Skip to content
BY 4.0 license Open Access Published by De Gruyter August 23, 2021

Multiplicity of positive solutions for a degenerate nonlocal problem with p-Laplacian

  • Pasquale Candito EMAIL logo , Leszek Gasiński , Roberto Livrea and João R. Santos Júnior

Abstract

We consider a nonlinear boundary value problem with degenerate nonlocal term depending on the Lq-norm of the solution and the p-Laplace operator. We prove the multiplicity of positive solutions for the problem, where the number of solutions doubles the number of “positive bumps” of the degenerate term. The solutions are also ordered according to their Lq-norms.

MSC 2010: 35J20; 35J25; 35Q74

1 Introduction

We study the following class of degenerate nonlocal boundary value problems with p-Laplacian

(P) a Ω u q d x Δ p u = f ( u )  in  Ω , u > 0  in  Ω , u = 0  on  Ω ,

where Ω is a bounded smooth domain in ℝN, N≥1, q≥1, 1 < p<+∞; a ∈ C ([0, ∞)) and f ∈ C ([0, t*]), with t* large enough, are functions satisfying some conditions which we will introduce later. The above problem generalizes the one considered in Gasínski-Santos Júnior [9] to the p-Laplacian case as well as with degenerate term a possibly sign changing.

Motivations for the nonlocal problems like (P) come from the biological models of the population diffusion where the velocity of the dispersion, i.e.,

v=aΩuqdxu

depends on the whole population (compare with Chipot-Rodrigues [6]). In this case u(x) denotes the population density at x and Ω is a pervious container of bacterias.

Another situation where a nonlocal model of this form is used (although in a much simpler setting, with p=q=2 and N=1) is a model for transversal vibrations of elastic strings where the displacements are not necessarily small (see Carrier [2]).

The main feature considered here is the degeneracy of the reaction term, namely, the function a appearing in front of the operator may have change sign several times. More precisely, we suppose that a has a multiplicity of "positive bumps" and possibly "other bumps" between them which can be negative (and as we will see later, they will not "generate" any solutions), namely:

(H0) a: [0, +∞) → ℝ is a continuous function and there exist positive numbers 0 ≕t0t1<t2t3<t4 ⩽ … ⩽ t2K−1<t2K(K=1) such that:a > 0 in (t2k−1, t2k) and a(t2k−1)=a(t2k)=0 for all k ∈ {1, …, K}.

In the past the nonlocal problems with degenerate nonlocal term were considered by Ambrosetti-Arcoya [1] (degeneracy appeared at zero and at infinity) and Santos Júnior-Siciliano [15] (where the problem was variational and in obtaining multiplicity of solutions the so called area condition was exploited). Moreover, we have two papers of Gasínski-Santos Júnior [9, 10], where a similar problem was considered with the Laplacian as a main operator on the left hand side. The authors proved existence, multiplicity as well as nonexistence results for the degenerate nonlocal term depending on the Lq-norm of the solution. For recent developments in problems with nonstandard growth and nonuniform ellipticity with an extensive focus on regularity theory, we mention [12] and the therein references.

Our paper here is the continuation of these works in the direction of a more general operator, namely the p-Laplacian. As far as we know, this is the first time where a p−Laplacian problem (1 < p<+∞) is investigated under the degenerate condition described in (H0). Although, on the right hand side, nonlinearity has a particular growth near zero, as in the case p=2, see [9]. More precisely, we assume:

(H1) There exists t*>0 such that f(t) > 0 in (0, t*), f(t*)=0, f ∈ C ([0, t*]) and the map (0, t*)∋ tf(t)/tp−1 is strictly decreasing.

To obtain our goal we need to put in good order different properties of the minus p−Laplacian, as the (S+) property and monotonicity, see [13], that are combined in a suitable way with the Diaz-Saa’s formula [7], truncation techniques, maximum principle and regularity theory, see for instance [3, 4, 5] and the therein references. Owing to these refined tools we are able to realize, also for 1 < p<+∞, the general strategy developed in [9]. In particular, we investigate the uniqueness, regularity and positivity of the solution uα of the auxiliary problem

(Pk,α,f) a ( α ) Δ p u = f ( u )  in  Ω , u = 0  on  Ω ,

with α ∈ (t2k−1, t2k), k=1, 2, ..., K.

