Volume 2010, Issue 1 536236
Research Article
Open Access

Sign-Changing Solutions for Nonlinear Elliptic Problems Depending on Parameters

Siegfried Carl

Corresponding Author

Siegfried Carl

Institut für Mathematik, Martin-Luther-Universität Halle-Wittenberg, 06099 Halle, Germany uni-halle.de

Search for more papers by this author
Dumitru Motreanu

Dumitru Motreanu

Département de Mathématiques, Université de Perpignan, 66860 Perpignan, France univ-perp.fr

Search for more papers by this author
First published: 09 February 2010
Academic Editor: Thomas Bartsch

Abstract

The study of multiple solutions for quasilinear elliptic problems under Dirichlet or nonlinear Neumann type boundary conditions has received much attention over the last decades. The main goal of this paper is to present multiple solutions results for elliptic inclusions of Clarke′s gradient type under Dirichlet boundary condition involving the p-Laplacian which, in general, depend on two parameters. Assuming different structure and smoothness assumptions on the nonlinearities generating the multivalued term, we prove the existence of multiple constant-sign and sign-changing (nodal) solutions for parameters specified in terms of the Fučik spectrum of the p-Laplacian. Our approach will be based on truncation techniques and comparison principles (sub-supersolution method) for elliptic inclusions combined with variational and topological arguments for, in general, nonsmooth functionals, such as, critical point theory, Mountain Pass Theorem, Second Deformation Lemma, and the variational characterization of the “beginning”of the Fu\v cik spectrum of the p-Laplacian. In particular, the existence of extremal constant-sign solutions and their variational characterization as global (resp., local) minima of the associated energy functional will play a key-role in the proof of sign-changing solutions.

1. Introduction

Let Ω ⊂ N be a bounded domain with a C2-boundary Ω, and let V = W1,p(Ω) and , 1 < p < +, denote the usual Sobolev spaces with their dual spaces V* and , respectively. We consider the following nonlinear multi-valued elliptic boundary value problem under Dirichlet boundary condition: find uV0∖{0} and parameters a, b such that
(1.1)
where Δpu = div (|∇u|p−2u) is the p-Laplacian, and sj(x, s, a, b) denotes Clarke′s generalized gradient of some locally Lipschitz function sj(x, s, a, b) which depends on x ∈ Ω and the parameters a,   b. For a = b = :λ problem (1.1) reduces to
(1.2)
which may be considered as a nonlinear and nonsmooth eigenvalue problem. We are going to study the existence of multiple solutions of (1.1) for two different classes of j which are in some sense complementary. Our presentation is based on and extends the authors′ recent results obtained in [13]. For the first class of j we let a = b = λ and assume the following structure of j:
(1.3)
where is such that f(·, ·, λ) : Ω × is a Carathéodory function. Problem (1.1) reduces then to the following nonlinear eigenvalue problem:
(1.4)
which will be considered in Section 2 when the parameter λ is small enough.
The second class of j has the following structure:
(1.5)
where s+ = max {s, 0} and s = max {−s, 0} is the positive and negative part of s, respectively, and G : Ω × is assumed to be the primitive of a measurable function g : Ω × that is merely bounded on bounded sets; that is, and G is given by
(1.6)
Problem (1.1) reduces then to the following parameter-dependent multi-valued elliptic problem:
(1.7)
which will be studied in Section 3 for parameters a and b large enough. Note that sG(x, s) stands for the generalized Clarke′s gradient of the locally Lipschitz function sG(x, s). Obviously, if g : Ω × is a Carathéodory function, that is, xg(x, s) is measurable in Ω for all s and sg(x, s) is continuous in for a.a. x ∈ Ω, then G(x, s) = {g(x, s)} is single-valued, and thus problem (1.7) reduces to the following nonlinear elliptic problem depending on parameters a and b: find uV0∖{0} and constants a, b such that
(1.8)
Multiple solution results for (1.8) were obtained by the authors in [4]. Furthermore, note that
(1.9)
Therefore, if one assumes, in addition, a = b = :λ, then (1.8) reduces to the nonlinear elliptic eigenvalue problem: find uV0∖{0} and a constant λ such that
(1.10)
In a recent paper (see [5]) the authors considered the eigenvalue problem (1.10) for a Carathéodory function g. Combining the method of sub-supersolution with variational techniques and assuming certain growth conditions of sg(x, s) at infinity and at zero the authors were able to prove the existence of at least three nontrivial solutions including one that changes sign. The results in [5] improve among others recent results obtained in [6]. For a = b = :λ, (1.7) reduces to the corresponding multivalued eigenvalue problem: find uV0∖{0} and a constant λ such that
(1.11)
The existence of multiple solutions for (1.11) has been shown recently in [7] where techniques for single-valued problems developed in [5] and hemivariational methods applied in [8] have been used. Multiplicity results for (1.11) have been obtained also in [9].

The existence of multiple solutions for semilinear and quasilinear elliptic problems has been studied by a number of authors, for example, [1024]. All these papers deal with nonlinearities (x, s) ↦ g(x, s) that are sufficiently smooth.

2. Problem (1.4) for λ being Small

The aim of this section is to provide an existence result of multiple solutions for all values of the parameter λ in an interval (0, λ0), with λ0 > 0, guaranteeing that for any such λ there exist at least three nontrivial solutions of problem (1.4), two of them having opposite constant sign and the third one being sign-changing (or nodal). More precisely, we demonstrate that under suitable assumptions there exist a smallest positive solution, a greatest negative solution, and a sign-changing solution between them, whereas the notions smallest and greatest refer to the underlying natural partial ordering of functions. This continues the works of Jin [25] (where p = 2 and f(x, s, λ) is Hölder continuous with respect to for every fixed λ) and of Motreanu-Motreanu-Papageogiou [26]. In these cited works one obtains three nontrivial solutions, two of which being of opposite constant sign, but without knowing that the third one changes sign. Here we derive the new information of having, in addition, a sign-changing solution by strengthening the unilateral condition for the right-hand side of the equation in (1.4) at zero. Furthermore, under additional hypotheses, we demonstrate that one can obtain two sign-changing solutions.

