎,𝑏] and 𝑊22[𝑎,𝑏], we get 𝑊12[𝑎,𝑏]𝑊22[𝑎,𝑏].

The Hilbert space 𝑊22[𝑎,𝑏] is called a reproducing kernel if for each fixed 𝑥[𝑎,𝑏] and any 𝑢(𝑦)𝑊22[𝑎,𝑏], there exist 𝐾(𝑥,𝑦)𝑊22[𝑎,𝑏] (simply 𝐾𝑥(𝑦)) and 𝑦[𝑎,𝑏] such that 𝑢(𝑦),𝐾𝑥(𝑦)𝑊22=𝑢(𝑥). The next theorem formulates the reproducing kernel function 𝐾𝑥(𝑦).

Theorem 2.4. The Hilbert space 𝑊22[𝑎,𝑏] is a reproducing kernel and its reproducing kernel function 𝐾𝑥(𝑦) can be written as 𝐾𝑥(𝑦)=4𝑖=1𝑝𝑖(𝑥)𝑦𝑖1,𝑦𝑥,4𝑖=1𝑞𝑖(𝑥)𝑦𝑖1,𝑦>𝑥,(2.3) where 𝑝𝑖(𝑥) and 𝑞𝑖(𝑥) are unknown coefficients of 𝐾𝑥(𝑦).

Proof. Through several integrations by parts for (2.1), we obtain 𝑢(𝑦),𝐾𝑥(𝑦)𝑊22=1𝑖=0𝑢(𝑖)(𝑎)(𝐾𝑥(𝑖)(𝑎)+(1)𝑖𝐾𝑥(3𝑖)(𝑎))+1𝑖=0(1)1𝑖𝑢(𝑖)(𝑏)𝐾𝑥(3𝑖)(𝑏)+𝑏𝑎𝑢(𝑦)𝐾𝑥(4)(𝑦)𝑑𝑦. Since 𝐾𝑥(𝑦)𝑊22[𝑎,𝑏], it follows that 𝐾𝑥(𝑎)=0. Also, since 𝑢𝑊22[𝑎,𝑏], one obtains 𝑢(𝑎)=0. Thus, if 𝐾𝑥(𝑖)(𝑏)=0, 𝑖=2,3, and 𝐾𝑥(𝑎)𝐾𝑥(𝑎)=0, then 𝑢(𝑦),𝐾𝑥(𝑦)𝑊22=𝑏𝑎𝑢(𝑦)𝐾𝑥(4)(𝑦)𝑑𝑦. Now, for each 𝑥[𝑎,𝑏], if 𝐾𝑥(𝑦) also satisfies 𝐾𝑥(4)(𝑦)=𝛿(𝑥𝑦), where 𝛿 is the dirac-delta function, then 𝑢(𝑦),𝐾𝑥(𝑦)𝑊22=𝑢(𝑥). Obviously, 𝐾𝑥(𝑦) is the reproducing kernel function of the space 𝑊22[𝑎,𝑏].
Next, we give the expression of the reproducing kernel function 𝐾𝑥(𝑦). The characteristic equation of 𝐾𝑥(4)(𝑦)=𝛿(𝑥𝑦) is 𝜆4=0, and their characteristic values are 𝜆=0 with 4 multiple roots. So, let the kernel 𝐾𝑥(𝑦) be as defined in (2.3).
On the other hand, let 𝐾𝑥(𝑦) satisfy 𝐾𝑥(𝑚)(𝑥+0)=𝐾𝑥(𝑚)(𝑥0), 𝑚=0,1,2. Integrating 𝐾𝑥(4)(𝑦)=𝛿(𝑥𝑦) from 𝑥𝜀 to 𝑥+𝜀 with respect to 𝑦 and letting 𝜀0, we have the jump degree of 𝐾𝑥(3)(𝑦) at 𝑦=𝑥 given by 𝐾𝑥(3)(𝑥0)𝐾𝑥(3)(𝑥+0)=1. Through the last descriptions the unknown coefficients of (2.3) can be obtained. This completes the proof.

By using Mathematica 7.0 software package, the representation of the reproducing kernel function 𝐾𝑥(𝑦) is provided by 𝐾𝑥1(𝑦)=6(𝑦𝑎)2𝑎2𝑦21+3𝑥(2+𝑦)𝑎(6+3𝑥+𝑦),𝑦𝑥,6(𝑥𝑎)2𝑎2𝑥2+3𝑦(2+𝑥)𝑎(6+𝑥+3𝑦),𝑦>𝑥.(2.4)

The following corollary summarizes some important properties of the reproducing kernel function 𝐾𝑥(𝑦).

Corollary 2.5. The reproducing kernel function 𝐾𝑥(𝑦) is symmetric, unique, and 𝐾𝑥(𝑥)0 for any fixed 𝑥[𝑎,𝑏].

Proof. By the reproducing property, we have 𝐾𝑥(𝑦)=𝐾𝑥(𝜉),𝐾𝑦(𝜉)=𝐾𝑦(𝜉),𝐾𝑥(𝜉)=𝐾𝑦(𝑥) for each 𝑥 and 𝑦. Now, let 𝐾1𝑥(𝑦) and 𝐾2𝑥(𝑦) be all the reproducing kernels of the space 𝑊22[𝑎,𝑏]; then 𝐾1𝑥(𝑦)=𝐾1𝑥(𝜉),𝐾2𝑦(𝜉)=𝐾2𝑦(𝜉),𝐾1𝑥(𝜉)=𝐾2𝑦(𝑥)=𝐾2𝑥(𝑦). Finally, we note that 𝐾𝑥(𝑥)=𝐾𝑥(𝜉),𝐾𝑥(𝜉)=𝐾𝑥(𝜉)20.

