Hostname: page-component-745bb68f8f-f46jp Total loading time: 0 Render date: 2025-02-11T15:23:03.729Z Has data issue: false hasContentIssue false

Positive solutions for a degenerate Kirchhoff problem

Published online by Cambridge University Press:  19 August 2021

David Arcoya
Affiliation:
Dpto. de Análisis Matemático, Univ. de Granada, Granada18071, Spain (darcoya@ugr.es)
João R. Santos Júnior
Affiliation:
Faculdade de Matemática, Instituto de Ciências Exatas e Naturais, Univ. Federal do Pará, Avda. Augusto Corrêa 01, Belém, PA66075-110, Brazil (joaojunior@ufpa.br)
Antonio Suárez
Affiliation:
Dpto. de Ecuaciones Diferenciales y Análisis Numérico, Facultad de Matemáticas, Sevilla, Spain (suarez@us.es)
Rights & Permissions [Opens in a new window]

Abstract

By assuming that the Kirchhoff term has $K$ degeneracy points and other appropriated conditions, we have proved the existence of at least $K$ positive solutions other than those obtained in Santos Júnior and Siciliano [Positive solutions for a Kirchhoff problem with vanishing nonlocal term, J. Differ. Equ. 265 (2018), 2034–2043], which also have ordered $H_{0}^{1}(\Omega )$-norms. A concentration phenomena of these solutions is verified as a parameter related to the area of a region under the graph of the reaction term goes to zero.

Type
Research Article
Copyright
Copyright © The Author(s), 2021. Published by Cambridge University Press on Behalf of The Edinburgh Mathematical Society

1. Introduction

In the last decade, non-local Kirchhoff problems have turned into an extremely active field of research in nonlinear partial differential equations. In particular, from different perspectives and approaches, many authors have investigated problems like (or its variations)

(P)\begin{equation} \left \{ \begin{array}{@{}ll} -m(\| u \|^{2})\Delta u = f(u) & \mbox{in}\ \Omega,\\ u=0 & \mbox{on}\ \partial\Omega, \end{array}\right. \end{equation}

where $\Omega \subset {\rm I}\hskip -0.85mm{\rm R} ^{N}$ ($N\geq 1$) is a smooth bounded domain, $\|u\|:=|\nabla u|_{2}$ is the usual norm in the Sobolev space $H^{1}_{0}(\Omega )$, $m:[0,\infty )\to {\rm I}\hskip -0.85mm{\rm R}$ and $f:{\rm I}\hskip -0.85mm{\rm R} \to {\rm I}\hskip -0.85mm{\rm R}$ are continuous functions with $m$ positive and away from zero. Without any intention to provide a complete overview about the matter, we just refer the reader to the following references [Reference Alves, Corrêa and Ma1, Reference He and Zou7, Reference Kirchhoff9Reference Perera and Zhang12] which can give a first glimpse about mathematical, physical and historical aspects involving stationary Kirchhoff models.

Nevertheless, most recently, it has grown the interest of some authors, attracted by inherent mathematical difficulties, in investigating the case in which $m$ is a degenerate function, i.e., when $m$ is not required to be bounded from below by a positive constant. In fact, Ambrosetti and Arcoya [Reference Ambrosetti and Arcoya2, Reference Ambrosetti and Arcoya3] took a first step towards the understanding of this class of problems. In [Reference Ambrosetti and Arcoya3], the function $m$ is allowed to verify $m(0)=0$ and/or $\lim _{t\to +\infty }m(t)=0$. It is worth to point out that in such a paper remains the assumption $m(t)>0$ for $t>0$. On the other hand, Theorem 2.1 (see also Remark 2.2) in [Reference Ambrosetti and Arcoya2] assures the existence of a non-trivial solution even if $m(t)=0$ for some $t>0$.

Inspired by [Reference Brown and Budin5, Reference Hess8], in [Reference Santos Júnior and Siciliano13], the authors considered the case in which $m$ has many different positive zeroes and they proved that the number of positive solutions of (P) is related to the number of points which vanish $m$. In order to state the result, we denote by $t_k$, $k=1,2,\ldots K$, the zeroes of $m$, $F(t): = \int _0^{t} f$, for every $t\geq 0$, $\lambda _1$ the first eigenvalue of Laplacian operator with homogeneous Dirichlet boundary condition and

(1.1)\begin{equation} \alpha_{k}:=\max_{u\in H_{0}^{1}(\Omega), \|u\|\leq t_{k}^{1/2}}\int_{\Omega}F^{{\ast}}(u)\,{\textrm{d}}x, \quad k\in\{1,\ldots, K\}. \end{equation}

where

(1.2)\begin{equation} F^{{\ast}}(t)=\int_{0}^{t}f^{{\ast}}(s)\,{\textrm{d}}s, \mbox{ with } f^{{\ast}}(t)=\left \{ \begin{array}{@{}ll} f(0) & \mbox{if} \ t< 0,\\ f(t) & \mbox{if} \ 0\leq t< s_{{\ast}},\\ 0 & \mbox{if} \ s_{{\ast}}\leq t. \end{array}\right. \end{equation}

and $s_{\ast }>0$ such that $f(t)>0$ in $(0, s_{\ast })$ and $f(s_{\ast })=0$.