2.1. Hypotheses and Example

Let Lq(Ω) +, 1 ≤ q ≤ +, denote the positive cone of Lq(Ω) given by
(2.1)
We impose the following hypotheses on the nonlinearity f(x, s, λ) in problem (1.4).
  • H(f) , with , is a function such that f(x, 0, λ) = 0 for a.a. x ∈ Ω, whenever , and one has the following.

  • (i)

    For all , f(·, ·, λ) is Carathéodory (i.e., f(·, s, λ) is measurable for all s and f(x, ·, λ) is continuous for almost all x ∈ Ω).

  • (ii)

    There are constants c > 0, r > p − 1, and functions a(·, λ) ∈ L(Ω) + () with ∥a(·, λ)∥ → 0 as λ ↓ 0 such that

    (2.2)

  • (iii)

    For all there exist constants μ0 = μ0(λ) > λ2, ν0 = ν0(λ) > μ0 and a set Ωλ ⊂ Ω with Ω∖Ωλ of Lebesgue measure zero such that

    (2.3)
    uniformly with respect to x ∈ Ωλ.

In H(f)(iii), λ2 denotes the second eigenvalue of (−Δp, V0). As mentioned in the Introduction, the strengthening with respect to [26] (see also [25]) of the unilateral condition for the right-hand side f in (1.4), which enables us to obtain, in addition, sign-changing solutions, consists in adding the part involving the limit superior in H(f)(iii).

Let us provide an example where all the assumptions formulated in H(f) are fulfilled.

Example 2.1. For the sake of simplicity we drop the x dependence for the function f in the right-hand side of (1.4). The function f : × (0, +) → given by

(2.4)
with c > 0 and r > p − 1, satisfies hypotheses H(f). Next we give an example of function f : × (0, +) → verifying assumptions H(f) which is generally not odd with respect to s:
(2.5)
with λ > 0, a1 ≥ 1, a2 ≥ 1, c1 > 0, c2 > 0, r1 > p − 1, r2 > p − 1.

2.2. Constant-Sign Solutions

The operator is maximal monotone and coercive; therefore there exists a unique solution eV0 of the Dirichlet problem
(2.6)
With s = max {−s, 0} for s, and using −eV0 as a test function, we see that
(2.7)
which implies that e ≥ 0. From the nonlinear regularity theory (cf., e.g., [27, Theorem 1.5.6]) we have . Then from the nonlinear strong maximum principle (see [28]) we infer that . Here denotes the interior of the positive cone in the Banach space , given by
(2.8)
where n = n(x) is the outer unit normal at xΩ.

Lemma 2.2. Let the data c, r, and a(·, λ) be as in H(f)(ii). Then for every constant θ > 0 there is with the property that if λ ∈ (0, λ0), one can choose ξ0 = ξ0(λ)∈(0, θ) such that

(2.9)

Proof. On the contrary there would exist a constant θ > 0 and a sequence λn ↓ 0 as n such that

(2.10)
Letting n we get for all ξ ∈ (0, θ) because we have ∥a(·,λ)∥ → 0 as λ ↓ 0. Since r > p − 1, a contradiction is achieved as ξ ↓ 0. Therefore (2.9) holds true.

We denote by λ1 the first eigenvalue of (−Δp, V0) and by φ1 the eigenfunction of (−Δp, V0) corresponding to λ1 satisfying
(2.11)

Lemma 2.3. Assume H(f)(i) and (ii) and the following weaker form of hypothesis H(f)(iii): for all there exist μ0 = μ0(λ) > λ1 and Ωλ ⊂ Ω with Ω∖Ωλ of Lebesgue measure zero such that

(2.12)
uniformly with respect to x ∈ Ωλ.

Fix a constant θ > 0 and consider the corresponding number obtained in Lemma 2.2. Then for any λ ∈ (0, λ0) the function , with ξ0 ∈ (0, θ) given by Lemma 2.2, is a supersolution for problem (1.4), and the function is a subsolution of problem (1.4) provided that the number ε > 0 is sufficiently small.

Proof. For a fixed λ ∈ (0, λ0), from (2.9) and H(f)(ii) we derive

(2.13)
which says that is a supersolution for problem (1.4).

On the other hand, by hypothesis we can find μ = μ(λ) > λ1 and δ = δ(λ) > 0 such that

(2.14)
Choose . Then by (2.14) we have
(2.15)
which ensures that is a subsolution of problem (1.4).

The following result which asserts the existence of two solutions of problem (1.4) having opposite constant sign and being extremal plays an important role in the proof of the existence of sign-changing solutions.

Theorem 2.4. Assume H(f)(i) and (ii) and the following weaker form of H(f)(iii): for all there exist constants μ0 = μ0(λ) > λ1, ν0 = ν0(λ) > μ0 and a set Ωλ ⊂ Ω with Ω∖Ωλ of Lebesgue measure zero such that

(2.16)
uniformly with respect to x ∈ Ωλ. Then for all b > 0 there exists a number with the property that if λ ∈ (0, λ0), then there is a constant ξ0 = ξ0(λ) ∈ (0, b/∥e) such that problem (1.4) has a least positive solution in the order interval [0, ξ0e] and a greatest negative solution in the order interval [−ξ0e, 0].

Proof. Since the proof of the existence of the greatest negative solution follows the same lines, we only provide the arguments for the existence of the least positive solution.

Applying Lemma 2.3 for θ = b/∥e  we find as therein. Fix λ ∈ (0, λ0). Lemma 2.3 ensures that is a supersolution for problem (1.4), with ξ0 ∈ (0, b/∥e) given by Lemma 2.2, and is a subsolution for problem (1.4) if ε > 0 is small enough. Passing eventually to a smaller ε > 0, we may assume that εφ1ξ0e. Then by the method of sub-supersolution we know that in the order interval [εφ1, ξ0e] there is a least (i.e., smallest) solution of problem (1.4) (see [29]).