Let us introduce some notation. Along the paper, λ1 is the first eigenvalue of the minus p-Laplacian operator with zero Dirichlet boundary condition, φ1 is the positive eigenfunction associated to λ1 normalized in W01,p(Ω) -norm and e1 is the positive eigenfunction associated to λ1 normalized in L(Ω)-norm.

The relation between the domains of a and f are stated in the following assumption:

H 2 _ t 2 K < t q Ω e 1 q d x .

Finally, to prove the multiplicity result for problem (P), namely the existence of 2K positive solutions corresponding to the "positive bumps" of the function a, we will also need the following two assumptions:

H 3 _ max t t 2 k 1 , t 2 k a ( t ) < γ / λ 1 , f o r a l l k { 1 , , K } , w h e r e γ = lim t 0 + f ( t ) / t p 1 ; H 4 _ max t t 2 k 1 , t 2 k t p a ( t ) > θ , f o r a l l k { 1 , , K } , w h e r e θ = | Ω | p t p ( q 1 ) λ 1 1 max t 0 , t [ t f ( t ) ] .

Remark 1.1

(a) All the hypotheses on a(t), namely (H0), (H3) and (H4) deal only with "even" intervals [t2k−1, t2k] (for k ∈ {1, …, K}). The function a(t) can be arbitrary in "odd" intervals (can be negative, zero or positive without satisfying bound conditions like (H3) or (H4)) and these intervals can be even degenerated to a point. In other words a(t) can be any continuous function defined on a positive interval and we can split its domain into "even" intervals where the above assumptions are satisfied and "odd" intervals (possibly degenerated to a single point), where the above assumptions need not be satisfied. The number of solutions obtained in the paper will double the number of "even" intervals.

(b) Condition (H1) implies that γ=limt0+f(t)/tp1 is well defined and it can be a positive number (see also assumption (H3)) or +∞. In particular, if f (0)>0 then γ=+∞.

(c) Condition (H3) is trivially satisfied if f(0) > 0 (see Remark 1.1(b)). If γ< +∞, (H3) basically means that the peaks of a(t) in even intervals are controlled from above by variation of f at 0. On the other hand, hypothesis (H4) means that the peaks of tp a(t) in “even” intervals are controlled from below by the maximum value of tf(t) in [0, t*] (see Fig. 1).

Fig. 1 Geometry of a(t) and tp a(t) satisfying (H1), (H3) and (H4).
Fig. 1

Geometry of a(t) and tp a(t) satisfying (H1), (H3) and (H4).

(d) Finally, we wish to explicitly observe that 1 < p<+∞ and q≥1 are not required to satisfy any particular further condition. As well as, we emphasize that the nonlinear term f has a behaviour that is prescribed only on [0, t*], so that, nor asymptotic conditions at infinity, neither critical/subcritical growth are involved.

2 Multiplicity of solutions

Hereinafter, ǁ·ǁ denotes the W01,p(Ω) -norm and ∣·∣r the Lr(Ω)-norm with r=1. In this Section, we will produce regular positive solutions of problem (P). In particular, they will belong to int(C+), where

C+=uC01(Ωˉ):u(x)0,xΩ and C01(Ωˉ)=uC1(Ωˉ):u(x)=0,xΩ.

For the sake of completeness, recall that

int(C+)=uC01(Ωˉ):u>0,xΩ,anduν(x)<0xΩ,

where ν=ν(x) denotes the outer unit normal for all x ∈ ∂Ω.

The main result of the present note can be stated as follows.

Theorem 2.1

If hypotheses (H0), (H1), (H2), (H3), and (H4) hold, then problem (P) has at least 2K positive solutions belonging to int(C+), with ordered Lq-norms, namely

t 2 k 1 < Ω u k , 1 q d x < Ω u k , 2 q d x < t 2 k f o r a l l k { 1 , , K } .

In the proof of Theorem 2.1 we will consider a suitable auxiliary problem involving the following truncation function

(2.1) f ( t ) = f ( 0 )  if  t 0 , f ( t )  if  0 t t 0  if  t t .

Fix k ∈ {1, …, K} and α ∈ (t2k−1, t2k), here is the announced auxiliary problem that will play a crucial role

(Pk,α) a ( α ) Δ p u = f ( u )  in  Ω , u = 0  on  Ω .