It is proved the following theorem.

Theorem 1.1 Santos Júnior and Siciliano [Reference Santos Júnior and Siciliano13]

If $m:[0, \infty )\to {\rm I}\hskip -0.85mm{\rm R}$ is a continuous function and $f:[0, \infty )\to {\rm I}\hskip -0.85mm{\rm R}$ is a locally Lipschitz continuous function satisfying$:$

  1. (m) there exist positive numbers $0=:t_{0}< t_{1}< t_{2}<\cdots < t_{K}$ such that

    \[ m(t_{k})=0 \ \mbox{and} m>0 \mbox{in} (t_{k-1}, t_{k}), \mbox{for all} k=1,\ldots, K; \]
  2. (f) there exists $s_{\ast }>0$ such that $f(t)>0$ in $(0, s_{\ast })$ and $f(s_{\ast })=0;$

  3. (A) $\alpha _{k}<\frac 12\int _{t_{k-1}}^{t_{k}}m(s)\,{\textrm {d}}s< |\Omega | \int _0^{s_{\ast }} f$, for all $k=1,\ldots , K$,

then$,$ problem (P) possesses at least $K$ non-trivial positive solutions. Furthermore$,$ these solutions are ordered in the $H_{0}^{1}(\Omega )$-norm$,$ i.e.$,$

\[ 0<\|u_{1}\|^{2}< t_{1}< \|u_{2}\|^{2}< t_{2}<\ldots< t_{K-1}<\|u_{K}\|^{2}< t_{K}. \]

In this paper, we improve Theorem 1.1 by proving under some appropriated conditions on functions $m$ and $f$, the existence of $K$ further positive solutions for (1.1). Moreover, we show concentration phenomena of these solutions at the zeroes of $m$, when the area $F(s_\ast )=\int _0^{s_{\ast }} f$ tends to zero. Specifically, we have the following result.

Theorem 1.2 Suppose that hypotheses (m) and (f) hold.

  1. (i) If $\alpha _k$ given by (1.1) satisfies that

    1. (A’) $\alpha _{k}\leq \frac 12\int _{t_{k-1}}^{t_{k}}m(s)\,{\textrm {d}}s,$ for all $k=1, \ldots , K,$

    then$,$ problem (P) possesses at least $K-1$ positive solutions which are ordered in $H_{0}^{1}(\Omega )$-norm$,$ i.e.$,$

    \[ t_{1}<\|v_{2}\|^{2}< t_{2}<\cdots< t_{K-1}<\|v_{K}\|^{2}< t_{K}. \]
  2. (ii) Furthermore$,$ if$,$ in addition$,$ the following condition holds

    1. (mf) there exists $\gamma :=\lim _{t\searrow 0} \frac {f(t)}{t}\in (0, \infty )$ and $m(0)<\gamma /\lambda _1,$

    then there is one more positive solution $v_{1}\in H_{0}^{1}(\Omega )$ with

    \[ 0<\|v_{1}\|^{2}< t_{1}. \]
  3. (iii) If we assume the stronger condition (A) instead of (A’), then for every $k=1, \ldots , K,$ the solution $u_k$ provided in Theorem 1.1 verifies $u_k\not = v_k$ and hence problem (P) has at least $2\,K$ positive solutions.

  4. (iv) Moreover$,$ if for each $\lambda >0,$ we denote by $(P_{\lambda })$ the problem (P) with $F(s_{\ast })=\lambda$ and by $v_{k, \lambda },\ k=1, \ldots , K,$ the solutions of $(P_{\lambda })$ obtained in (i)–(ii), then$,$

    \[ \lim_{\lambda\searrow 0}\|v_{k, \lambda}\|^{2}=t_{k-1},\quad \forall\,k=1, \ldots, K. \]

Observe that hypothesis (A’) is weaker than (A).

Let $f_{\lambda }$ be a function $f$ satisfying (f), (mf), (A’) and $F(s_{\ast })=\lambda$, where $\lambda$ is a positive parameter. In Theorem 1.2-iv), we are interested in proving that if $f^{\ast }_{\lambda }:=(f_{\lambda })^{\ast }$ and $\lambda$ goes to zero, then the square of the $H_{0}^{1}$-norm of the solutions of

(P\unicode{x03BB})\begin{equation} \left \{ \begin{array}{@{}ll} -m(\| u \|^{2})\Delta u = f^{{\ast}}_{\lambda}(u) & \mbox{in}\ \Omega,\\ u=0 & \mbox{on}\ \partial\Omega, \end{array}\right. \end{equation}

are concentrating at the points which vanish function $m$. This concentration result is somewhat more general than those commonly found in the literature, mainly because the nonlinearity is not required tending to zero in a uniform way. In fact, in [Reference Clément and Sweers6], for example, in order to study concentration phenomena of the solutions, when $\lambda$ tends to infinity, of the problem

\[ \left \{ \begin{array}{@{}ll} -\Delta u = \lambda f(u) & \mbox{in}\ \Omega,\\ u=0 & \mbox{on}\ \partial\Omega, \end{array}\right. \]

the authors require $\lambda f$ goes uniformly to infinity. Instead, we are just assuming $F(s_{\ast })\to 0$. This assumption allows $f^{\ast }_\lambda$ going to zero in different ways for each $t\in (0, s_{\ast })$.