We thus obtain that for every positive integer n sufficiently large there is a least solution of problem (1.4) in the order interval [(1/n)φ1, ξ0e]. Clearly, we have

(2.17)
with some function u+ : Ω → satisfying 0 ≤ u+ξ0e. First we claim that
(2.18)
Taking into account that un solves (1.4), and the fact that un belongs to the order interval [0, ξ0e], from H(f)(ii) we see that
(2.19)
which implies the boundedness of the sequence (un) in V0. Then due to (2.17) we have that u+V0 as well as
(2.20)
Since un solves problem (1.4), one has
(2.21)
Setting φ = unu+ in (2.21) gives
(2.22)
As already noticed that the sequence (f(·, un(·), λ) is uniformly bounded on Ω, so (2.20) and (2.22) yield
(2.23)
The S+-property of −Δp on V0 implies
(2.24)
The strong convergence in (2.24) and Lebesgue′s dominated convergence theorem permit to pass to the limit in (2.21) that results in (2.18).

By (2.18) and the nonlinear regularity theory (cf., e.g., Theorem 1.5.6 in [27]) it turns out . The choice of ξ0 guarantees that

(2.25)
Thus, from (2.18), assumptions H(f)(ii) and (iii), and the boundedness of u+, we get
(2.26)
with a constant . Applying the nonlinear strong maximum principle (cf. [28]) we conclude that either u+ = 0 or .

We claim that

(2.27)
Assume on the contrary that u+ = 0. Then (2.17) becomes
(2.28)
Since un ≥ (1/n)φ1, we may consider
(2.29)
Along a relabelled subsequence we may suppose
(2.30)
for some . Moreover, one can find a function wLp(Ω) + such that for almost all x ∈ Ω. Relation (2.21) reads
(2.31)
Setting leads to
(2.32)
By H(f)(iii) we know that there exist constants c0 = c0(λ) > λ1 and α = α(λ) > 0 such that
(2.33)
while H(f)(ii) entails
(2.34)
for a.a. x ∈ Ω and for all |s| ≥ α. Combining the two estimates gives
(2.35)
with a constant c1 = c1(λ) > 0. Since un ∈ [(1/n)φ1, ξ0e], r > p − 1 and (2.35) holds, there exists a constant C > 0 such that
(2.36)
We see from (2.36) that
(2.37)
Then, because the right-hand side of the above inequality is in L1(Ω), by means of (2.30) and (2.36) we can apply Lebesgue′s dominated convergence theorem to get
(2.38)
Consequently, from (2.32) we obtain
(2.39)
The S+-property of −Δp on V0 implies
(2.40)
On the basis of (2.31) and (2.40) it follows
(2.41)
Notice from (2.36) that
(2.42)
for a.a. x ∈ Ω and for all φV0. We are thus allowed to apply Fatou′s lemma which in conjunction with (2.28), (2.30), and (2.16) ensures
(2.43)
for all φV0,+ : = V0Lp(Ω) +. Thanks to (2.41) we obtain
(2.44)
Owing to (2.42) we may once again use Fatou′s lemma; so according to (2.28), (2.30), and the last part of (2.16), we find
(2.45)
for all φV0,+. Then (2.41) ensures
(2.46)
Combining (2.44) and (2.46) results in
(2.47)
which guarantees to have (see [27, Theorem 1.5.5]). Since by (2.47) we know that , we are in a position to address Theorem 1.5.6 in [27], which provides with some β ∈ (0,1). This regularity up to the boundary and the fact that a.e. in Ω and (2.47) enable us to refer to the strong maximum principle (see Theorem 5 of Vázquez [28]). Recalling that does not vanish identically on Ω (because ) we deduce that for all x ∈ Ω and for all xΩ which amounts to saying . Consequently, there exist constants k0 > 0 and k1 > 0 such that
(2.48)
Following [30] let us denote
(2.49)
whenever (u, v) ∈ DI, where
(2.50)
Relation (2.48) justifies that . Then Proposition 1 of Anane [30] implies . On the other hand a direct computation based on (2.48) and (2.47) shows
(2.51)
This contradiction proves that the claim in (2.27) holds true.

In view of (2.18) it remains to establish that u+ is the smallest positive solution of problem (1.4) in the interval . Let uV0 be a positive solution to (1.4) in . Since uL(Ω), then (1.4) and H(f)(ii) allow to deduce that −ΔpuL(Ω). Using Theorem 1.5.6 of [27] leads to . Then, as u is a solution to (1.4) and , with , by means of hypotheses H(f)(ii) and (iii), we are able to apply the strong maximum principle. So we get , hence for n sufficiently large. The fact that un is the least solution of (1.4) in ensures unu. Taking into account (2.17), we obtain u+u. This completes the proof.

2.3. Sign-Changing Solution

The main result of this section is as follows.

Theorem 2.5. Under hypotheses H(f), for all b > 0, there exists a number with the property that if λ ∈ (0, λ0), then problem (1.4) has a (positive) solution , a (negative) solution and a nontrivial sign-changing solution satisfying ∥u+ < b, ∥u < b, ∥u0 < b.

Proof. Let b > 0. Consider the positive number λ0 given by Theorem 2.4 and fix λ ∈ (0, λ0). Let and be the two extremal solutions determined in Theorem 2.4. We introduce on Ω × the truncation functions

(2.52)
and then define the following associated functionals:
(2.53)
It is clear that E+, E, E0C1(V0).

We observe that if v is a critical point of E+, then

(2.54)
which implies vu+. Similarly, it follows that v ≥ 0. This leads to
(2.55)

Since the function E+ is coercive and weakly lower semicontinuous, there exists a global minimizer z+V0 of it. Using (2.14), it is seen that

(2.56)
and so z+ ≠ 0. Relation (2.55) shows that z+ is a nontrivial solution of problem (1.4) belonging to the order interval [0, u+]. Via assumptions H(f)(ii) and (iii) and the boundedness of z+, we may apply the strong maximum principle which ensures z+ > 0 on Ω. In view of the minimality property of u+ as stated in Theorem 2.4, it follows that z+ = u+. In fact, u+ is the unique global minimizer of E+.

