Abstract

A second-order boundary value problem with nonlinear and mixed two-point boundary conditions is considered, , , , in which is a formally self-adjoint second-order differential operator. Under appropriate assumptions on , , and , existence and uniqueness of solutions is established by the method of upper and lower solutions and Leray-Schauder degree theory. The general quasilinearization method is then applied to this problem. Two monotone sequences converging quadratically to the unique solution are constructed.

1. Introduction

The investigation of boundary value problems (denoted as BVPs for short) of ordinary differential equations is of great significance. On one hand, it makes a great impact on the studies of partial differential equations [1]. On the other hand, BVPs of ordinary differential equations can be used to describe a large number of mechanical, physical, biological, and chemical phenomena; see [25] for example. So far a lot of work has been carried out, including second-order, third-order, and higher-order BVPs with various boundary conditions.

As far as we know, for a long term most of works focused on existence and uniqueness of solutions. The works relating to approximation of solutions are relatively rare. In recent years, some approximate methods, such as the shooting method [6], monotone iterative technique [7], homotopy analysis method [8], and general quasilinearization method have been applied to BVPs for obtaining approximations of solutions. Among these methods, the general quasilinearization method becomes more and more popular.

The quasilinearization method was originally proposed by Bellman and Kalaba [9]. It is a very powerful approximation technique and unlike perturbation methods, is not dependent on the existence of a small or large parameter. The method, whose sequence of solutions of linear problems convergences to the solution of the original nonlinear problem, is quadratic and monotone, which is one of the reasons for the popularity of this technique. This method was generalized by Lakshmikantham and Vatsala [10] in which the convexity or concavity assumption on the nonlinear functions involved in the problems is relaxed.

So far, the general quasilinearization method, coupled with the method of upper and lower solutions, has been applied to obtain approximation of solutions for a large number of nonlinear problems, for example, BVPs of ordinary differential equations, such as first-order BVP with nonlinear boundary condition [11] and second-order BVPs with Dirichlet boundary condition [12], periodic boundary condition [13], three-point boundary condition [14], four-point boundary condition [15], and -point boundary condition [16]; BVPs of partial differential equations, such as parabolic initial-boundary value problem [17], elliptic problems with nonlinear boundary condition [18] and -Laplacian equations with nonlinear boundary condition [19], and so forth; BVPs of impulsive differential equations [20] and impulse functional differential equations with anti-periodic boundary condition [21]; BVPs of some practically nonlinear problems, such as Duffing equation involving both integral and nonintegral forcing terms with Robin boundary condition [22] and forced Duffing equation with discontinuous-type integral boundary condition [23]; as well as some abstract problems such as fixed point theorems in ordered Banach space [24].

Recently, El-Gebeily and O’Regan [25, 26] consider the singular and nonsingular second-order ordinary differential equations with nonlinear boundary conditions as follows: in which is a formally self-adjoint second-order differential operator. This type of BVPs arises in a variety of problems in applied mathematics and physics [27]. By defining the upper and lower solutions of BVP (1.1)–(1.3) suitably, Gebeily and Regan established the existence and uniqueness of solutions and constructed two monotonic iterative sequences converging to the unique solution quadratically. However, it can be noted that the and terms are not involved in (1.2) and (1.3), respectively. For second-order two-point BVPs, if all the terms and are involved in the boundary conditions, it seems quite difficult to deal with.

In this paper, we consider a second-order BVP with nonlinear and mixed two-point boundary conditions as follows: where , in which , , and , .

In BVP (1.4)–(1.6), it can be found that the boundary condition (1.5) is dependent on all the , and terms. First, existence and uniqueness of solutions of BVP (1.4)–(1.6) is established by combining the method of upper and lower solutions with Leray-Schauder degree theory. Then, the general quasilinearization method is applied to construct the approximations of the unique solution. Two monotone sequences of iterations converging to the unique solution quadratically are obtained.

2. Preliminaries

In this section, several definitions and lemmas needed to the main results are given first.

Definition 2.1. are called the upper and lower solutions of BVP (1.4)–(1.6), respectively, if

Definition 2.2. Let be a subset of ; it is said that the right-hand side function of (1.4) satisfies Nagumo condition on if holds for and

Lemma 2.3 (see [28]). Let be a continuous function satisfying Nagumo condition on where are continuous functions such that for all . Then there exists a constant such that every solution of second-order equations with satisfies , in which is called the Nagumo constant.

Lemma 2.4. Boundary value problem as follows: has only the trivial solution.