We start solving problem (Pk,α) by the variational approach. Observe that, after removing the nonlocal term of (P) we obtained a variational problem, so that we can look for solutions of (Pk,α) searching functions uW01,p(Ω) such that

(2.2) a(α)Ω|u|p2|u|vdxΩf(u)vdx=0vW01,p(Ω).

Proposition 2.2

If hypotheses (H0), (H1) and (H3) hold, then for each k ∈ {1, …, K} and α ∈ (t2k−1, t2k) fixed, problem (Pk,α) has a unique solution uα ∈ int(C+) such that 0 < uαt* in Ω. In particular, uα solves the problem

(Pk,α,f) a ( α ) Δ p u = f ( u ) i n Ω , u = 0 o n Ω .

Proof. Fix k ∈ {1, …, K} and α ∈ (t2k−1, t2k). Let us define the energy functional Ik,α:W01,p(Ω)R corresponding to problem (Pk,α), namely

(2.3) Ik,α(u)=1pa(α)upΩF(u)dx,

where F(t)=0tf(s)ds . Since f* is bounded and continuous, it is clear that Ik, α is coercive and weakly lower semicontinuous. Therefore Ik, α has a minimizer uα which is a solution of (Pk,α). Since φ1 ∈ int(C+) (recall that φ1 is the positive eigenfunction of the minus p-Laplacian associated to first eigenvalue λ1 normalized in W01,p(Ω) -norm), observing that

(2.4) Ik,α(tφ1)tp=1pa(α)ΩF(tφ1)(tφ1)pφ1pdx,

by assumption (H3), we can pass to the limit in the previous and obtain

limt0+Ik,α(tφ1)=1pa(α)γλ1<0,

namely, for t > 0 sufficiently small we obtain

(2.5) Ik,α(uα)Ik,α(tφ1)<0,

and so uα is nontrivial.

First, let us verify that 0 ⩽ uαt*. Indeed, if in (2.2) we take v=−(uα), we get

a(α)uαp=Ωf(x,uα)uα0,

and, since a(α)>0, uα=0 , that is 0 ⩽ uα. Analogously, if in (2.2) we take v=(uαt*)+, we get

a(α)(uαt)+p=Ωf(x,uα)(uαt)+dx=0,

so, again recalling that a(α)>0, (uαt*)+=0 and uαt*. Hence, taking in mind the definition of f*, uα solves (Pk,α,f).

Since uα is bounded, by standard regularity arguments due to Lieberman [11, Theorem 1], we conclude that uαC01,β(Ωˉ) (for some β ∈ (0, 1)). Recalling that f ∈ C ([0, t*]) and observing that f(uα)⩾0, one has Δpuα ∈ L and Δpuα⩽0. Hence, by the Maximum Principle (see Gasiński-Papageorgiou [8, Theorem 6.2.8], Vázquez [16] or Pucci-Serrin [14]) we get that uα ∈ int(C+).

Finally, suppose that vα is a second solution to (Pk,α), with uαvα. Arguing as for uα, one can point out that 0 < vαt* in Ω, as well as that vα solves (Pk, α, f). Hence, in view of assumption (H1), from [7] one can derive

0a(α)ΩΔpuαuαp1+Δpvαvαp1(uαpvαp)dx=Ωf(uα)uαp1f(vα)vαp1(uαpvαp)dx<0

and we get a contradiction. □

In order to better describe the strategy that we will follow, observe that when (H0), (H1) and (H3) hold, the previous proposition allow us to define suitable maps Sk:(t2k1,t2k)C01(Ωˉ) for every k=1, …, K by putting

Sk(α)=uα

for every α ∈ (t2k−1, t2k), where uα, that is a minimizer of Ik, α, is the unique solution of (Pk,α,) such that 0 < uαt*. At this point, for every k=1, …, K, we introduce a real function Pk:(t2k1,t2k)R defined by

(2.6) Pk(α)=|Sk(α)|qq=Ωuαqdx

for all α ∈ (t2k−1, t2k), where q≥1. The role of the map Pk is revealed by the following claim

(ᘓ)  if αFixPk, then Sk(α) is a solution of problem (P),

where Fix(Pk)={α(t2k1,t2k):Pk(α)=α} . Indeed, let α ∈ (t2k−1, t2k) be such that Pk(α)=α , then, recalling that Sk(α)=uα is a solution of (Pk,α,f) one can conclude that

aΩuαqdxΔpuα=a(α)Δpuα=f(uα)inΩ

and the claim (ᘓ) holds.