The paper is organized as follows. In §2, we provide a technical lemma which plays an important role in the existence result. In §3, we prove items (i), (ii) and (iii) of Theorem 1.2 by looking for critical points with norm in the interval $(t_{k-1}^{1/2}, t_k^{1/2})$ of a truncated functional $I_{k}$ defined in $H_0^{1}(\Omega )$. In §4, we study the part (iv) of the cited theorem, that is, the concentration phenomena of the solutions obtained in Theorem 1.2 as $\lambda =F(s_{\ast })$ tends to zero.

2. A technical lemma

Taking into account that $f^{\ast }$ is continuous and bounded, we deduce by the compact embedding Rellich–Kondrachov theorem that the function $J: H_{0}^{1}(\Omega )\to {\rm I}\hskip -0.85mm{\rm R}$ defined by

\[ J(v)= \int_{\Omega}F^{{\ast}}(v)\,{\textrm{d}}x, \quad \forall\,v\in H_0^{1}(\Omega ) \]

is weakly continuous and then it attains its maximum in any closed ball in $H_0^{1}(\Omega )$. In particular, the map $\alpha :[0, \infty )\to {\rm I}\hskip -0.85mm{\rm R}$ given by

(2.1)\begin{equation} \alpha(t)=\max_{u\in H_{0}^{1}(\Omega), \|u\|\leq t^{1/2}}\int_{\Omega}F^{{\ast}}(u)\,{\textrm{d}}x, \quad \forall\,t\geq 0, \end{equation}

is well defined; i.e., the set

(2.2)\begin{equation} S_t:=\{u\in H_{0}^{1}(\Omega): \|u\|^{2}\leq t \ \mbox{and} \ \int_{\Omega}F^{{\ast}}(u)\,{\textrm{d}}x=\alpha(t)\} \end{equation}

is non-empty for every $t\geq 0$.

The following technical lemma collects the main properties of $\alpha$ and $S_t$. In order to state it, we define the function $M_{k} :{\rm I}\hskip -0.85mm{\rm R} \to {\rm I}\hskip -0.85mm{\rm R}$ by

(2.3)\begin{equation} M_{k} (t) = \int_0^{t} m_k , \mbox{ with } m_{k}(t)=\left \{ \begin{array}{@{}ll} m(t), & \mbox{if} \ t_{k-1}\leq t\leq t_{k},\\ 0, & \mbox{otherwise}. \end{array}\right. \end{equation}

Lemma 2.1 Assume that (f) holds and consider the map $\alpha :[0, \infty )\to {\rm I}\hskip -0.85mm{\rm R}$ defined by (2.1) and the set $S_t$ given by (2.2) for $t\geq 0$. We have$:$

  1. (i) for every $t>0,$ $\alpha (t)>0$ and each maximizer $u\in S_t,$ satisfies

    • $\|u\|^{2}=t$, $0< u\leq s_{\ast },$

    • $u\in C^{2, \beta }(\overline {\Omega })$ for some $\beta \in (0, 1)$ and

    • $\partial u/\partial \tau <0$ on $\partial \Omega ,$ where $\tau$ stands for the outward unit normal vector$;$

  2. (ii) the map $\alpha$ is differentiable in $(0,\infty )$ with $\alpha '(t)=(1/2t)\max _{u\in S_t}\int _{\Omega }f^{\ast }(u)u\,{\textrm {d}}x$ and $\alpha '$ is upper semicontinuous. In particular$,$ $\alpha$ is increasing. Moreover$,$ if there exists $\gamma = \lim _{t\to 0} f(t)/t,$ then

    \[ \lim_{t\searrow 0}\alpha'(t)=\frac{\gamma}{2\lambda_1} \quad \mbox{and} \quad {\liminf_{t\to t_{{\ast}}}\alpha'(t)>0},\quad \forall\,t_{{\ast}}>0; \]
  3. (iii) if$,$ in addition$,$ condition (m) holds$,$ then for every $k\in \{2, \ldots , K\}$, there exists $\delta _{k-1}>0$ such that the map $g_{k}(t)=(1/2)M_{k}(t)-\alpha (t)$ is decreasing in $(0, t_{k-1}+\delta _{k-1})$. Moreover$,$ if (mf) holds$,$ the same is true in the case $k=1$.

Proof. (i) Fix $t>0$. For any positive function $\varphi \in H_{0}^{1}(\Omega )$ with $\|\varphi \|^{2}\leq t$, we have

\[ 0<\int_{\Omega}F^{{\ast}}(\varphi)\,{\textrm{d}}x\leq \max_{v\in H_{0}^{1}(\Omega), \|v\|\leq t^{1/2}}\int_{\Omega}F^{{\ast}}(v)\,{\textrm{d}}x =\alpha (t), \]

and then $\alpha (t) >0$. In particular, if $u\in S_t$, then $u \not = 0$. Since

\[ \int_{\Omega}F^{{\ast}}(v)\,{\textrm{d}}x\leq \int_{\Omega}F^{{\ast}}(|v|)\,{\textrm{d}}x \quad \mbox{and} \quad \|v\|=\||v|\|, \]

with strict inequality if $v\not \equiv |v|$, we derive that a maximizer $u\in S_t$ is also non-negative.