Since , there exists a neighborhood 𝒰 of u+ in the space such that . Therefore E0 = E+ on 𝒰, which guarantees that u+ is a local minimizer of E0 on . It results that u+ is also a local minimizer of E0 on the space V0 (see [27], pages 655-656 ). Employing the functional E and proceeding as in the case of u+, we establish that u is a local minimizer of E0 on V0.

As in the case of (2.55), we verify that every critical point of E0 belongs to the set {uV0 : u(x) ≤ u(x) ≤ u+(x)  a.e.  x ∈ Ω}, which implies that every critical point of E0 is a solution to problem (1.4). The functional E0 is coercive, weakly lower semicontinuous, with . Thus E0 has a global minimizer y0V0 with y0 ≠ 0. The above properties ensure that y0 is a nontrivial solution of problem (1.4) belonging to the order interval [u, u+]. Assume y0u+ and y0u. We claim that y0 changes sign. Indeed, if not, y0 would have constant sign, for instance y0 ≥ 0 a.e. on Ω. Using assumptions H(f)(ii) and (iii) and the boundedness of y0, we may apply the strong maximum principle which leads to y0 > 0 on Ω. This is impossible because it contradicts the minimality property of the solution u+ as given by Theorem 2.4. According to the claim, we obtain the conclusion of the theorem setting u0 = y0.

Thus, the proof reduces to consider the cases y0 = u+ or y0 = u. To make a choice, suppose y0 = u+. We may also admit that u is a strict local minimizer of E0. This is true since on the contrary we would find (infinitely many) critical points x0 of E0 belonging to the order interval [u, u+] which are different from 0, u, u+, and if x0 does not change sign, taking into account the strong maximum principle, the extremality properties of the solutions u, u+ given in Theorem 2.4 will be contradicted. A straightforward argument allows then to find ρ ∈ (0, ∥u+u∥) such that

(2.57)
where Bρ(u) = {uV0   :   uu∥ = ρ}. Relation (2.57) in conjunction with the Palais-Smale condition (which holds for E0 due to its coercivity) enables us to apply the mountain pass theorem to the functional E0 (see, e.g., [31]). In this way we get u0V0 satisfying E0(u0) = 0 and
(2.58)
where
(2.59)
We infer from (2.57) and (2.58) that u0u  and u0u+.

The next step in the proof is to show that

(2.60)
By the equality in (2.58), it suffices to produce a path such that
(2.61)
Let , where , and be endowed with the topologies induced by V0 and , respectively. We set
(2.62)
Making use of the first inequality in assumption H(f) (iii), we fix numbers μ > λ2 and δ > 0 such that (2.14) holds, and then let ρ0 ∈ (0, μλ2). We recall the following variational expression for λ2 given by Cuesta et al. [32]:
(2.63)
where
(2.64)
By (2.63) there exists γ ∈ Γ0 such that
(2.65)
Choose some number r with 0 < r ≤ (λ2 + ρ0) 1/p − (λ2 + ρ0/2) 1/p. The density of SC in S implies that Γ0,C is dense in Γ0; so there is γ0 ∈ Γ0,C satisfying
(2.66)
Then the choice of r establishes
(2.67)
The boundedness of the set in   ensures the existence of some ε1 > 0 such that
(2.68)
Since (see Theorem 2.4), for every uγ0([−1,1]) and any bounded neighborhood Vu of u in there exist positive numbers hu and ju such that
(2.69)
whenever hhu, jju, and vVu. This fact and the compactness of γ0([−1,1]) in allow to determine a number ε0 > 0 for which one has
(2.70)
We now focus on the continuous path εγ0 in joining −εφ1  and εφ1 with a fixed constant ε satisfying 0 < ε < min{ε0, ε1}. By (2.70), (2.67), (2.68), (2.14) with μ > λ2, and taking into account the choice of ρ0 as well as we obtain
(2.71)

At this point we apply the second deformation lemma (see, e.g., [27, page 366]) to the C1 functional E+ : V0. Towards this let us denote

(2.72)
It was already shown that u+ is the unique global minimizer of E+, and so we have m+ < c+. Taking into account (2.55), E+ has no critical values in the interval (m+, c+] (for, otherwise, the minimality of the positive solution u+ of (1.4) would be contradicted). Using also that the functional E+ satisfies the Palais-Smale condition (because it is coercive), the second deformation lemma can be applied to E+  yielding a continuous mapping such that η(0, u) = u and η(1, u) = u+ for all as well as E+(η(t, u)) ≤ E+(u) whenever t ∈ [0,1] and . Introducing γ+ : [0,1] → V0 by
(2.73)
for all t ∈ [0,1], it is seen that γ+ is a continuous path in V0 joining εφ1 and u+. (Note the mapping ww+ is continuous from V0 into itself.) The properties of the deformation η imply
(2.74)
for all t ∈ [0,1]. Similarly, applying the second deformation lemma to the functional E, we construct a continuous path γ : [0,1] → V0 joining u and −εφ1 such that
(2.75)
The union of the curves γ, εγ0, and γ+ gives rise to a path . We see from (2.75), (2.71), and (2.74) that (2.61) is satisfied. Hence (2.60) holds, and so u0 ≠ 0. Recalling that the critical points of E0 are in the order interval {uV0   :   u(x) ≤ u(x) ≤ u+(x)  a.e.  x ∈ Ω} we derive that u0 is a nontrivial solution of (1.4) distinct from u and u+, with uu0u+. By the nonlinear regularity theory we have that . The extremality properties of the constant sign solutions u and u+ as described in Theorem 2.4 force u0 to be sign-changing. This completes the proof.

2.4. Two Sign-Changing Solutions

The goal of this section is to show that under hypotheses stronger than those in Theorem 2.5, problem (1.4) possesses at least two sign-changing solutions.