Next lemma will be useful in the proof of Proposition 2.4 which provides the continuity of the map Pk . First observe that, since the map (0, t*)∋ tψ(t)=f(t)/tp−1 is strictly decreasing (see hypothesis (H1)), there exists the inverse ψ−1:(0, γ) → (0, t*), where γ is as in (H3).

Lemma 2.3

Assume that hypotheses (H0), (H1) and (H3) hold. Put

aˉ:=max1kKmaxt[t2k1,t2k]a(t)

and let ε(0,γλ1aˉ) . Then, for every k=1, …, K and α ∈ (t2k−1, t2k) one has

(2.7) cα1pε(ψ1(λ1a(α)+ε))pΩe1pdx

where cα=Ik,α(uα)=minuW01,p(Ω)Ik,α(u) .

Proof. We only treat the case γ<+∞. The case γ=+∞ is analogous. First, by hypothesis (H3), since 0<aˉ<γλ1 , it makes sense the choice of ε(0,γλ1aˉ) . Fix k ∈ {1, …, K} and α ∈ (t2k−1, t2k) and, observing that ε<γλ1aˉγλ1a(α) , we can consider the function

yα:=ψ1(λ1a(α)+ε)e1.

From assumption (H1), we have

(2.8) F(t)1pf(t)tt0.

Hence,

I k , α y a ψ 1 λ 1 a ( α ) + ε p = 1 p a ( α ) e 1 p Ω F y a ψ 1 λ 1 a ( α ) + ε p d x ( 2.8 ) 1 p a ( α ) e 1 p Ω f y α ψ 1 λ 1 a ( α ) + ε p y a d x = 1 p a ( α ) e 1 p Ω f y a y α p 1 e 1 p d x H 1 1 p a ( α ) e 1 p Ω f ψ 1 λ 1 a ( α ) + ε ψ 1 λ 1 a ( α ) + ε p 1 e 1 p d x = 1 p a ( α ) e 1 p λ 1 a ( α ) + ε Ω e 1 p d x = 1 p ε Ω e 1 p d x .

Therefore, (2.7) holds. □

Proposition 2.4

If hypotheses (H0), (H1), and (H3) hold, then for each k ∈ {1, …, K} the map Pk:(t2k1,t2k)R defined in (2.6) is continuous.

Proof. Let {αn}⊂(t2k−1, t2k) be a sequence such αn → α*, for some α* ∈ (t2k−1, t2k). For simplicity, put un=Sk(αn) . Taking in mind (2.5),

(2.9) 1pa(αn)unpΩF(un)dx=Ik,αn(un)<0,

that implies

unppF(t)|Ω|a(αn)nN.

Therefore, {un} is bounded in W01,p(Ω) and, passing to a subsequence if necessary, we may assume that

(2.10) unu in W01,p(Ω),unu in L1(Ω) and un(x)u(x) a.e. in Ω

for some uW01,p(Ω) . Moreover, for every nN one has

(2.11) a(αn)Ω|un|p2unvdx=Ωf(un)vdxvW01,p(Ω).

Testing the previous with v=unu* and passing to the limsup, in view of the continuity of α and f* as well as (2.10) and the Lebesgue’s dominated convergence theorem, we get

a(α)lim supnΔpun,unu0.

The (S+) property of −Δpu assures that un → u* in W01,p(Ω) , so that, passing to the limit in (2.11), we can conclude that u* is a nonnegative weak solution of (Pk,α) with α=α*. We need to show that u*≠ 0. By Lemma 2.3, there exists ε>0, small enough, such that

Ik,αn(un)1pεψ1(λ1a(αn)+ε)pΩe1pdxnN.

So, passing to the limit and using (2.10), we obtain

Ik,α(u)1pεψ1(λ1a(α)+ε)pΩe1pdx<0.