In addition, we have $\|u \|=t^{1/2}$. Indeed, if, by contradiction, it would be satisfied that $\|u\|< t^{1/2}$, then $u$ would be a local maximum point in the interior of the ball $B_{t^{1/2}}(0)$ of the functional $J\in C^{1}( H_{0}^{1}(\Omega ))$. We would get

\[ 0=J'(u)v=\int_{\Omega}f^{{\ast}}(u)v\,{\textrm{d}}x, \quad \forall\,v\in H_{0}^{1}(\Omega), \]

and then $f^{\ast }(u)=0$, contradicting that $u$ is non-negative and non-trivial and proving the claim.

In particular, by Lagrange multipliers theorem, there exists $\lambda \in {\rm I}\hskip -0.85mm{\rm R}$ verifying

\[ \int_{\Omega}f^{{\ast}}(u)v\,{\textrm{d}}x=2\lambda\int_{\Omega}\nabla u\nabla v\,{\textrm{d}}x, \quad \forall\, v\in H_{0}^{1}(\Omega). \]

Choosing $v=u$ in the previous equality, we obtain

\[ \lambda=\frac{1}{2t}\int_{\Omega}f^{{\ast}}(u)u\,{\textrm{d}}x>0. \]

Thus, $u$ is a non-trivial and non-negative weak solution of

\[ \left \{ \begin{array}{@{}ll} -\Delta u = \displaystyle\dfrac{1}{2\lambda} f^{{\ast}}(u) & \mbox{in}\ \Omega,\\ u=0 & \mbox{on}\ \partial\Omega. \end{array}\right. \]

To see that $u\leq s_{\ast }$, it is sufficient to choose $v=(u-s_{\ast })^{+}$ as test function in the above boundary value problem. Since $f^{\ast }(u)=f(u)\in L^{\infty }(\Omega )$, it also follows from [Reference Ambrosetti and Prodi4, Theorem 0.5(ii)] that $u\in C^{1, \beta }(\overline {\Omega })$ for some $0<\beta <1$. Now, since $f$ is locally Lipschitz continuous, we get $f(u)\in C^{0,\beta }(\overline {\Omega })$. Hence, by [Reference Ambrosetti and Prodi4, Theorem 0.5(iii)], we have $u\in C^{2, \beta }(\overline {\Omega })$. Applying then the strong maximum principle and Hopf lemma, we obtain $u>0$ in $\Omega$ and $\partial u/\partial \tau <0$ on $\partial \Omega$ and case (i) is proved.

(ii) We begin by proving the continuity of $\alpha$ in $(0, \infty )$. Let $\{t_{n}\}$ be a sequence of positive numbers with $t_{n}\to t_{\ast }$, $t_{\ast }>0$, and $u_{n}\in S_{t_{n}}$. Since $\{u_{n}\}$ is bounded in $H_{0}^{1}(\Omega )$, there exists $u_{\ast }\in H_{0}^{1}(\Omega )$ such that, up to a subsequence, $\{ u_n\}$ is weakly convergent to $u_{\ast }$. Using that the norm is weakly lower semicontinuous and the weak continuity of $J$, we get

\[ \|u_{{\ast}}\|^{2}\leq\liminf_{n\to\infty}\|u_{n}\|^{2}= t_{{\ast}} \mbox{ and } \lim_{n\to\infty} \alpha(t_{n}) = \lim_{n\to\infty}\int_{\Omega}F^{{\ast}}(u_{n})\,{\textrm{d}}x= \int_{\Omega}F^{{\ast}}(u_{{\ast}})\,{\textrm{d}}x. \]

Hence, by definition of $\alpha$, it is clear that

\[ \int_{\Omega}F^{{\ast}}(u_{{\ast}})\leq \alpha(t_{{\ast}}). \]

On the other hand, for the converse inequality, choosing $u_{0}\in S_{t_{\ast }}$ we have

\[ \int_{\Omega}F^{{\ast}} \bigg(\frac{t_{n}^{1/2}}{t_{{\ast}}^{1/2}}u_{0} \bigg)\,{\textrm{d}}x\leq \int_{\Omega}F^{{\ast}}(u_{n})\,{\textrm{d}}x,\quad \forall\,n\in{\rm I}\hskip -0.85mm{\rm N}, \]

and passing to the limit as $n$ goes to infinity, we get

\[ \int_{\Omega}F^{{\ast}}(u_{0})\,{\textrm{d}}x = \alpha(t_{{\ast}})\leq \int_{\Omega}F^{{\ast}}(u_{{\ast}})\,{\textrm{d}}x, \]

concluding that

\[ \lim_{n\to\infty} \alpha(t_{n})=\lim_{n\to\infty} \int_{\Omega}F^{{\ast}}(u_{n})\,{\textrm{d}}x\to\int_{\Omega}F^{{\ast}}(u_{{\ast}})\,{\textrm{d}}x=\alpha(t_{{\ast}}), \]

and the continuity of $\alpha$ has been proved.