The new hypotheses on the nonlinearity f(x, s, λ) in problem (1.4) are the following.
  • H(f) , with , is a function such that f(x, 0, λ) = 0 for a.a. x ∈ Ω, whenever .

  • (i)

    For all , .

  • (ii)

    There are constants c > 0, r ∈ (p − 1, p* − 1), and functions a(·, λ) ∈ L(Ω) + () with ∥a(·,λ)∥ → 0 as λ ↓ 0 such that

    (2.76)

  • (iii)

    For all there exist constants μ0 = μ0(λ) > λ2, ν0 = ν0(λ) > μ0 and a set Ωλ ⊂ Ω with Ω∖Ωλ of Lebesgue measure zero such that

    (2.77)
    uniformly with respect to x ∈ Ωλ.

  • (iv)

    There exist constants b < 0 < b+ such that for all we have

    (2.78)

  • (v)

    For every , there exist M = M(λ) > 0 and μ = μ(λ) > p such that

(2.79)

We notice that hypotheses H'(f) are stronger than H(f). In particular, for every , we added the Ambrosetti-Rabinowitz condition for f(·, ·, λ) (see hypothesis H'(f)(v)).

We state now the main result of this section, which produces two sign-changing solutions for problem (1.4).

Theorem 2.6. Assume that hypotheses H'(f) are fulfilled. Then there exists a number with the property that if λ ∈ (0, λ0), then problem (1.4) has a minimal (positive) solution , a maximal (negative) solution and two nontrivial sign-changing solutions satisfying ∥u+ < b, ∥u < b, uu0u+ a.e. in Ω (so ∥u0 < b) and ∥w0b, where b : = min {b+, |b|}.

Proof. Since hypotheses H'(f) are stronger than H(f), we can apply Theorem 2.5 with b = min {b+, |b|}, which ensures the existence of a number such that for every λ ∈ (0, λ0), problem (1.4) possesses a positive solution , a negative solution and a sign-changing solution with −b < uu0u+ < b. The proof of Theorem 2.5 shows that u+ and u can be chosen to be the minimal positive solution and the maximal negative solution, respectively.

On the other hand, hypotheses H'(f) enable us to apply Theorem 1.1 of Bartsch et al. [33]. It follows that there exists a sign-changing solution (by the nonlinear regularity theory) with and . Therefore, we have ∥w0b, which shows that the sign-changing solutions u0 and w0 are different. This completes the proof.

Remark 2.7. In fact, under hypotheses H'(f), for λ ∈ (0, λ0), problem (1.4) admits at least six nontrivial solutions: two positive solutions, two negative solutions, and two sign-changing solutions, as seen in Theorem 5 in [34].

3. Problem (1.7) for Parameters a and b being Large

The main goal of this section is to provide a detailed multiplicity analysis of the nonsmooth elliptic problem (1.7) in dependence of the two parameters a and b. Conditions in terms of the Fučik spectrum are formulated that ensure the existence of sign-changing solutions. As for the precise formulation of this result we recall the Fučik spectrum, see, for example, [13].

The set Σp of those points (μ1, μ2) ∈ 2 for which the problem
(3.1)
has a nontrivial solution is called the Fučik spectrum of the negative p-Laplacian on Ω. Hence, Σp clearly contains the two lines λ1 × and × λ1 with λ1 being the first Dirichlet eigenvalue of −Δp. In addition, the spectrum σ(−Δp) of the negative p-Laplacian has an unbounded sequence of variational eigenvalues (λl), l, satisfying a standard min-max characterization, and Σp contains the corresponding sequence of points (λl, λl), l. A first nontrivial curve 𝒞 in Σp through (λ2, λ2) asymptotic to λ1 × and × λ1 at infinity was constructed and variationally characterized by a mountain-pass procedure by Cuesta et al. [32] (see Figure 1), which implies the existence of a continuous path in {uV0 : I(a,b)(u) < 0,   up = 1} joining −φ1 and φ1 provided (a, b) is above the curve 𝒞. Here the functional I(a,b) on V0 is given by
(3.2)
The hypothesis on the parameters a and b that will finally ensure the existence of sign-changing solutions is as follows.
  • (H)

    Let be above the curve 𝒞 of the Fučik spectrum constructed in [32]; see Figure 1.

Details are in the caption following the image
Fučik Spectrum.

3.1. Hypotheses, Definitions, and Preliminaries

We impose the following hypotheses on the nonlinearity g : Ω × whose primitive is G of problem (1.7)
  • (g1) (x, s) ↦ g(x, s) is measurable in each variable separately.

  • (g2) There exists c > 0, and q ∈ [p, p*) such that

    (3.3)
    for a.a. x ∈ Ω and for all s, where p* denotes the critical Sobolev exponent which is p* = Np/(Np) if p < N, and p* = + if pN.

  • (g3) One has

    (3.4)

  • (g4) One has

    (3.5)

In view of assumptions (g1) and (g2) the function sG(x, s) is locally Lipschitz and the functional 𝒢 : Lq(Ω) → defined by
(3.6)
is well defined and locally Lipschitz continuous as well. The generalized gradients G(x, ·) and 𝒢 can be characterized as follows: Define for every (x, t) ∈ Ω × ,
(3.7)
Proposition 1.7 in [35] ensures that
(3.8)
while Theorem 4.5.19 of [36] implies
(3.9)
with q : = q/(q − 1). The next result is an immediate consequence of [37, Proposition 2.1.5].

Lemma 3.1. Suppose unu in V0, wnw in Lq(Ω), and wn𝒢(un) for all n. Then w𝒢(u).

Definition 3.2. A function uV0 is called a solution of (1.7) if there is an ηLq(Ω) such that

(3.10)

Remark 3.3. Due to assumption (g3) we have g1(x, 0) ≤ 0 ≤ g2(x, 0) for almost all x ∈ Ω. Hence, in view of (3.8), problem (1.7) always possesses the trivial solution.