3. Introduction to a Linear Operator and a Normal Orthogonal System in 𝑊22[𝑎,𝑏]

In this section, we construct an orthogonal function system of 𝑊22[𝑎,𝑏]. Also, representation of the solution of (1.1) and (1.2) is given in the reproducing kernel space 𝑊22[𝑎,𝑏].

To do this, we define a differential operator 𝐿𝑊22[𝑎,𝑏]𝑊12[𝑎,𝑏] such that 𝐿𝑢(𝑥)=𝑢(𝑥). After homogenization of the initial condition (1.2), IDEs (1.1) and (1.2) can be converted into the equivalent form as follows: 𝑢𝐿𝑢(𝑥)=𝐹(𝑥,𝑢(𝑥),𝑇𝑢(𝑥),𝑆𝑢(𝑥)),𝑎𝑥𝑏,(𝑎)=0,(3.1) such that 𝐹(𝑥,𝑢(𝑥),𝑇𝑢(𝑥),𝑆𝑢(𝑥))=𝑔(𝑥)𝑓(𝑥)𝑁(𝑥,𝑢(𝑥))𝑇𝑢(𝑥)𝑆𝑢(𝑥), 𝑇𝑢(𝑥)=𝑏𝑎𝑘1(𝑥,𝑡)𝐺1(𝑢(𝑡))𝑑𝑡, and 𝑆𝑢(𝑥)=𝑥𝑎𝑘2(𝑥,𝑡)𝐺2(𝑢(𝑡))𝑑𝑡, where 𝑢(𝑥)𝑊22[𝑎,𝑏] and 𝐹(𝑥,𝑤,𝑦,𝑧)𝑊12[𝑎,𝑏] for 𝑥[𝑎,𝑏] and 𝑤=𝑤(𝑥),𝑦=𝑦(𝑥),𝑧=𝑧(𝑥)𝑊22[𝑎,𝑏], <𝑤,𝑦,𝑧<. It is easy to show that 𝐿 is a bounded linear operator from 𝑊22[𝑎,𝑏] to 𝑊12[𝑎,𝑏].

Now, we construct an orthogonal function system of 𝑊22[𝑎,𝑏]. Let 𝜑𝑖(𝑥)=𝑇𝑥𝑖(𝑥) and 𝜓𝑖(𝑥)=𝐿𝜑𝑖(𝑥), where {𝑥𝑖}𝑖=1 is dense on [𝑎,𝑏] and 𝐿 is the adjoint operator of 𝐿. From the properties of the reproducing kernel 𝑇𝑥(𝑦), we have 𝑢(𝑥),𝜑𝑖(𝑥)𝑊12=𝑢(𝑥),𝑇𝑥𝑖(𝑥)𝑊12=𝑢(𝑥𝑖) for every 𝑢(𝑥)𝑊12[𝑎,𝑏]. In terms of the properties of 𝐾𝑥(𝑦), one obtains 𝑢(𝑥),𝜓𝑖(𝑥)𝑊22=𝑢(𝑥),𝐿𝜑𝑖(𝑥)𝑊22=𝐿𝑢(𝑥),𝜑𝑖(𝑥)𝑊12=𝐿𝑢(𝑥𝑖), 𝑖=1,2,.

It is easy to see that 𝜓𝑖(𝑥)=𝐿𝜑𝑖(𝑥)=𝐿𝜑𝑖(𝑥),𝐾𝑥(𝑦)𝑊22=𝜑𝑖(𝑥),𝐿𝑦𝐾𝑥(𝑦)𝑊12=𝐿𝑦𝐾𝑥(𝑦)|𝑦=𝑥𝑖. Thus, 𝜓𝑖(𝑥) can be expressed in the form 𝜓𝑖(𝑥)=𝐿𝑦𝐾𝑥(𝑦)|𝑦=𝑥𝑖, where 𝐿𝑦 indicates that the operator 𝐿 applies to the function of 𝑦.

Theorem 3.1. For (3.1), if {𝑥𝑖}𝑖=1 is dense on [𝑎,𝑏], then {𝜓𝑖(𝑥)}𝑖=1 is the complete function system of the space 𝑊22[𝑎,𝑏].

Proof. Clearly, 𝜓𝑖(𝑥)𝑊22[𝑎,𝑏]. For each fixed 𝑢(𝑥)𝑊22[𝑎,𝑏], let 𝑢(𝑥),𝜓𝑖(𝑥)𝑊22=0, 𝑖=1,2,, which means that 𝑢(𝑥),𝜓𝑖(𝑥)𝑊22=𝑢(𝑥),𝐿𝜑𝑖(𝑥)𝑊22=𝐿𝑢(𝑥),𝜑𝑖(𝑥)𝑊12=𝐿𝑢(𝑥𝑖)=0. Note that {𝑥𝑖}𝑖=1 is dense on [𝑎,𝑏]; therefore, 𝐿𝑢(𝑥)=0. It follows that 𝑢(𝑥)=0 from the existence of 𝐿1. So, the proof of the theorem is complete.