Therefore u*≠ 0. Arguing as in the proof of Proposition 2.2 we can show that u* ∈ int(C+). Since such a solution is unique, we conclude that u=uα=Sk(α) . Finally, again from Proposition (2.2) one has that 0 < unt* for every nN and we can use (2.10) and the Lebesgue’s dominated convergence theorem in order to conclude that Pk(αn)Pk(α) . This proves the continuity of Pk . □

Remark 2.5

A deeper analysis allow us to assure a strongest convergence of the sequence {un} considered in the previous proof. Indeed, since every un solves (Pk,α), with α=αn, one has

Δpun=f(un)a(αn)=:gn.

Recalling that f* is bounded and a(αn) is away from zero, we have

|gn|CnN,

for some C > 0. By the regularity theory (see Lieberman [11]), we have that

unC01,β(Ω)C,

for some C > 0 and β ∈ (0, 1). The compactness of the embedding C1,β(Ω)C1(Ω) , passing to a subsequence if necessary, permits to obtain that

unuinC01(Ω),

(the limit function u* is the same as in (2.10), by the uniqueness of the limit function).

Next lemma will be needed in Proposition 2.7, where the fixed points of Pk are provided.

Lemma 2.6

If hypotheses (H0), (H1) and (H3) hold, then, for every k ∈ {1, …, K} and α ∈ (t2k−1, t2k) one has

(2.12) uαzα:=ψ1(λ1a(α))e1

Proof. Fix k ∈ {1, …, K} and α ∈ (t2k−1, t2k). From hypothesis (H1), the definition of ψ−1 and observing that, in view of (H3), 0 < zαψ−1(λ1a(α))<t*, we have

λ1a(α)=f(ψ1(λ1a(α)))ψ1(λ1a(α))p1f(zα)zαp1,

so

(2.13) a ( α ) Δ p z α = λ 1 a ( α ) z α p 1 f z α  in Ω .

Therefore zα is a subsolution of (Pk,α). Now, put

(2.14) f ( x , t ) = f z α ( x )  if  t z α ( x ) , f ( t )  if  z α ( x ) < t

for every (x, t) ∈ Ω  ×  ℝN and

I˜k,α(u)=a(α)pupΩF˜(x,u)dx

for every uW01,p(Ω) , where F ~ ( x , t ) = 0 t f ~ ( x , s ) d s . The direct methods assure the existence of a minimizer vα for the functional Ĩk, α. Moreover, since φ1, zα ∈ int(C+), for t > 0 small enough one has 0 < tφ1zα in Ω and

I˜k,α(vα)I˜k,α(tφ1)=ttp1pa(α)Ωf(zα)φ1dx<0.

Thus vα is a nontrivial function such that

(2.15) a(α)Ω|vα|p2|vα|wdxΩf˜(x,vα)wdx=0wW01,p(Ω).

Testing (2.15) with −(vα) and (vαt*)+ we obtain that

0vαt.

In particular, we wish to show that

(2.16) vαzα.

Indeed, acting in (2.15) with w=(zαvα)+ and taking in mind (2.13), we obtain

a(α)Δpvα,(zαvα)+=Ωf˜(x,vα)(zαvα)+dx=Ωf(zα)(zαvα)dxa(α)Δpzα,(zαvα)+.

Hence,

a(α)za> vαzαp2zαvαp2vα,zαvαRNdx0.

and, taking into account [13, Lemma A.0.5], (2.16) holds. At this point, it is easy to observe that, because of the defintion of f˜ , one has

a(α)Δpvα=f˜(x,vα)=f(vα)=f(vα),

namely vα is a solution of (Pk,α) such that 0 < vαt*. Finally, Proposition 2.2 assures that vα=uα and the proof is complete in view of (2.16).

Proposition 2.7

If hypotheses (H0), (H1), (H2), (H3), and (H4) hold, then, for every k ∈ {1, …, K} the map Pk has at least two fixed points t2k−1<α1,k<α2,k<t2k.

Proof. Fix k ∈ {1, …, K}. First we show that

(2.17) limαt2k1+Pk(α)> t2k1andlimαt2kPk(α)> t2k.

From Lemma 2.6, we have

Pk(α)=ΩuαqdxΩzαqdx=(ψ1(λ1a(α)))qΩe1qdxα(t2k1,t2k).

Hence, by hypotheses (H1) and (H2), we have

limαt2k1+Pk(α)tqΩe1qdx> t2K> t2k1,limαt2kPk(α)tqΩe1qdx> t2Kt2k,

which proves (2.17).