Definition 3.4. A function is called a subsolution of (1.7) if , and if there is an such that

(3.11)

Similarly, we define a supersolution as follows.

Definition 3.5. A function is called a supersolution of (1.7) if , and if there is an such that

(3.12)

Lemma 3.6. Let e be the uniquely defined solution of (2.6). If a > λ1, then there exists a constant αa > 0 such that for any b the function αae is a positive supersolution of problem (1.7).

Proof. Let a > λ1. By (g4) there is sa > 0 such that

(3.13)
and by (g2) we get
(3.14)
which implies
(3.15)
and thus in view of the definition of g2 we obtain
(3.16)
Let , where αa is a positive constant to be specified. Then we get
(3.17)
which shows that for the function αae is in fact a supersolution of (1.7) with .

In a similar way the following lemma on the existence of a negative subsolution can be proved.

Lemma 3.7. Let e be the uniquely defined solution of (2.6). If b > λ1, then there exists a constant βb > 0 such that for any a the function −βbe is a negative subsolution of problem (1.7).

In the next lemma we demonstrate that small constant multiples of φ1 may be sub- and supersolutions of (1.7). More precisely we have the following result.

Lemma 3.8. Let φ1 be the normalized positive eigenfunction corresponding to the first eigenvalue λ1 of (−Δp, V0). If a > λ1, then for ε > 0 sufficiently small and any b the function εφ1 is a positive subsolution of problem (1.7). If b > λ1, then for ε > 0 sufficiently small and any a the function −εφ1 is a negative supersolution of problem (1.7).

Proof. By (g3) there is a constant δa > 0 such that

(3.18)
which implies
(3.19)
Define with ε > 0. Applying (3.19) and the definition of g1 we get
(3.20)
provided 0 ≤ εφ1δa. The latter can be satisfied by choosing ε sufficiently small such that , where stands for the supremum-norm of φ1. This proves that εφ1 is a subsolution if . In a similar way one can show that for ε sufficiently small the function −εφ1 is a negative supersolution.

Applying a recently obtained comparison result that holds for even more general elliptic inclusions (see [38, Theorem 4.1, Corollary 4.1] we immediately obtain the following theorem.

Theorem 3.9. Let hypotheses (g1)-(g2) be satisfied and assume the existence of a subsolution and supersolution of (1.7) such that . Then there exist extremal solutions of (1.7) within .

3.2. Extremal Constant-Sign Solutions and Their Variational Characterization

Combining the results of Lemmas 3.6, 3.7, and 3.8 and Theorem 3.9 we immediately deduce the existence of nontrivial positive solutions of problem (1.7) provided the parameter a satisfies a > λ1 that and the existence of negative solutions of problem (1.7) provided that the parameter b satisfies b > λ1. Our main goal of this section is to show that problem (1.7) has a smallest positive solution and a greatest negative solution . More precisely the following result will be shown.

Theorem 3.10. Let hypotheses (g1)–(g4) be fulfilled. For every a > λ1 and b there exists a smallest positive solution of (1.7) within the order interval [0, αae] with the constant αa > 0 as in Lemma 3.6. For every b > λ1 and a there exists a greatest negative solution of (1.7) within the order interval [−βb  e, 0] with the constant βb > 0 as in Lemma 3.7.

Proof. Let a > λ1. Lemmas 3.6 and 3.8 ensure that is a supersolution of problem (1.7) and is a subsolution of problem (1.7) provided that ε > 0 is sufficiently small. We may choose ε > 0 such that, in addition, εφ1αae. Thus by Theorem 3.9 there exists a smallest and a greatest solution of (1.7) within the ordered interval [εφ1, αae]. Let us denote the smallest solution by uε. Moreover, the nonlinear regularity theory for the p-Laplacian (cf., e.g., [27, Theorem 1.5.6]) and Vázquez′s strong maximum principle [28] ensure that . Thus for every positive integer n sufficiently large there is a smallest solution of problem (1.7) within [(1/n)φ1, αae]. In this way we inductively construct a sequence (un) of smallest solutions which is monotone decreasing; that is, we have

(3.21)
with some function u+ : Ω → satisfying 0 ≤ u+αae.

Claim 1. u+ is a solution of problem (1.7).

As and un are solutions of (1.7) we have
(3.22)
where ηnLq(Ω) and ηn(x) ∈ G(x, un(x)) for almost all x ∈ Ω. Since un ∈ [(1/n)φ1, αae], the last equation together with (g2) implies that the sequence (un) is bounded in V0. Taking into account (3.21) we obtain that u+V0 and
(3.23)
The solution un of (1.7) satisfies
(3.24)
which yields with φ = unu+ in (3.24) the equation
(3.25)
Using the convergence properties (3.23) of (un) and (g2) as well as the uniform boundedness of the sequence (un), we get by applying Lebesgue′s dominated convergence theorem
(3.26)
which by the S+-property of −Δp on V0 implies
(3.27)
Since un are uniformly bounded, from (g2) we see that there exists a constant c > 0 such that
(3.28)
and thus we get (for some subsequence if necessary) ηnη+ in Lq(Ω). By the strong convergence (3.27), Lemma 3.1 can be applied to show that η+(x) ∈ G(x, u+(x)) for almost every x ∈ Ω. Passing to the limit in (3.24) (for some subsequence if necessary) proves Claim 1.

As u+ belongs, in particular, to L(Ω), Claim 1 and Assumption (g2) implies Δpu+L(Ω). The nonlinear regularity theory (cf., e.g., Theorem 1.5.6 in [27]) ensures that for some γ ∈ (0,1), so . In view of (g2) (g3) a constant can be found such that

(3.29)
which yields in conjunction with Claim 1 that
(3.30)
We now apply Vázquez′s strong maximum principle [28] where in its statement the function β is chosen as for all s > 0, which is possible because . This result guarantees that if u+ ≠ 0, then u+ > 0 in Ω and u+/n < 0 on Ω which means that .