Fix now $t_{0}>0$. For each $t\in (-t_0,\infty )$, we take a function $u_{t_{0}}$ (respectively, $u_{t_{0}+t}$) in $S_{t_{0}}$ (respectively, in $S_{t+t_{0}}$). Thus, by definition of $S_{t+t_{0}}$,

\begin{align*} \frac{\alpha(t_{0}+t)-\alpha(t_{0})}{t}& =\frac{1}{t}\int_{\Omega}[F^{{\ast}}(u_{t_{0}+t})-F^{{\ast}}(u_{t_{0}})]\,{\textrm{d}}x\\ & \geq \frac{1}{t}\int_{\Omega}\left[F^{{\ast}}\Big( \Big(1+ \frac{t}{t_{0}}\Big)^{1/2}u_{t_{0}} \Big)-F^{{\ast}}(u_{t_{0}})\right]\,{\textrm{d}}x\\ & =\frac{1}{2t_{0}(1+tc_{t}/t_{0})^{1/2}}\int_{\Omega}f^{{\ast}}\Big( \Big(1+\frac{tc_{t}}{t_{0}} \Big)^{1/2}u_{t_{0}}\Big)u_{t_{0}}\,{\textrm{d}}x, \end{align*}

where $0<|c_{t}|<1$. Consequently,

\[ \liminf_{t\to 0}\left[\frac{\alpha(t_{0}+t)-\alpha(t_{0})}{t}\right]\geq \frac{1}{2t_{0}}\int_{\Omega}f^{{\ast}}(u_{t_{0}})u_{t_{0}}\,{\textrm{d}}x,\quad \forall\,u_{t_{0}}\in S_{t_{0}}. \]

Since $f^{\ast }$ is bounded, it is a consequence of the definition of $S_t$ that $\max _{u\in S_{t_{0}}}\int _{\Omega }f^{\ast }(u)u\,{\textrm {d}}x$ is attained. Therefore,

(2.4)\begin{equation} \liminf_{t\to 0}\left[\frac{\alpha(t_{0}+t)-\alpha(t_{0})}{t}\right]\geq \frac{1}{2t_{0}}\max_{u\in S_{t_{0}}}\int_{\Omega}f^{{\ast}}(u)u\,{\textrm{d}}x. \end{equation}

On the other hand, by definition of $S_{t_{0}}$, we get

\begin{align*} \frac{\alpha(t_{0}+t)-\alpha(t_{0})}{t}& \leq \frac{1}{t}\int_{\Omega}\left[F^{{\ast}}\left(u_{t_{0}+t}\right)-F^{{\ast}}\left(\left(1- \frac{t}{t_{0}+t}\right)^{1/2}u_{t_{0}+t}\right)\right]\,{\textrm{d}}x\\ & =\frac{1}{2(t_{0}+t)(1-td_{t}/(t_{0}+t))}\\ & \quad \int_{\Omega}f^{{\ast}}\left(\left(1-\frac{td_{t}}{t_{0}+t}\right)^{1/2} u_{t_{0}+t}\right) \left( 1-\frac{td_{t}}{t_{0}+t}\right)^{1/2}u_{t_{0}+t}\,{\textrm{d}}x, \end{align*}

where $0<|d_{t}|<1$. Therefore,

(2.5)\begin{equation} \frac{\alpha(t_{0}+t)-\alpha(t_{0})}{t}\leq \frac{1}{2(t_{0}+t)(1-td_{t}/(t_{0}+t))}\max_{u\in S_{t_{0}+t(1-d_{t})}}\int_{\Omega}f^{{\ast}}(u)u\,{\textrm{d}}x. \end{equation}

Claim $\limsup _{t\to 0}[\max _{u\in S_{t_{0}+t(1-d_{t})}}\int _{\Omega }f^{\ast }(u)u\,{\textrm {d}}x]\leq \max _{u\in S_{t_{0}}}\int _{\Omega }f^{\ast }(u)u\,{\textrm {d}}x$.

Let $\{t_{n}\}$ be a sequence of real numbers converging to zero. For each $n$, choose $u_{n}\in S_{t_{0}+t_{n}(1-d_{t_{n}})}$ such that

\[ \int_{\Omega}f^{{\ast}}(u_{n})u_{n}\,{\textrm{d}}x=\max_{u\in S_{t_{0}+t_{n}(1-d_{t_{n}})}}\int_{\Omega}f^{{\ast}}(u)u\,{\textrm{d}}x. \]

Since $\|u_{n}\|^{2}\leq t_{0}+t_{n}(1-d_{t_{n}}),$ we have that $\{u_{n}\}$ is bounded in $H_{0}^{1}(\Omega )$ and, as before, there exists $u_{0}\in H_{0}^{1}(\Omega )$ such that, up to a subsequence,

\[ u_{n}\rightharpoonup u_{0}\ \mbox{in} H_{0}^{1}(\Omega), \]

and

\[ \|u_{0}\|^{2}\leq \liminf_{n\to \infty}\|u_{n}\|^{2}\leq t_{0}. \]

Moreover, by the continuity of $\alpha$,

\[ \alpha(t_{0})=\lim_{n\to\infty}\alpha(t_{0}+t_{n}(1-d_{t_{n}}))=\lim_{n\to\infty}\int_{\Omega}F^{{\ast}}(u_{n})\,{\textrm{d}}x=\int_{\Omega}F^{{\ast}}(u_{0})\,{\textrm{d}}x. \]