Next, we show that for some α0 ∈ (t2k−1, t2k) we have

(2.18) Pk(α0)<α0.

For each α ∈ (t2k−1, t2k), let wα be the unique (positive) solution of the problem

(2.19) Δ p u = u α q 1  in  Ω , u = 0  on  Ω ,

where uα=Sk(α) is the (unique) positive solution of (Pk,α). Multiplying by uα, integrating by parts, using Cauchy-Schwarz inequality and Hölder inequality, we have

(2.20) Pk(α)=Ωuαqdx=Ω|wα|p2wαuαdxΩ|wα|p1|uα|dxwαp1uα.

By the variational property of uα, we get

a(α)Ω|uα|pdx=Ωf(uα)uαdx,

thus

(2.21) uα|Ω|1/pa(α)1/pmax[0,t][tf(t)]1/p,

while, taking in mind that wα solves (2.19), by the Hölder inequality and the variational characterization of λ1, we have

wαp=Ω|wα|pdx=Ωuαq1wα(Ωuα(q1)pdx)1/p(Ωwαpdx)1/p|Ω|1/ptq1λ11/pwα,

where 1p+1p=1 , so

(2.22) wαp1|Ω|1/ptq1λ11/p.

Applying (2.21) and (2.22) in (2.20), we obtain

Pk(α)1a(α)1/pmax[0,t][tf(t)]1/p|Ω|tq1λ11/p=θ1/pa(α)1/pα(t2k1,t2k).

Using hypothesis (H4), for some α0 ∈ (t2k−1, t2k) one has

α0pa(α0)=maxt[t2k1,t2k][tpa(t)]>θ

and we get

Pk(α0)θ1/pa(α0)1/p<α0,

namely (2.18) holds.

From the continuity of Pk (see Proposition 2.4), since (2.17) implies the existence of α1 and α2 in (t2k−1, t2k) with P(α1)>t2k−1 and P(α2)>t2k, (2.18) and the intermediate value theorem for continuous real functions, we conclude that Pk has at least two fixed points in the interval (t2k−1, t2k). □

Proof of Theorem 2.1

Fix k ∈ {1, …, K}, recall claim (ᘓ) and observe that the two fixed points of Pk obtained in Proposition 2.7 produce two positive solutions uk,1 and uk,2 of problem (P), with uk,1, uk,2 ∈ int(C+) and satisfying

t2k1<Ωuk,1qdx<Ωuk,2qdx< t2k,

for any k ∈ {1, …, K}. This finishes the proof.

We conclude presenting a concrete example of possible nonlocal problem that can be considered

Example 2.8

Let K be a positive integer, 1 < p<+∞, q=1 and fix t*, M > 0 such that

t>((2K1)π)1/q|e1|qandM|Ω|ptpqepλ1,

where e1 and λ1 are as defined in Section 1, so that (H2) holds. Let a:[0,+)R be a continuous function such that

a(t)=Msint

for t ∈ [0, (2K − 1)π]. Obviously a satisfies hypothesis (H0) for the partition

0=t0=t1< t2=π< t3=2π<<t2K1=2(K1)π< t2K=(2K1)π.

Consider a continuos function f:RR such that

f(t)=tp1lnttift(0,t]0ift=0.

It is simple to verify that

  1. tf(t)/tp−1 is decreasing on (0, t*),

  2. limt0+f(t)/tp1=+ ,

  3. f (0)=f(t*)=0, f(t) > 0 for t ∈ (0, t*).

Namely, hypotheses (H1) and (H3) hold. Finally, a direct computation shows that

maxt[0,t][tf(t)]=tpep.

Hence, for every k ∈ {1, …, K} one has

maxt[t2k1,t2k][tpa(t)]M(π/2)p> M|Ω|ptpqepλ1=|Ω|ptp(q1)λ1maxt[0,t][tf(t)]=θ.

That is (H4) holds too.


The author is a member of the Gruppo Nazionale per l’Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM).

Acknowledgements

The authors would like to thank the anonymous Referees for their valuable comments which helped to improve the manuscript.

The paper is partially supported by PRIN 2017–Progetti di Ricerca di rilevante Interesse Nazionale, "Nonlinear Differential Problems via Variational, Topological and Set-valued Methods" (2017AYM8XW).