Claim 2. .

Suppose that Claim 2 does not hold. Then by Vázquez′s strong maximum principle we must have u+ = 0, and thus the sequence (un) satisfies
(3.31)
Setting
(3.32)
we may suppose that along a relabelled subsequence one has
(3.33)
with some , and there is a function wLp(Ω) + such that
(3.34)
Since un are positive solutions of (1.7), we get for the following variational equation:
(3.35)
With the special test function in (3.35) we obtain
(3.36)
From (3.29) and (3.34) we get the estimate
(3.37)
As the right-hand side of the last inequality is in L1(Ω), we may apply Lebesgue′s dominated convergence theorem, which in conjunction with (3.33) yields
(3.38)
From (3.33) and (3.36) we conclude
(3.39)
which in view of the S+-property of −Δp on V0 results in
(3.40)
and therefore, in particular, . Taking into account (g3), (3.31), and (3.40), we may pass to the limit in (3.35) which results in
(3.41)
As , relation (3.41) expresses the fact that is an eigenfunction of (−Δp, V0) corresponding to the eigenvalue a. As a > λ1, this is impossible according to Anane [30], because must change sign. This contradiction proves that Claim 2 holds true. Note that unlike in the proof of Theorem 2.4, here the contradiction is achieved by the sign-changing property of eigenfunctions belonging to eigenvalues bigger than λ1.

Claim 3. is the smallest positive solution of (1.7) in [0, αae].

We already know that u+ ∈ [0, αae]. Assume that uV0 is any positive solution of (1.7) belonging to [0, αae]. Since uL(Ω), then by (1.7) and (g3) we deduce ΔpuL(Ω). Using [27, Theorem 1.56] we derive , and applying Vázquez′s strong maximum principle [28] we infer , which yields u ∈ [(1/n)φ1, αae] for n sufficiently large. This in conjunction with the fact that un is the least solution of (1.7) in [(1/n)φ1, αae] ensures unu if n is large enough. In view of (3.21), we obtain u+u, which proves Claim 3.

The proof of the existence of the greatest negative solution of (1.7) within the ordered interval [−βbe, 0] can be done in a similar way. This completes the proof of the theorem.

Under hypotheses (g1)–(g4), Theorem 3.10 ensures the existence of extremal positive and negative solutions of (1.7) for all a > λ1 and b > λ1 denoted by and , respectively. In what follows we are going to characterize these extremal solutions as global (local) minimizers of certain locally Lipschitz functionals that are generated by truncation procedures. To this end let us introduce truncation functions related to the extremal solutions u+ and u as follows:
(3.42)
The truncations τ+, τ, τ0 : Ω × are continuous, uniformly bounded, and Lipschitzian with respect to s. The extremal positive and negative solutions u+ and u of (1.7), respectively, ensured by Theorem 3.10 satisfy
(3.43)
where η+, ηLq(Ω) and
(3.44)
for a.a. x ∈ Ω. By means of η+, η we introduce the following truncations of the nonlinearity g : Ω × :
(3.45)
and define functionals E+,   E,   E0 by
(3.46)
Due to (g2) the functionals E+, E, E0 : V0 are locally Lipschitz continuous. Moreover, in view of the truncations involved these functionals are bounded below, coercive, and weakly lower semicontinuous such that their global minimizers exist. The following lemma provides a characterization of the critical points of these functionals.

Lemma 3.11. Let u+ and u be the extremal constant-sign solutions of (1.7). Then the following holds.

  • (i)

    A critical point vV0 of E+ is a (nonnegative) solution of (1.7) satisfying 0 ≤ vu+.

  • (ii)

    A critical point wV0 of E is a (nonpositive) solution of (1.7) satisfying uw ≤ 0.

  • (iii)

    A critical point zV0 of E0 is a solution of (1.7) satisfying uzu+.

Proof. To prove (i) let v be a critical point of E+, that is, 0 ∈ E+(v), which results in

(3.47)
for some wLq(Ω) and such that w(x) ∈ G+(x, v(x)) almost everywhere in Ω, with
(3.48)
Let us show that vu+ holds. As u+ is a positive solution of (1.7), it satisfies the first equation in (3.43), and by subtracting that equation from (3.47) and applying the special test function φ = (vu+) + we get
(3.49)
By the definition of the truncations introduced above we have and w(x) − η+(x) = 0 for a.a. x ∈ {v > u+}, and thus the right-hand side of (3.49) is zero which leads to
(3.50)
and hence it follows ∇(vu+) + = 0. Because (vu+) +V0, this implies (vu+) + = 0 which proves vu+. To prove 0 ≤ v we test (3.47) with φ = v = max {−v, 0} ∈ V0 and get
(3.51)
which results in , and thus v = 0, that is, 0 ≤ v. Thus the critical point v of E+ which is a solution of the Dirichlet problem (3.47) satisfies 0 ≤ vu+, and therefore τ+(x, v) = v. Because G+(x, v(x)) ⊂ G(x, v(x)), it follows wG(x, v(x)), and therefore v must be a solution of (1.7). This proves (i). In just the same way one can prove also (ii) and (iii) which is omitted.

The following lemma provides a variational characterization of the extremal constant-sign solutions u+ and u.

Lemma 3.12. Let a > λ1 and b > λ1. Then the extremal positive solution u+ of (1.7) is the unique global minimizer of the functional E+, and the extremal negative solution u of (1.7) is the unique global minimizer of the functional E. Both u+ and u are local minimizers of E0.