Consequently, $u_{0}\in S_{t_{0}}$ and, by $(i)$, $\|u_{0}\|^{2}=t_{0}$. The claim follows now by observing that the weak convergence of $u_n$ to $u_0$ implies that

\begin{align*} \limsup_{n\to \infty} \max_{u\in S_{t_{0}+t_{n}(1-d_{t_{n}})}}\int_{\Omega}f^{{\ast}}(u)u\,{\textrm{d}}x & = \limsup_{n\to \infty}\int_{\Omega}f^{{\ast}}(u_{n})u_{n}\,{\textrm{d}}x\\ & =\int_{\Omega}f^{{\ast}}(u_{0})u_{0}\,{\textrm{d}}x\\ & \leq \max_{u\in S_{t_{0}}}\int_{\Omega}f^{{\ast}}(u)u\,{\textrm{d}}x. \end{align*}

Now, by previous claim, passing to the upper limit as $t$ tends to zero in (2.5), we get

\[ \limsup_{t\to 0}\left[\frac{\alpha(t_{0}+t)-\alpha(t_{0})}{t}\right]\leq \frac{1}{2t_{0}}\max_{u\in S_{t_{0}}}\int_{\Omega}f^{{\ast}}(u)u\,{\textrm{d}}x. \]

By comparing this inequality with (2.4), we conclude that $\alpha$ is differentiable in $t_{0}$ and

\[ \alpha'(t_{0})=\frac{1}{2t_{0}}\max_{u\in S_{t_{0}}}\int_{\Omega}f^{{\ast}}(u)u\,{\textrm{d}}x. \]

To prove the upper semicontinuity of the derivative, let $\{t_{n}\}$ be a sequence of real numbers converging to a positive number $t_{0}$. For each $n$, choose $u_{n}\in S_{t_{n}}$ such that

\[ \int_{\Omega}f^{{\ast}}(u_{n})u_{n}\,{\textrm{d}}x=\max_{u\in S_{t_{n}}}\int_{\Omega}f^{{\ast}}(u)u\,{\textrm{d}}x. \]

Since $u_{n}\in S_{t_{n}}$, as previously, it follows that there exists $u_{0}\in H_{0}^{1}(\Omega )$ such that, up to a subsequence,

\begin{align*} & u_{n}\rightharpoonup u_{0}\ \mbox{in} H_{0}^{1}(\Omega),\\ & \|u_{0}\|^{2}\leq \liminf_{n\to \infty}\|u_{n}\|^{2}\leq t_{0},\\ & \lim_{n\to\infty}\alpha(t_n) = \lim_{n\to\infty} \int_{\Omega}F^{{\ast}}(u_{n})\,{\textrm{d}}x = \int_{\Omega}F^{{\ast}}(u_{0})\,{\textrm{d}}x,\\ & \lim_{n\to\infty} \int_{\Omega}f^{{\ast}}(u_{n})u_{n}\,{\textrm{d}}x= \int_{\Omega}f^{{\ast}}(u_{0})u_{0}\,{\textrm{d}}x. \end{align*}

Using the continuity of $\alpha$,

\[ \alpha(t_{0})=\int_{\Omega}F^{{\ast}}(u_{0})\,{\textrm{d}}x, \]

and consequently

\[ u_{0}\in S_{t_{0}}, \]

and by the computation of the derivative $\alpha '(t_0)$, we obtain

\[ \limsup_{n\to\infty}\alpha'(t_{n})=\limsup_{n\to\infty}\frac{1}{2t_{n}}\int_{\Omega}f^{{\ast}}(u_{n})u_{n} \,{\textrm{d}}x=\frac{1}{2t_{0}}\int_{\Omega}f^{{\ast}}(u_{0})u_{0}\,{\textrm{d}}x\leq \alpha'(t_{0}). \]

To prove the last part of $(ii)$, let $t_{n}\searrow 0$ and $u_{n}\in S_{t_{n}}$ be such that

\[ \alpha'(t_{n})=\frac{1}{2t_{n}}\int_{\Omega}f^{{\ast}}(u_{n})u_{n}\,{\textrm{d}}x. \]

Define $v_{n}:=u_{n}/t_{n}^{1/2}$ and observe that, by (i), $\|u_{n}\|=t_{n}^{1/2}$ and thus $\{v_{n}\}$ is bounded in $H_{0}^{1}(\Omega )$. Consequently, there exists $v_{\ast }$ such that, up to a subsequence,

(2.6)\begin{equation} v_{n}\rightharpoonup v_{{\ast}}\quad \mbox{in}\ H_{0}^{1}(\Omega). \end{equation}

Thus, by Lebesgue Dominated Convergence Theorem and the definition of $\gamma$, we get

(2.7)\begin{equation} \alpha'(t_{n})=\frac{1}{2t_{n}}\int_{\Omega}f^{{\ast}}(u_{n})u_{n}\,{\textrm{d}}x=\frac{1}{2} \int_{\Omega}\left[\frac{f^{{\ast}}(t_{n}^{1/2}v_{n})}{t_{n}^{1/2}v_{n}}\right]v_{n}^{2}\,{\textrm{d}}x\to \frac{\gamma}{2}\int_{\Omega}v_{{\ast}}^{2}\,{\textrm{d}}x. \end{equation}