The orthonormal function system {𝜓𝑖(𝑥)}𝑖=1 of the space 𝑊22[𝑎,𝑏] can be derived from Gram-Schmidt orthogonalization process of {𝜓𝑖(𝑥)}𝑖=1 as follows: 𝜓𝑖(𝑥)=𝑖𝑘=1𝛽𝑖𝑘𝜓𝑘(𝑥),(3.2) where 𝛽𝑖𝑘 are orthogonalization coefficients given as 𝛽11=1/𝜓1, 𝛽𝑖𝑖=1/𝑑𝑖𝑘, and 𝛽𝑖𝑗=(1/𝑑𝑖𝑘)𝑖1𝑘=𝑗𝑐𝑖𝑘𝛽𝑘𝑗 for 𝑗<𝑖 in which 𝑑𝑖𝑘=𝜓𝑖2𝑖1𝑘=1𝑐2𝑖𝑘, 𝑐𝑖𝑘=𝜓𝑖,𝜓𝑘𝑊22, and {𝜓𝑖(𝑥)}𝑖=1 is the orthonormal system in the space 𝑊22[𝑎,𝑏].

Theorem 3.2. For each 𝑢(𝑥)𝑊22[𝑎,𝑏], the series 𝑖=1𝑢(𝑥),𝜓𝑖(𝑥)𝜓𝑖(𝑥) is convergent in the norm of 𝑊22[𝑎,𝑏]. On the other hand, if {𝑥𝑖}𝑖=1 is dense on [𝑎,𝑏] and 𝑢(𝑥) is the exact solution of (3.1), then 𝑢(𝑥)=𝑖𝑖=1𝑘=1𝛽𝑖𝑘𝐹𝑥𝑘𝑥,𝑢𝑘𝑥,𝑇𝑢𝑘𝑥,𝑆𝑢𝑘𝜓𝑖(𝑥).(3.3)

Proof. Since 𝑢(𝑥)𝑊22[𝑎,𝑏], 𝑖=1𝑢(𝑥),𝜓𝑖(𝑥)𝜓𝑖(𝑥) is the Fourier series expansion about normal orthogonal system {𝜓𝑖(𝑥)}𝑖=1, and 𝑊22[𝑎,𝑏] is the Hilbert space, then the series 𝑖=1𝑢(𝑥),𝜓𝑖(𝑥)𝜓𝑖(𝑥) is convergent in the sense of 𝑊22. On the other hand, using (3.2), we have 𝑢(𝑥)=𝑖=1𝑢(𝑥),𝜓𝑖(𝑥)𝑊22𝜓𝑖(𝑥)=𝑖=1𝑢(𝑥),𝑖𝑘=1𝛽𝑖𝑘𝜓𝑘(𝑥)𝑊22𝜓𝑖=(𝑥)𝑖𝑖=1𝑘=1𝛽𝑖𝑘𝑢(𝑥),𝜓𝑘(𝑥)𝑊22𝜓𝑖(𝑥)=𝑖𝑖=1𝑘=1𝛽𝑖𝑘𝑢(𝑥),𝐿𝜑𝑘(𝑥)𝑊22𝜓𝑖=(𝑥)𝑖𝑖=1𝑘=1𝛽𝑖𝑘𝐿𝑢(𝑥),𝜑𝑘(𝑥)𝑊12𝜓𝑖(𝑥)=𝑖𝑖=1𝑘=1𝛽𝑖𝑘𝑥𝐿𝑢𝑘𝜓𝑖(𝑥).(3.4) But since 𝑢(𝑥) is the exact solution of (3.1), then 𝐿𝑢(𝑥)=𝐹(𝑥,𝑢(𝑥),𝑇𝑢(𝑥),𝑆𝑢(𝑥)) and 𝑢(𝑥)=𝑖𝑖=1𝑘=1𝛽𝑖𝑘𝐹𝑥𝑘𝑥,𝑢𝑘𝑥,𝑇𝑢𝑘𝑥,𝑆𝑢𝑘𝜓𝑖(𝑥).(3.5) So, the proof of the theorem is complete.

Note that we denote to the approximate solution of 𝑢(𝑥) by 𝑢𝑛(𝑥)=𝑛𝑖𝑖=1𝑘=1𝛽𝑖𝑘𝐹𝑥𝑘𝑥,𝑢𝑘𝑥,𝑇𝑢𝑘𝑥,𝑆𝑢𝑘𝜓𝑖(𝑥).(3.6)

Theorem 3.3. If 𝑢(𝑥)𝑊22[𝑎,𝑏], then there exists 𝑀>0 such that ||𝑢(𝑖)(𝑥)||𝐶𝑀||𝑢(𝑥)||𝑊22, 𝑖=0,1, where ||𝑢(𝑥)||𝐶=max𝑎𝑥𝑏|𝑢(𝑥)|.