Proof. Assume that is an arbitrarily nontrivial solution of BVP (2.5)-(2.6). From the boundary conditions (2.6), it can be concluded that can achieve its positive maximum or negative minimum in the interior of , suppose at , .
If achieves its positive maximum, then which means that On the other hand, it can be derived from (2.5) that It is a contradiction.
If achieves its negative minimum, similar arguments lead to a contradiction too. Hence, BVP (2.5)-(2.6) has only the trivial solution.

Lemma 2.5 (see [26]). Define a linear operator by Then exists and is continuous.

Lemma 2.6. Assume that(1) are the lower and upper solutions of BVP (1.4)–(1.6), respectively;(2) is continuous on and is strictly decreasing in on (3) is continuously differentiable on , strictly decreasing in the first and second variables, and nondecreasing and nonincreasing, respectively, in the third and forth variables.
Then,

Proof. Suppose that for some . Then there exist some such that
If , then , , and and consequently, However, it follows from Definition 2.1 and mean value theorem that in which , and the last inequality follows from the strictly decreasing property of in . It is a contradiction.
If , then and By the definitions of the lower and upper solutions, we have Moreover, Consequently, in view of the monotonicity of in its variables, it follows that which is a contradiction.
If , similar deductions lead to a contradiction too. Hence, Lemma 2.6 is proved.

3. Existence and Uniqueness of Solutions

Theorem 3.1. Assume that(1) are the lower and upper solutions of BVP (1.4)–(1.6), respectively;(2) is continuous on and is strictly decreasing in on and satisfies Nagumo condition on ;(3) is continuously differentiable, strictly decreasing in the first and second variables, nondecreasing and nonincreasing, respectively, in the third and forth variables on .

Then there exists a unique solution of BVP (1.4)–(1.6) such that

Proof. Define Introduce the following auxiliary BVPs with homotopy character: in which is called the embedded parameter.
By the continuity of and , and by the boundedness of and , we can select a sufficiently large constant such that
In what follows, the proof of the existence of solutions is divided into four steps.
Step 1. Show that, for , every solution of BVPs (3.4) satisfies
Suppose that the estimate is not true. Then, there must be some points in such that either or For the former case, has a positive maximum, suppose at , that is, can be assumed. We have three cases to consider.
Case 1 (). In this case, , and which leads to . On the other hand, since the definition of yields ; hence, for , we have in which the last inequality is obtained by the inequality (3.6). It is a contradiction.
For , it can be derived that which is also a contradiction.
Case 2 (). In this case, , Furthermore, we have The definition of means that and . Hence, for , in view of and , it follows from the monotonicity of in its third and forth variables and the inequality (3.7) that which is a contradiction.
For we can obtain from the second equation in (3.4) that which is a contradiction too.
Case 3 (). In this case, , , and furthermore, The same deductions with those in Case 2 yield, for , that It is also a contradiction. For it can be deduced from the boundary conditions of (3.4) that which is a contradiction too.
Step 2. Show that there exists a positive constant such that, for , every solution of BVPs (3.4) satisfies
Let Define Consequently, (1.4) can be rewritten as Since the function satisfies Nagumo condition on , is obvious if the boundedness of and , , , and is kept in mind. That is, satisfies Nagumo condition on . Hence, by Lemma 2.3, the estimate , , can be obtained.
Step 3. Show that for , BVP (3.4) has at least one solution .
Define a nonlinear operator by with Consequently, BVPs (3.4) are equivalent to the following operator equations: in which is the linear operator defined in Lemma 2.5, and is the unit operator.
Define the norm in as for , in which are two real numbers.
Let be the bounded sequence on . It then follows from Steps 1 and 2 that and are both uniformly bounded in the sense of the norm defined above. Thus, is equicontinuous on . Consequently, Arzela-Ascoli theorem yields that is compact on . Therefore, is a completely continuous operator.
Furthermore, the operator defined by is also a completely continuous operator.
Define a bounded and open domain as follows: It follows from Steps 1 and 2 that Therefore, the degree is well defined. Then the invariance of degree under homotopy yields
Since the operator equation is equivalent to BVP (2.5)-(2.6) which has only the trivial solution, therefore Consequently, the operator equation gives which has at least one solution
Step 4. Show that every solution of BVP (3.36) satisfies
The right-hand side of this inequality is first proved. For the sake of contradiction, suppose that for some . Define , then has a positive maximum at some .
Case 1 (). In this case, , , and and consequently, .
On the other hand, which is a contradiction.
Case 2 (). In this case, Furthermore, we have The definition of yields and . In view of and , it then follows from the monotonicity of in its third and forth variables that It is also a contradiction.Case 3 (). In this case, Furthermore, Similar deductions lead to a contradiction too.
Therefore, the inequality holds. In a similar way, can be proved.
Till now, by Steps 14, the proof of the existence of solutions is completed. In what follows, we turn to the proof of the uniqueness of solutions.
We may assume that and are two arbitrary solutions of BVP (1.4)–(1.6). Define If for , the uniqueness of solutions is obvious. Otherwise, there must be some points in such that either or . For the former case, we define Similarly, we only need to consider the following three cases.
Case 1 (). In this case, Thus, . On the other hand, the mean value theorem and the strictly decreasing property of in results that which is a contradiction, where .Case 2 (). In this case, Moreover, Nevertheless, in view of the monotonicity of in its variables, it follows from the mean value theorem that which is a contradiction, in which , , and can be located by analogy.Case 3 (). For this case, in the same way, it can be proved that this case is also impossible.
Consequently, the conclusion in Theorem 3.1 is proved.

