Abstract

The singular semilinear elliptic problem in , in , on , is considered, where is a bounded domain with smooth boundary in , , and are three positive constants. Some existence or nonexistence results are obtained for solutions of this problem by the sub-supersolution method.

1. Introduction and Main Results

In this paper, we study the existence or the nonexistence of solutions to the following singular semilinear elliptic problem where is a bounded domain with boundary for some , , and are three nonnegative constants. This problem arises in the study of non-Newtonian fluids, chemical heterogeneous catalysts, in the theory of heat conduction in electrically conducting materials (see [17] and their references).

Many authors have considered this problem. For examples, when in , problem (1.1) was studied in [3, 811]; when in , problem (1.1) was considered in [1214]. Particularly, when , it has been established in Zhang [14] that there exists such that problem (1.1) has at least one solution in for all and has no solution in if . After that Shi and Yao in [13] have also obtained the same results with and in . Recently, Ghergu and Rădulescu in [12] considered more general sublinear singular elliptic problem with .

In this paper, we consider the case that , and may have zeros in . The following main results are obtained by the sub-supersolution method with restriction on the boundary in Cui [15].

Theorem 1.1. Suppose that , and . Assume that and , then there exists such that problem (1.1) has at least one solution and for all , and problem (1.1) has no solution in if . Moreover, problem (1.1) has a maximal solution which is increasing with respect to for all .

Remark 1.2. Theorem 1.1 generalizes Theorem in [13] in coefficient of the singular term. Consequently, it also generalizes Theorem in [14]. Moreover, there are functions satisfying our Theorem 1.1 and not satisfying Theorem in [13]. For example, let where , and Certainly, this example does not satisfy Theorem in [12] yet.

Theorem 1.3. Suppose that and in . If , problem (1.1) has no solution in for all and .

Remark 1.4. Obviously, Theorem 1.3 is a generalization of Theorem in [14]. There are also functions satisfying our Theorem 1.3 and not satisfying Theorem in [14] and Theorem in [12]. For example, let where , is any positive constant and is the diameter of .

2. Proof of Theorems

Consider the more general semilinear elliptic problem where the function is locally Hölder continuous in and continuously differentiable with respect to the variable . A function is called to be a subsolution of problem (2.1) if , and A function is called to be a supersolution of problem (2.1) if , and

According to Lemma in the study of Cui [15], we can easily have the following basic existence of classical solution to problem (2.1).

Lemma 2.1. Let be continuously differentiable with respect to the variable . Suppose that problem (2.1) has a supersolution and a subsolution such that then problem (2.1) has at least one solution satisfying

Let be the first eigenvalue of the eigenvalue problem and in the corresponding eigenfunction. Then . Moreover one has the following lemma.

Lemma 2.2 (see [10]). One has if and only if .

Now we give the proof of our theorems.

Proof of Theorem 1.1. Let , and let denote the unique solution of where belongs to (see [16]). Then is a solution of where and . Then fix and set thus we can easily obtain that is a supersolution of problem (1.1).
Now, we want to find a subsolution of problem (1.1). Let where is a positive constant; now we will prove that is a subsolution of problem (1.1). By Hopf's maximum principle in [17], there exist and such that where On , we choose , then we have where for . On , we choose , then one obtains Thus, we choose , then fixing , let it follows from (2.13) and (2.14) that
Thus we proved that is a subsolution of problem (1.1) for all . According to Lemma in [14], there exists a positive constant such that Set , then we have Thus we choose ; via Lemma 2.1, problem (1.1) has at least one solution and satisfying for all .
Since in for all and , according to Lemma 2.2 one has So we obtain .
Let , and let be the unique solution of for , and with , where We claim that is nonincreasing with respect to in for all . Indeed, since is a supersolution of problem (1.1) for all , then we have for all . Since in , so by the maximum principle, one has in . So when our claim is true. We assume that our claim is true when ; that is, in . Then we obtain for all . Since in , so by the maximum principle, one has in . Thus by the induction, one obtains for all . Then by the monotonicity of , we have for all and . According to the definitions of and , we obtain that is a supersolution of problem (1.1) for all . Let be a classical solution of problem (1.1), thus one has Assume that for all , then by standard elliptic arguments (see [17]) it follows that is a solution of problem (1.1), and in for any . Therefore, is the maximal solution of problem (1.1). According to the above arguments, problem (1.1) has a maximal solution for .
To complete the proof of Theorem 1.1, setting then , . It suffices to prove that if , then ; that is, assume that , then problem (1.1) has at least one solution. Let be a solution of problem (1.1) corresponding to , then is a subsolution of problem (1.1) with every fixed . Since is a supersolution of problem (1.1) for any , then one has for all . According to Lemma 2.1, problem (1.1) has at least one solution for all . Moreover, Consequently, the maximal solution of problem (1.1) is increasing with respect to for all . So the proof of Theorem 1.1 is completed.

Proof of Theorem 1.3. Suppose to the contrary that there exists such that problem (1.1) has one solution . Let be the unique solution of . By the maximum principle, in . We claim that for any solution of problem (1.1), there exists a constant such that Indeed, let , then one obtains for all . Since , by the maximum principle we have According to Lemma in [14], there exists a positive constant such that Since , from Lemma 2.2, it follows that Thus we obtain
Set and , then and , satisfying for all and . Consequently, integrating (2.38) we have noting that where denotes the outward normal to . From (2.39) and (2.40), letting , one has where denotes the Lebesgue measure of . According to (2.36) and in , one obtains But this is impossible, by Hopf's maximum principle, we have for all , where denotes the outward normal to at . Therefore Theorem 1.3 is true.

Acknowledgment

This paper is supported by NNSF of China under Grant 10771173 and the Natural Science Foundation of Education of Guizhou Province under Grant 2008067.