Proof. For any 𝑥,𝑦[𝑎,𝑏], we have 𝑢(𝑖)(𝑥)=𝑢(𝑦),𝐾𝑥(𝑖)(𝑦)𝑊22,𝑖=0,1. By the expression of 𝐾𝑥(𝑦), it follows that 𝐾𝑥(𝑖)(𝑦)𝑊22𝑀𝑖, 𝑖=0,1. Thus, |𝑢(𝑖)(𝑥)|=|𝑢(𝑥),𝐾𝑥(𝑖)(𝑥)𝑊22|𝑢(𝑥)𝑊22𝐾𝑥(𝑖)(𝑥)𝑊22𝑀𝑖𝑢(𝑥)𝑊22, 𝑖=0,1. Hence, ||𝑢(𝑖)(𝑥)||𝐶𝑀||𝑢(𝑥)||𝑊22,  𝑖=0,1, where 𝑀=max{𝑀0,𝑀1}. The proof is complete.

Corollary 3.4. The approximate solution 𝑢𝑛(𝑥) and its derivative 𝑢𝑛(𝑥) are uniformly convergent.

Proof. By Theorems 3.2 and 3.3, for any 𝑥[𝑎,𝑏], we get ||𝑢𝑛||=|||(𝑥)𝑢(𝑥)𝑢𝑛(𝑥)𝑢(𝑥),𝐾𝑥(𝑥)𝑊22|||𝐾𝑥(𝑥)𝑊22𝑢𝑛(𝑥)𝑢(𝑥)𝑊22𝑀0𝑢𝑛(𝑥)𝑢(𝑥)𝑊22.(3.7)
On the other hand, ||𝑢𝑛(𝑥)𝑢||=|||𝑢(𝑥)𝑛(𝑥)𝑢(𝑥),𝐾𝑥(𝑥)𝑊22|||𝐾𝑥(𝑥)𝑊22𝑢𝑛(𝑥)𝑢(𝑥)𝑊22𝑀1𝑢𝑛(𝑥)𝑢(𝑥)𝑊22.(3.8)
Hence, |𝑢𝑛(𝑖)(𝑥)𝑢(𝑖)(𝑥)|𝑢𝑛(𝑖)(𝑥)𝑢(𝑖)(𝑥)𝐶𝑀𝑖𝑢𝑛(𝑥)𝑢(𝑥)𝑊22, where 𝑀0 and 𝑀1 are positive constants. Hence, if 𝑢𝑛(𝑥)𝑢(𝑥)𝑊220 as 𝑛, the approximate solutions 𝑢𝑛(𝑥) and 𝑢𝑛(𝑥) converge uniformly to the exact solution 𝑢(𝑥) and its derivative, respectively.

4. Iterative Method and Convergence Theorem

In this section, an iterative method of obtaining the solution of (3.1) is presented in the reproducing kernel space 𝑊22[𝑎,𝑏].

First of all, we will mention the following remark in order to solve (1.1) and (1.2) numerically. If (1.1) is linear, then the exact and approximate solutions can be obtained directly from (3.5) and (3.6), respectively. On the other hand, if (1.1) is nonlinear, then the exact and approximate solutions can be obtained using the following iterative method.

According to (3.5), the representation of the solution of (1.1) can be denoted by 𝑢(𝑥)=𝑖=1𝐴𝑖𝜓𝑖(𝑥),(4.1) where 𝐴𝑖=𝑖𝑘=1𝛽𝑖𝑘𝐹(𝑥𝑘,𝑢(𝑥𝑘),𝑇𝑢(𝑥𝑘),𝑆𝑢(𝑥𝑘)). In fact, 𝐴𝑖, 𝑖=1,2,, in (4.1) are unknown, and we will approximate 𝐴𝑖 using known 𝐵𝑖. For a numerical computation, we define initial function 𝑢0(𝑥1)=0 and the 𝑛-term approximation to 𝑢(𝑥) by 𝑢𝑛(𝑥)=𝑛𝑖=1𝐵𝑖𝜓𝑖(𝑥),(4.2) where the coefficients 𝐵𝑖 are given as 𝐵1=𝛽11𝐹𝑥1,𝑢0𝑥1,𝑇𝑢0𝑥1,𝑆𝑢0𝑥1,𝑢1(𝑥)=𝐵1𝜓1𝐵(𝑥),2=2𝑘=1𝛽2𝑘𝐹𝑥𝑘,𝑢𝑘1𝑥𝑘,𝑇𝑢𝑘1𝑥𝑘,𝑆𝑢𝑘1𝑥𝑘,𝑢2(𝑥)=2𝑖=1𝐵𝑖𝜓𝑖𝑢(𝑥),𝑛1(𝑥)=𝑛1𝑖=1𝐵𝑖𝜓𝑖(𝐵𝑥),𝑛=𝑛𝑘=1𝛽𝑛𝑘𝐹𝑥𝑘,𝑢𝑘1𝑥𝑘,𝑇𝑢𝑘1𝑥𝑘,𝑆𝑢𝑘1𝑥𝑘.(4.3)

We mention here the following remark: in the iterative process of (4.2), we can guarantee that the approximation 𝑢𝑛(𝑥) satisfies the initial condition (1.2).

Now, the approximate solution 𝑢𝑁𝑛(𝑥) can be obtained by taking finitely many terms in the series representation of 𝑢𝑛(𝑥) and 𝑢𝑁𝑛(𝑥)=𝑁𝑖𝑖=1𝑘=1𝛽𝑖𝑘𝐹𝑥𝑘,𝑢𝑛1𝑥𝑘,𝑇𝑢𝑛1𝑥𝑘,𝑆𝑢𝑛1𝑥𝑘𝜓𝑖(𝑥).(4.4)

Next, we will prove that 𝑢𝑛(𝑥) in the iterative formula (4.2) is convergent to the exact solution 𝑢(𝑥) of (1.1).