4. Approximations of the Unique Solution

In this section, a series of boundary value problems whose sequence of solutions converging to the unique solution of BVP (1.4)–(1.6) is constructed.

Theorem 4.1. Assume that (1)there exist the upper and lower solutions of BVP (1.4)–(1.6);(2) is continuous on and satisfies Nagumo condition on (3) exists and is continuous on and , on , where is chosen sufficiently large, and is the Nagumo constant;(4), and it is strictly decreasing in and , no-decreasing and nonincreasing, respectively, in and .Then there exist a monotone nonincreasing sequence and a monotone nondecreasing sequence converging uniformly to the unique solution of BVP (1.4)–(1.6). Moreover, the rate of the convergence is quadratic.

Proof. Define a piecewise function for , and introduce on and for .
Obviously, is linear in if is viewed as a known quantity. Moreover, and both satisfy the Nagumo condition.
We begin by considering the following BVP: In view of the definition of the upper solution of BVP (1.4)–(1.6), it can be verified directly that Hence, is the upper solution of BVP (4.4).
According to condition , it can be deduced that in which has been taken into account, and . Consequently, by taking the definition of the lower solution of BVP (1.4)–(1.6) into account, we have which indicates that is the lower solution of BVP (4.4). Since BVP (4.4) satisfies all the requirements in Theorem 3.1, hence BVP (4.4) has a unique solution with and , , where is a constant.
Next, we consider It can be shown that in which . Thus, is the upper solution of BVP (4.8) since are obvious if we keep that is the solution of BVP (4.4) in mind.
Similarly, we can show that is the lower solution of BVP (4.8). Hence, according to Theorem 3.1, there exists a unique solution of BVP (4.8) such that and , , where is a constant.
By repeating the above procedure iteratively, we obtain a monotone nonincreasing sequence with where are, respectively, the unique solution of the BVPs as follows:
Formula (4.11), the properties of the operator , and the continuity of yield the uniform boundedness of for . Therefore, the sequences are uniformly bounded and equicontinuous on . Hence, Ascoli-Arzela theorem guarantees the existence of subsequences and a function with converging uniformly to on as .
Now let The continuity of the functions and imply that, as , The Lebeague dominated convergence theorem then yields for each as
Integrating the equation from to results in Passing to the limit in (4.17), we obtain Hence, is the unique solution of BVP (1.4)–(1.6).
To show the quadratic rate of convergence, we define the error function as follows: Hence, for in which lies between and , , is a negative constant, is a constant with for , and .
Moreover, since are the solutions of BVPs (4.12) and BVP (1.4)–(1.6), respectively, therefore, mean value theorem yields in which , , where , and so forth.
Define a constant function by Then, satisfies and moreover, since is a nonnegative constant function, therefore, we have and , . Consequently, the condition () guarantees that in which , , are defined in (4.21).
If holds uniformly for , the quadratic convergence of the iteration has been proved. Otherwise, there must be some points such that . We then introduce a function and define Case 1 (). In this case, , , and . Hence, . On the other hand, it can be deduced that which is a contradiction.Case 2 (). In this case, , . Furthermore, we can obtain Consequently, the monotonicity of given in the condition () yields However, equation (4.21) minus inequality (4.25) gives which contradicts to (4.31).
Similarly, we can prove that Case , that is, , is also impossible.
Hence, the inequality holds uniformly for , that is, holds uniformly, in which and

5. An Example

Consider a nonlinear BVP as follows: in which is a constant to be determined, that is, in this BVP.

It can be verified that and are, respectively, the lower and upper solutions of BVP (5.1) if is chosen. Moreover, it is easy to show that all the assumptions in Theorem 4.1 are satisfied. Therefore, by Theorem 4.1, BVP (5.1) has a unique solution and it can be approximated quadratically by two monotone sequences.

Acknowledgment

This work is supported by the National Natural Science Foundation of Fujian Province of China under Grant (no. S0650010).