  1. Conflict of interest statement: Authors state no conflict of interest.

References

[1] A. Ambrosetti, D. Arcoya, Positive solutions of elliptic Kirchhoff equations, Adv. Nonlinear Stud. 17 (2017), no. 1, 3–16.10.1515/ans-2016-6004Search in Google Scholar

[2] G.F. Carrier, On the non-linear vibration problem of the elastic string, Q. J. Appl. Math. 3 (1945), 151–165.10.1090/qam/12351Search in Google Scholar

[3] P. Candito, L. Gasínski, R. Livrea, Three solutions for parametric problems with nonhomogeneous (a, 2)−type differential operators and reaction terms sublinear at zero, J. Math. Anal. Appl. 480 (2019), no. 1, 123398, 24 pp.10.1016/j.jmaa.2019.123398Search in Google Scholar

[4] P. Candito, S. Carl, R. Livrea, Critical points in open sublevels and multiple solutions for parameter-depending quasilinear elliptic equations, Adv. Differential Equations 19 (2014), no. 11-12, 1021–1042.10.57262/ade/1408367287Search in Google Scholar

[5] P. Candito, S. Carl, R. Livrea, Multiple solutions for quasilinear elliptic problems via critical points in open sublevels and truncation principles, J. Math. Anal. Appl. 395 (2012), 156–163.10.1016/j.jmaa.2012.05.003Search in Google Scholar

[6] M. Chipot, J.F. Rodrigues, On a class of nonlocal nonlinear elliptic problems, RAIRO - Modélisation mathématique et analyse numérique 26 (1992), no. 3, 447–467.10.1051/m2an/1992260304471Search in Google Scholar

[7] J. I. Diaz, J. E. Saa, Existence et unicité de solutions positives pour certaines équations elliptiques quasilinéaires, C. R. Acad. Sci. Paris, t. 305 Série I (1987), 521–524.Search in Google Scholar

[8] L. Gasinski, N.S. Papageorgiou, Nonlinear analysis, Chapman Hall/CRC, Boca Raton, FL, 2006.Search in Google Scholar

[9] L. Gasínski, J.R. Santos Júnior,Multiplicity of positive solutions for an equation with degenerate nonlocal diffusion, Comput. Math. Appl. 78 (2019), 136–143.10.1016/j.camwa.2019.02.029Search in Google Scholar

[10] L. Gasínski, J.R. Santos Júnior, Nonexistence and multiplicity of positive solutions for an equation with degenerate nonlocal diffusion, Bull. Lond. Math. Soc. 52 (2020), no. 3, 489–497.10.1112/blms.12342Search in Google Scholar

[11] G.M. Lieberman, Boundary regularity for solutions of degenerate elliptic equations, Nonlinear Anal. 12 (1988), 1203–1219.10.1016/0362-546X(88)90053-3Search in Google Scholar

[12] G. Mingione, V. Radulescu, Recent developments in problems with nonstandard growth and nonuniform ellipticity, J. Math. Anal. Appl. (2021), Paper 125197.10.1016/j.jmaa.2021.125197Search in Google Scholar

[13] I. Peral, Multiplicity of Solutions for the p-Laplacian, ICTP Lecture Notes of the Second School of Nonlinear Functional Analysis and Applications to Differential Equations. Trieste (1997).Search in Google Scholar

[14] P. Pucci, J. Serrin, The Maximum Principle, in: Progress in Nonlinear Differential Equations and their Applications, 73, Birkhäuser Verlag, Basel, 2007.10.1007/978-3-7643-8145-5Search in Google Scholar

[15] J.R. Santos Júnior, G. Siciliano, Positive solutions for a Kirchhoff problem with vanishing nonlocal term, J. Differential Equations 265 (2018), no. 5, 2034–2043.10.1016/j.jde.2018.04.027Search in Google Scholar

[16] J. L. Vázquez, A strong maximum principle for some quasilinear elliptic equations, Appl. Math. Optim. 12 (1984), no. 3, 191–202.10.1007/BF01449041Search in Google Scholar

Received: 2021-03-26
Accepted: 2021-06-28
Published Online: 2021-08-23

© 2021 Pasquale Candito et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 24.4.2024 from https://www.degruyter.com/document/doi/10.1515/anona-2021-0200/html
Scroll to top button