Lemma 4.1. If 𝑢𝑛(𝑥)𝑢(𝑥)𝑊220, 𝑥𝑛𝑦 as 𝑛 and 𝐹(𝑥,𝑣,𝑤,𝑧) is continuous in [𝑎,𝑏] with respect to 𝑥,𝑣,𝑤,𝑧, for 𝑥[𝑎,𝑏] and 𝑣,𝑤,𝑧(,), then as 𝑛, one has 𝐹(𝑥𝑛,𝑢𝑛1(𝑥𝑛),𝑇𝑢𝑛1(𝑥𝑛),𝑆𝑢𝑛1(𝑥𝑛))𝐹(𝑦,𝑢(𝑦),𝑇𝑢(𝑦),𝑆𝑢(𝑦)).

Proof. Firstly, we will prove that 𝑢𝑛1(𝑥𝑛)𝑢(𝑦) in the sense of 𝑊22. Since ||𝑢𝑛1𝑥𝑛||=||𝑢𝑢(𝑦)𝑛1𝑥𝑛𝑢𝑛1(𝑦)+𝑢𝑛1||||𝑢(𝑦)𝑢(𝑦)𝑛1𝑥𝑛𝑢𝑛1(||+||𝑢𝑦)𝑛1(||,𝑦)𝑢(𝑦)(4.5) by reproducing kernel property of 𝐾𝑥(𝑦), we have 𝑢𝑛1(𝑥𝑛) = 𝑢𝑛1(𝑥),𝐾𝑥𝑛(𝑥) and 𝑢𝑛1(𝑦)=𝑢𝑛1(𝑥),𝐾𝑦(𝑥). Thus, |𝑢𝑛1(𝑥𝑛)𝑢𝑛1(𝑦)| = |𝑢𝑛1(𝑥),𝐾𝑥𝑛(𝑥)𝐾𝑦(𝑥)𝑊22|𝑢𝑛1(𝑥)𝑊22𝐾𝑥𝑛(𝑥)𝐾𝑦(𝑥)𝑊22. From the symmetry of 𝐾𝑥(𝑦), it follows that 𝐾𝑥𝑛(𝑥)𝐾𝑦(𝑥)𝑊220 as 𝑛. Hence, |𝑢𝑛1(𝑥𝑛)𝑢𝑛1(𝑦)|0 as soon as 𝑥𝑛𝑦.
On the other hand, by Corollary 3.4, for any 𝑦[𝑎,𝑏], it holds that |𝑢𝑛1(𝑦)𝑢(𝑦)|0. Therefore, 𝑢𝑛1(𝑥𝑛)𝑢(𝑦) in the sense of 𝑊22 as 𝑥𝑛𝑦 and 𝑛.
Thus, by means of the continuation of 𝐺1, 𝐺2, and 𝑁, it is obtained that 𝐺1(𝑢𝑛1(𝑥𝑛))𝐺1(𝑢(𝑦)), 𝐺2(𝑢𝑛1(𝑥𝑛))𝐺2(𝑢(𝑦)), and 𝑁(𝑥𝑛,𝑢𝑛1(𝑥𝑛))𝑁(𝑦,𝑢(𝑦)) as 𝑛. This shows that 𝑇𝑢𝑛1(𝑥𝑛)𝑇𝑢(𝑦) and 𝑆𝑢𝑛1(𝑥𝑛)𝑆𝑢(𝑦) as 𝑛. Hence, the continuity of 𝐹 gives the result.

Lemma 4.2. {𝑢𝑛}𝑛=1 in (4.2) is monotonically increasing in the sense of the norm of 𝑊22[𝑎,𝑏].

Proof. By Theorem 3.1, {𝜓𝑖}𝑖=1 is the complete orthonormal system in the space 𝑊22[𝑎,𝑏]. Hence, we have 𝑢𝑛2𝑊22=𝑢𝑛(𝑥),𝑢𝑛(𝑥)𝑊22=𝑛𝑖=1𝐵𝑖𝜓𝑖(𝑥),𝑛𝑖=1𝐵𝑖𝜓𝑖(𝑥)𝑊22=𝑛𝑖=1(𝐵𝑖)2. Therefore, 𝑢𝑛𝑊22 is monotonically increasing.

Lemma 4.3. One has 𝐿𝑢𝑛(𝑥𝑗)=𝐹(𝑥𝑗,𝑢𝑗1(𝑥𝑗),𝑇𝑢𝑗1(𝑥𝑗),𝑆𝑢𝑗1(𝑥𝑗)),𝑗𝑛.