Moreover, by the definition of $\gamma$ again and the L'Hospital rule, we have $\lim _{t\searrow 0}F^{\ast }(t)/t^{2}=\gamma /2$. Thence, from (2.6), up to a subsequence, we obtain

\[ \lim_{n\to\infty} \frac{F^{{\ast}}(u_{n})}{t_{n}}=\lim_{n\to\infty}\left[\frac{F^{{\ast}}(t_{n}^{1/2}v_{n})}{(t_{n}^{1/2}v_{n})^{2}}\right]v_{n}^{2}=\frac{\gamma}{2} v_{{\ast}}^{2} \ \mbox{a.e. in}\ \Omega. \]

and

\[ \left| \frac{F^{{\ast}}(u_{n})}{t_{n}}\right|=\left|\frac{F^{{\ast}}(t_{n}^{1/2}v_{n})}{(t_{n}^{1/2}v_{n})^{2}}\right||v_{n}^{2}|\leq \gamma h(x) \ \mbox{a.e. in}\ \Omega, \]

for some $h\in L^{1}(\Omega )$. (Observe that the convergence (2.6) implies, passing to a subsequence if necessary, the existence of such a function $h(x)$). Therefore, by Lebesgue Dominated Convergence Theorem

(2.8)\begin{equation} \int_{\Omega}\frac{F^{{\ast}}(u_n)}{t_{n}}\,{\textrm{d}}x\to \frac{\gamma}{2}\int_{\Omega}v_{{\ast}}^{2}\,{\textrm{d}}x. \end{equation}

If it were $v_{\ast }=0$, then by (2.7) and (2.8), we would conclude that

(2.9)\begin{equation} 0=\lim_{n\to\infty}\alpha'(t_{n})=\frac{1}{2}\lim_{n\to\infty} \frac{1}{t_{n}}\int_{\Omega}f^{{\ast}}(u_{n})u_{n}\,{\textrm{d}}x=\lim_{n\to\infty}\int_{\Omega}\frac{F^{{\ast}}(u_n)}{t_{n}}\,{\textrm{d}}x. \end{equation}

On the other hand, since $u_{n}\in S_{t_{n}}$, it follows again from Lebesgue Dominated Convergence Theorem and L'Hospital rule, that

(2.10)\begin{equation} \lim_{n\to\infty}\int_{\Omega}\frac{F^{{\ast}}(u_n)}{t_{n}}\,{\textrm{d}}x\geq \lim_{n\to\infty}\int_{\Omega} \frac{F^{{\ast}}(t_{n}^{1/2}\varphi_1)}{(t_{n}^{1/2}\varphi_1)^{2}}\varphi^{2}_{1}\,{\textrm{d}}x=\frac{\gamma}{2\lambda_1}, \end{equation}

where $\varphi _{1}$ is the eigenfunction of Laplacian operator with homogeneous Dirichlet boundary condition, associated with $\lambda _1$ with $\|\varphi _{1}\|=1$. By comparing (2.9) and (2.10), we get a contradiction proving that $v_{\ast }\neq 0$ $(v_{\ast }\geq 0)$.

Finally, since $u_{n}\in S_{t_{n}}$ and $\|u_{n}\|=t_{n}^{1/2}$, we can again use the Lagrange multipliers theorem (see proof of item $(i)$) to conclude that

\[ \left \{ \begin{array}{@{}ll} -\Delta u_n = \displaystyle\dfrac{1}{2\alpha'(t_{n})} f^{{\ast}}(u_n) & \mbox{in}\ \Omega,\\ u_n=0 & \mbox{on}\ \partial\Omega. \end{array}\right. \]

Consequently,

(2.11)\begin{equation} \int_{\Omega}\nabla v_{n}\nabla w \,{\textrm{d}}x=\displaystyle\frac{1}{2\alpha'(t_{n})} \int_{\Omega}\left[\frac{f^{{\ast}}(t_{n}^{1/2}v_{n})}{t_{n}^{1/2}v_{n}}\right]v_{n} w \,{\textrm{d}}x,\quad \forall\, w\in H_{0}^{1}(\Omega). \end{equation}

By passing to the limit as $n$ goes to infinity in (2.11), it follows from (2.6) and (2.7) that

\[ \int_{\Omega}\nabla v_{{\ast}}\nabla w \,{\textrm{d}}x= \frac{1}{\int_{\Omega}v_{{\ast}}^{2}\,{\textrm{d}}x} \int_{\Omega}v_{{\ast}} w \,{\textrm{d}}x, \quad \forall\, w\in H_{0}^{1}(\Omega). \]

Since $v^{*}\geq 0$, $v^{*}\not \equiv 0$, we have $\int _{\Omega }v_{\ast }^{2}\,{\textrm {d}}x=1/\lambda _{1}$ and (2.7) implies that $\lim _{n\to \infty } \alpha ' (t_n) = \gamma /(2\lambda _1)$.