Proof. The functional E+ : V0 is bounded below, coercive, and weakly lower semicontinuous. Thus there exists a global minimizer v+V0 of E+, that is,

(3.52)
As v+ is a critical point of E+, so by Lemma 3.11 it is a nonnegative solution of (1.7) satisfying 0 ≤ v+u+. Since aλ1 > 0, there is a νa > 0 such that aλ1νa > 0. By (g3) we infer the existence of a such that
(3.53)
and thus for ε > 0 sufficiently small such that
(3.54)
we obtain (note ∥φ1p = 1)
(3.55)
Hence it follows that E+(v+) < 0, and thus v+ ≠ 0. Applying nonlinear regularity theory for the p-Laplacian (cf., e.g., [27, Theorem 1.5.6]) and Vázquez′s strong maximum principle, we see that . As u+ is the smallest positive solution of (1.7) in [0, αae] and 0 ≤ v+u+, it follows v+ = u+, which shows that the global minimizer v+ must be unique and equal to u+. By similar arguments one can show that the global minimizer v of E must be unique and v = u. It remains to prove that u+ and u are local minimizers of E0. Let us show this last assertion for u+ only. By definition we have
(3.56)
Since u+ is a global minimizer of E+ and , it follows that u+ is a local minimizer of E0 with respect to the C1 topology. Due to a result by Motreanu and Papageorgiou in [39, Proposition 4], we conclude that u+ is also a local minimizer of E0 with respect to the V0 topology. This completes the proof of the lemma.

Lemma 3.13. The functional E0 : V0 has a global minimizer v0 which is a nontrivial solution of (1.7) satisfying uv0u+.

Proof. One easily verifies that E0 : V0 is coercive and weakly lower semicontinuous, and thus a global minimizer v0 exists which is a critical point of E0. Apply Lemma 3.11(iii) and note that, for example, E0(u+) = E+(u+) < 0, which shows that v0 ≠ 0.

3.3. Sign-Changing Solutions

Theorem 3.10 ensures the existence of a smallest positive solution in [0, αa  e] and a greatest negative solution of (1.7) in [−βbe, 0]. The idea to show the existence of sign-changing solutions is to prove the existence of nontrivial solutions u0 of (1.7) satisfying uu0u+ with u0u and u0u+, which then must be sign-changing, because u+ and u are the extremal constant-sign solutions.

Theorem 3.14. Let hypotheses (g1)–(g4) and (H) be satisfied. Then problem (1.7) has a smallest positive solution in [0, αa  e], a greatest negative solution in [−βbe, 0], and a nontrivial sign-changing solution with uu0u+.

Proof. Clearly the existence of the extremal positive and negative solution u+ and u follows from Theorem 3.10, because (H), in particular, implies that a > λ1 and b > λ1. As for the existence of a sign-changing solution we first note that by Lemma 3.13 it follows that the global minimizer v0 of E0 is a nontrivial solution of (1.7) satisfying uv0u+. Therefore, if v0u+ and v0u, then v0 = u0 must be a sign-changing solution as asserted, because u is the greatest negative and u+ is the smallest positive solution of (1.7). Thus, we still need to prove the existence of sign-changing solutions in case that either v0 = u or v0 = u+.

Let us consider the case v0 = u+ only, since the case v0 = u can be treated quite similarly. By Lemma 3.12, u is a local minimizer of E0. Without loss of generality we may even assume that u is a strict local minimizer of E0, because on the contrary we would find (infinitely many) critical points z of E0 that are sign-changing solutions thanks to uzu+ and the extremality of the solutions u, u+ obtained in Theorem 3.10 which proves the assertion.

Therefore, it remains to prove the existence of sign-changing solutions under the assumptions that the global minimizer v0 of E0 is equal to u+, and u is a strict local minimizer of E0. This implies the existence of ρ ∈ (0, ∥uu+∥) such that

(3.57)
where Bρ(u) = {uV0   :   uu∥ = ρ}. The functional E0 satisfies the Palais-Smale condition, because it is bounded below, locally Lipschitz continuous, and coercive; see, for example, [40, Corollary 2.4]. Thus in view of (3.57) we may apply the nonsmooth version of Ambrosetti-Rabinowitz′s Mountain-Pass Theorem (see, e.g., [41, Theorem 3.4]) which ensures the existence of a critical point u0V0 satisfying 0 ∈ E0(u0) and
(3.58)
where
(3.59)
It is clear from (3.57) and (3.58) that u0u  and u0u+, and thus u0 is a sign-changing solution provided u0 ≠ 0. To prove the latter we claim
(3.60)
for which it suffices to construct a path such that
(3.61)
The construction of such a path can be done by adopting an approach due to the authors in [3] and applying the Second Deformation Lemma for locally Lipschitz functionals as it can be found in [42, Theorem 2.10]. This completes the proof.

Remark 3.15. The multiplicity results and the existence of sign-changing solutions obtained in this section generalize very recent results due to the authors obtained in [3, 4, 7]. Moreover, if the function tg(x, t) is continuous on and a = b = λ, then G(x, ξ) = {g(x, ξ)}, and problem (1.7) reduces to

(3.62)
However, even in this setting the results obtained here are more general than obtained in [6, Theorem 3.9], because we do not assume that g(x, t)t ≥ 0 for all t.

Remark 3.16. Theorem 3.14 improves also Corollary 3.2 of [8]. In fact, let p = 2, let uV0 be a solution of (1.7) in case a = b = λ and g(x, t) ≡ g(t), (x, t) ∈ Ω × with ηLq(Ω) satisfying η(x) ∈ G(u(x)). By definition of Clarke′s gradient we have, for any φV0,

(3.63)
As u is a solution, the following holds: uV0 and (p = 2),
(3.64)
which yields
(3.65)
That is, u turns out to be a solution of the hemivariational inequality studied in [8]. Since the hypotheses of [8, Corollary 3.2] imply (g1)–(g4), the assertion follows.

Remark 3.17. Multiplicity results for a nonsmooth version of problem (1.4) in form of (1.2) can be established under structure conditions for Clarke′s gradient j similar to H(f).

Multiplicity and sign-changing solutions results have been obtained recently in [43] for the following nonlinear Neumann-type boundary value problem: find uV∖{0} and parameters a, b such that

(3.66)
For problem (3.66) conditions on the parameters have been given in terms of the “Steklov-Fučik” spectrum to ensure multiplicity results.

    The full text of this article hosted at iucr.org is unavailable due to technical difficulties.