Proof. The proof will be obtained by mathematical induction as follows: if 𝑗𝑛, then 𝐿𝑢𝑛(𝑥𝑗) = 𝑛𝑖=1𝐵𝑖𝐿𝜓𝑖(𝑥𝑗) = 𝑛𝑖=1𝐵𝑖𝐿𝜓𝑖(𝑥),𝜑𝑗(𝑥)𝑊12 = 𝑛𝑖=1𝐵𝑖𝜓𝑖(𝑥),𝐿𝑗𝜑(𝑥)𝑊22 = 𝑛𝑖=1𝐵𝑖𝜓𝑖(𝑥),𝜓𝑗(𝑥)𝑊22. Thus, 𝐿𝑢𝑛𝑥𝑗=𝑛𝑖=1𝐵𝑖𝜓𝑖(𝑥),𝜓𝑗(𝑥)𝑊22.(4.6)
Multiplying both sides of (4.6) by 𝛽𝑗𝑙, summing for 𝑙 from 1 to 𝑗, and using the orthogonality of {𝜓𝑖(𝑥)}𝑖=1 yield that 𝑗𝑙=1𝛽𝑗𝑙𝐿𝑢𝑛𝑥𝑙=𝑛𝑖=1𝐵𝑖𝜓𝑖(𝑥),𝑗𝑙=1𝛽𝑗𝑙𝜓𝑙(𝑥)𝑊22=𝑛𝑖=1𝐵𝑖𝜓𝑖(𝑥),𝜓𝑗(𝑥)𝑊22=𝐵𝑗=𝑗𝑙=1𝛽𝑗𝑙𝐹𝑥𝑙,𝑢𝑙1𝑥𝑙,𝑇𝑢𝑙1𝑥𝑙,𝑆𝑢𝑙1𝑥𝑙.(4.7)
Now, if 𝑗=1, then 𝐿𝑢𝑛(𝑥1)=𝐹(𝑥1,𝑢0(𝑥1),𝑇𝑢0(𝑥1),𝑆𝑢0(𝑥1)). On the other hand, if 𝑗=2, then 𝛽21𝐿𝑢𝑛(𝑥1)+𝛽22𝐿𝑢𝑛(𝑥2)=𝛽21𝐹(𝑥1,𝑢0(𝑥1),𝑇𝑢0(𝑥1),𝑆𝑢0(𝑥1))+𝛽22𝐹(𝑥2,𝑢1(𝑥2),𝑇𝑢1(𝑥2),𝑆𝑢1(𝑥2)). Thus, 𝐿𝑢𝑛(𝑥2)=𝐹(𝑥2,𝑢1(𝑥2),𝑇𝑢1(𝑥2),𝑆𝑢1(𝑥2)). It is easy to see that 𝐿𝑢𝑛(𝑥𝑗)=𝐹(𝑥𝑗,𝑢𝑗1(𝑥𝑗),𝑇𝑢𝑗1(𝑥𝑗),𝑆𝑢𝑗1(𝑥𝑗)) by using mathematical induction.

Lemma 4.4. One has 𝐿𝑢𝑛(𝑥𝑗)=𝐿𝑢(𝑥𝑗),𝑗𝑛.

Proof. It is clear that on taking limits in (4.2) 𝑢(𝑥)=𝑖=1𝐵𝑖𝜓𝑖(𝑥). Therefore, 𝑢𝑛(𝑥)=𝑃𝑛𝑢(𝑥), where 𝑃𝑛 is an orthogonal projector from 𝑊22[𝑎,𝑏] to Span{𝜓1,𝜓2,,𝜓𝑛}. Thus, 𝐿𝑢𝑛(𝑥𝑗) = 𝐿𝑢𝑛(𝑥),𝜑𝑗(𝑥)𝑊12 = 𝑢𝑛(𝑥),𝐿𝑗𝜑(𝑥)𝑊22 = 𝑃𝑛𝑢(𝑥),𝜓𝑗(𝑥)𝑊22 = 𝑢(𝑥),𝑃𝑛𝜓𝑗(𝑥)𝑊22 = 𝑢(𝑥),𝜓𝑗(𝑥)𝑊22 = 𝐿𝑢(𝑥),𝜑𝑗(𝑥)𝑊12=𝐿𝑢(𝑥𝑗).

Theorem 4.5. Suppose that 𝑢𝑛𝑊22 is bounded in (4.2). If {𝑥𝑖}𝑖=1 is dense on [𝑎,𝑏], then the 𝑛-term approximate solution 𝑢𝑛(𝑥) in the iterative formula (4.2) converges to the exact solution 𝑢(𝑥) of (3.1) in the space 𝑊22[𝑎,𝑏] and 𝑢(𝑥)=𝑖=1𝐵𝑖𝜓𝑖(𝑥), where 𝐵𝑖 is given by (4.3).

Proof. First of all, we will prove the convergence of 𝑢𝑛(𝑥). From (4.2), we infer that 𝑢𝑛+1(𝑥)=𝑢𝑛(𝑥)+𝐵𝑛+1𝜓𝑛+1(𝑥). The orthogonality of {𝜓𝑖(𝑥)}𝑖=1 yields that ||||𝑢𝑛+1||||2𝑊22=||||𝑢𝑛||||2𝑊22+𝐵𝑛+12=||||𝑢𝑛1||||2𝑊22+𝐵𝑛2+𝐵𝑛+12||||𝑢==0||||2𝑊22+𝑛+1𝑖=1𝐵𝑖2.(4.8)
From Lemma 4.2, the sequence 𝑢𝑛𝑊22 is monotonically increasing. Due to the condition that 𝑢𝑛𝑊22 is bounded, 𝑢𝑛𝑊22 is convergent as 𝑛. Then, there exists a constant 𝑐 such that 𝑖=1(𝐵𝑖)2=𝑐. This implies that 𝐵𝑖=𝑖𝑘=1𝛽𝑖𝑘𝐹(𝑥𝑘,𝑢𝑘1(𝑥𝑘),𝑇𝑢𝑘1(𝑥𝑘),𝑆𝑢𝑘1(𝑥𝑘))𝑙2,  𝑖=1,2,.
If 𝑚>𝑛, using (𝑢𝑚𝑢𝑚1)(𝑢𝑚1𝑢𝑚2)(𝑢𝑛+1𝑢𝑛), then one gets ||||𝑢𝑚(𝑥)𝑢𝑛||||(𝑥)2𝑊22=||||𝑢𝑚(𝑥)𝑢𝑚1(𝑥)+𝑢𝑚1(𝑥)+𝑢𝑛+1(𝑥)𝑢𝑛||||(𝑥)2𝑊22=||||𝑢𝑚(𝑥)𝑢𝑚1||||(𝑥)2𝑊22||||𝑢++𝑛+1(𝑥)𝑢𝑛||||(𝑥)2𝑊22.(4.9) Furthermore, ||𝑢𝑚(𝑥)𝑢𝑚1(𝑥)||2𝑊22=(𝐵𝑚)2. Consequently, ||𝑢𝑚(𝑥)𝑢𝑛(𝑥)||2𝑊22=𝑚𝑖=𝑛+1(𝐵𝑖)20 as 𝑛,𝑚. Considering the completeness of the space 𝑊22[𝑎,𝑏], there exists a 𝑢(𝑥)𝑊22[𝑎,𝑏] such that 𝑢𝑛(𝑥)𝑢(𝑥) as 𝑛 in the sense of 𝑊22.
Secondly, we will prove that 𝑢(𝑥) is the solution of (3.1). Since {𝑥𝑖}𝑖=1 is dense on [𝑎,𝑏], for any 𝑥[𝑎,𝑏], there exists subsequence {𝑥𝑛𝑗}𝑗=1 such that 𝑥𝑛𝑗𝑥 as 𝑗. From Lemmas 4.3 and 4.4, it is easy to see that 𝐿𝑢(𝑥𝑛𝑗)=𝐹(𝑥𝑛𝑗,𝑢𝑛𝑗1(𝑥𝑘),𝑇𝑢𝑛𝑗1(𝑥𝑘),𝑆𝑢𝑛𝑗1(𝑥𝑘)). Hence, letting 𝑗, by Lemma 4.1 and the continuity of 𝐹, we have 𝐿𝑢(𝑥)=𝐹(𝑥,𝑢(𝑥),𝑇𝑢(𝑥),𝑆𝑢(𝑥)). That is, 𝑢(𝑥) is the solution of (3.1).
Since 𝜓𝑖(𝑥)𝑊22[𝑎,𝑏], clearly, 𝑢(𝑥) satisfies the initial condition (1.2). In other words, 𝑢(𝑥) is the solution of (1.1) and (1.2), where 𝑢(𝑥)=𝑖=1𝐵𝑖𝜓𝑖(𝑥) and 𝐵𝑖 is given by (4.3). The proof is complete.

Theorem 4.6. Assume that 𝑢(𝑥) is the solution of (3.1) and 𝑟𝑛(𝑥) is the difference between the approximate solution 𝑢𝑛(𝑥) and the exact solution 𝑢(𝑥). Then, 𝑟𝑛(𝑥) is monotonically decreasing in the sense of the norm of 𝑊22[𝑎,𝑏].

Proof. It obvious that ||𝑟𝑛(𝑥)||2𝑊22=||𝑢(𝑥)𝑢𝑛(𝑥)||2𝑊22=𝑖=𝑛+1𝐵𝑖𝜓𝑖(𝑥)2𝑊22=𝑖=𝑛+1(𝐵𝑖)2 and ||𝑟𝑛1(𝑥)||2𝑊22=𝑖=𝑛(𝐵𝑖)2. Thus, ||𝑟𝑛(𝑥)||𝑊22||𝑟𝑛1(𝑥)||𝑊22; consequently, the difference 𝑟𝑛(𝑥) is monotonically decreasing in the sense of 𝑊22. So, the proof of the theorem is complete.

5. Numerical Examples

In this section, some numerical examples are studied to demonstrate the accuracy and applicability of the present method. Results obtained are compared with the exact solution of each example and are found to be in good agreement with each other. In the process of computation, all the symbolic and numerical computations were performed by using Mathematica 7.0 software package.

Example 5.1. Consider the nonlinear Fredholm-Volterra IDE: 𝑢1(𝑥)+𝑢(𝑥)410𝑡𝑢31(𝑡)𝑑𝑡+2𝑥0𝑥𝑢2(𝑡)𝑑𝑡=𝑔(𝑥),0𝑥,𝑡1,𝑢(0)=0,(5.1) where 𝑔(𝑥)=(1/10)𝑥6+𝑥2+2𝑥(1/32). The exact solution is 𝑢(𝑥)=𝑥2.

Using RKHS method, taking 𝑥𝑖=(𝑖1)/(𝑁1), 𝑖=1,2,,𝑁, with the reproducing kernel function 𝐾𝑥(𝑦) on [0,1], the approximate solution 𝑢𝑁𝑛(𝑥) is calculated by (4.4). The numerical results at some selected grid points for 𝑁=26 and 𝑛=5 are given in Table 1.

As we mention, we used the grid nodes mentioned earlier in order to obtain approximate solutions. Moreover, it is possible to pick any point in [𝑎,𝑏] and as well the approximate solution and its derivative will be applicable. Next, the numerical results for Example 5.1 at some selected gird nodes in [0,1] of 𝑢(𝑥) are given in Table 2.

Table 3 shows, a comparison between the absolute errors of our method together with triangular functions method [6], operational matrix with block-pulse functions method [7], and Hybrid Legendre polynomials and block-pulse functions method [8]. As it is evident from the comparison results, it was found that our method in comparison with the mentioned methods is better with a view to accuracy and utilization.

Example 5.2. Consider the nonlinear Fredholm-Volterra IDE: 𝑢(𝑥)+2𝑥𝑢(𝑥)10(𝑥𝑡)𝑢(𝑡)𝑑𝑡𝑥0(𝑥+𝑡)𝑢3(𝑡)𝑑𝑡=𝑔(𝑥),0𝑥,𝑡1,𝑢(0)=1,(5.2) where 𝑔(𝑥)=((2/3)𝑥+(1/9))𝑒3𝑥+(2𝑥+1)𝑒𝑥+((4/3)𝑒)𝑥+(8/9). The exact solution is 𝑢(𝑥)=𝑒𝑥.

Using RKHS method, taking 𝑥𝑖=(𝑖1)/(𝑁1), 𝑖=1,2,,𝑁, with the reproducing kernel function 𝐾𝑥(𝑦) on [0,1], the approximate solution 𝑢𝑁𝑛(𝑥) is calculated by (4.4). The numerical results at some selected grid points for 𝑁=26 and 𝑛=1 are given in Table 4.

The comparison among the RKHS solution besides the solutions of triangular functions [6], operational matrix with block-pulse functions solution [7], and exact solutions are shown in Table 5.

Example 5.3. Consider the nonlinear Fredholm-Volterra IDE: 𝑢(𝑥)+𝑢2(𝑥)10(𝑡+1)sinh(𝑢(𝑡)1)𝑑𝑡𝑥0𝑥𝑒𝑢(𝑡)𝑑𝑡=𝑔(𝑥),0𝑥,𝑡1,𝑢(0)=1,(5.3) where 𝑔(𝑥)=(ln(𝑥+1)+1)2+(𝑥+1)1(𝑒/2)(𝑥+2)𝑥2(2/3). The exact solution is 𝑢(𝑥)=1+ln(𝑥+1).

Using RKHS method, taking 𝑥𝑖=(𝑖1)/(𝑁1), 𝑖=1,2,,𝑁, with the reproducing kernel function 𝐾𝑥(𝑦) on [0,1], the approximate solution 𝑢𝑁𝑛(𝑥) is calculated by (4.4). The numerical results at some selected grid points for 𝑁=51 and 𝑛=1 are given in Table 6.

Example 5.4. Consider the nonlinear Fredholm-Volterra IDE: 𝑢(𝑥)+cos(𝑢(𝑥))10(𝑡+𝑥)𝑢(𝑡)𝑑𝑡𝑥0𝑡𝑒𝑢(𝑡)𝑑𝑡=𝑔(𝑥),0𝑥,𝑡1,𝑢(0)=1,(5.4) where 𝑔(𝑥)=cos(1𝑥)+𝑒1𝑥(1+𝑥)(1/2)𝑥(7/6)𝑒. The exact solution is 𝑢(𝑥)=1𝑥.

Using RKHS method, taking 𝑥𝑖=(𝑖1)/(𝑁1), 𝑖=1,2,,𝑁, with the reproducing kernel function 𝐾𝑥(𝑦) on [0,1], the approximate solution 𝑢𝑁𝑛(𝑥) is calculated by (4.4). The numerical results at some selected grid points for 𝑁=51 and 𝑛=5 are given in Table 7.

Example 5.5. Consider the nonlinear Fredholm-Volterra IDE: 𝑢(𝑥)+𝑥𝑢2(𝑥)10(𝑡+𝑥)1+𝑢2(𝑡)𝑑𝑡𝑥0𝑥cos(𝑢(𝑡))𝑑𝑡=𝑔(𝑥),0𝑥,𝑡1,𝑢(0)=0,(5.5) where 𝑔(𝑥)=𝑥sin(𝑥)+𝑥3(4/3)𝑥+(1/4). The exact solution is 𝑢(𝑥)=𝑥.
Using RKHS method, taking 𝑥𝑖=(𝑖1)/(𝑁1), 𝑖=1,2,,𝑁 with the reproducing kernel function 𝐾𝑥(𝑦) on [0,1], the approximate solution 𝑢𝑁𝑛(𝑥) is calculated by (4.4). The numerical results at some selected grid points for 𝑁=26 and 𝑛=5 are given in Table 8.

6. Conclusion

In this paper, the RKHS method was employed to solve the nonlinear Fredholm-Volterra IDEs (1.1) and (1.2). The solution 𝑢(𝑥) and the approximate solution 𝑢𝑛(𝑥) are represented in the form of series in the space 𝑊22[𝑎,𝑏]. Moreover, the approximate solution and its derivative converge uniformly to the exact solution and its derivative, respectively. Meanwhile, the error of the approximate solution is monotonically decreasing in the sense of the norm of 𝑊22[𝑎,𝑏].