Finally, suppose by contradiction that there exists a sequence of positive numbers $\{t_{n}\}$ and $\{u_{n}\} \subset S_{t_n}$ such that $t_{n}\to t_{\ast }$ (for some $t_{\ast }>0$), $\|u_{n}\|= t_{n}^{1/2}$, $\int _{\Omega }F^{\ast }(u_{n})\,{\textrm {d}}x=\alpha (t_{n})$ and

(2.12)\begin{equation} \alpha' (t_n)=\frac{1}{2t_{n}}\int_{\Omega}f^{{\ast}}(u_{n})u_{n}\,{\textrm{d}}x\to 0. \end{equation}

In particular, $\{u_{n}\}$ is bounded and hence there exists $u_{\ast }\in H_{0}^{1}(\Omega )$ with $u_{\ast }\geq 0$ (because $u_{n}>0$) and such that, up to a subsequence,

\[ u_{n}\rightharpoonup u_{{\ast}}\ \mbox{in} H_{0}^{1}(\Omega). \]

The weak continuity of $J$ and of the map $\int _{\Omega }f^{\ast }(v)v\,{\textrm {d}}x$, (2.12), $(i)$ and the continuity of $\alpha$ imply by taking limits when $n$ tends to infinity that

(2.13)\begin{equation} \int_{\Omega}F^{{\ast}}(u_{{\ast}})\,{\textrm{d}}x=\alpha (t_{{\ast}})>0 \end{equation}

and

(2.14)\begin{equation} \frac{1}{2t_{{\ast}}} \int_{\Omega}f^{{\ast}}(u_{{\ast}})u_{{\ast}}\,{\textrm{d}}x=0 \end{equation}

Identity (2.14) implies that $u_{\ast }=0$, but that leads us to a contradiction with (2.13). The result is now proved.

(iii) Let us consider first $k=2,\ldots , K$. By (ii), $g_{k}: (0, \infty )\to {\rm I}\hskip -0.85mm{\rm R}$ is differentiable with $g'_k (t) =m_k (t) - \alpha '(t)$ for every $t>0$. If $t\in (0, t_{k-1}]$, by the definition of $m_k$ we have $g_{k}(t)=-\alpha (t)$ and thus

\[ g_{k}'(t)={-}\alpha'(t)={-}\frac{1}{2t}\max_{u\in S_t}\int_{\Omega}f^{{\ast}}(u)u\,{\textrm{d}}x<0,\quad \forall \, t\in (0, t_{k-1}]. \]

On the other hand, by (m), $\lim _{t\searrow t_{k-1}} m_k(t)= m_k(t_{k-1})=0$ and $(ii)$

\[ \limsup_{t\searrow t_{k-1}}g_{k}'(t)=m_k(t_{k-1})-\liminf_{t\searrow t_{k-1}}\alpha'(t)={-}\liminf_{t\searrow t_{k-1}}\alpha'(t)<0. \]

Finally, assuming hypothesis (mf), if $k=1$ then, by $g_{1}(0)=0$ and $(ii)$, we have

\[ \lim_{t\searrow 0}g_{1}'(t)=\frac{1}{2}\left[m(0)-\frac{\gamma}{\lambda_1}\right]<0. \]

The result follows.

3. Existence of $K$ further solutions for (P)

As in [Reference Santos Júnior and Siciliano13], our approach to prove the existence of a solution $v_k$ for each $k\in \{1,\ldots ,K\}$ is based on looking for critical points with norm in the interval $(t_{k-1}^{1/2}, t_k^{1/2})$ of the truncated functional $I_{k}: H_{0}^{1}(\Omega )\to {\rm I}\hskip -0.85mm{\rm R}$ defined by

(3.1)\begin{equation} I_{k}(u)=\frac{1}{2}M_{k}(\|u\|^{2})-\int_{\Omega}F^{{\ast}}(u)\,{\textrm{d}}x, \quad \forall\,u\in H_{0}^{1}(\Omega), \end{equation}

where $F^{\ast }$ is given by (1.2) and $M_{k}$ by (2.3).

Indeed, every critical point $v_{k}\in H_{0}^{1}(\Omega )$ of $I_k$ with $\| v_{k}\|^{2}\in (t_{k-1}, t_k)$ satisfies

\[ \int_{\Omega}\nabla v_{k}\nabla\varphi \,{\textrm{d}}x=\frac{1}{m(\|v_{k}\|^{2})}\int_{\Omega}f^{{\ast}}(v_{k})\varphi \,{\textrm{d}}x\geq 0,\quad \forall\,\varphi\in H_{0}^{1}(\Omega), \varphi\geq 0, \]

(because $m_k(\|v_{k}\|^{2})=m(\|v_{k}\|^{2})$). Hence, by the maximum principle, we conclude that $v_{k}\geq 0$ in $\Omega$. In addition, by choosing $\varphi = (v_{k}-s_{\ast })^{+}$ as a test function, we get

\[ 0\leq m(\|v_{k}\|^{2})\|(v_{k}-s_{{\ast}})^{+}\|^{2}=\int_{\Omega}f^{{\ast}}(v_{k})(v_{k}-s_{{\ast}})^{+}\,{\textrm{d}}x=0. \]

Since $m(\|v_{k}\|^{2})>0$ (by (m)), we conclude that $0\leq v_{k}\leq s_{\ast }$ is a weak solution of (P). By elliptic regularity and the maximum principle, $v_{k}>0$ is a classical solution.

Proof of case (i) of Theorem 1.2. Fix $k=2, \ldots , K$. It follows from (m) that $M_k\geq 0$ and thus by the definition of $\alpha _k$ that

\[ I_{k}(u)\geq{-}\int_{\Omega}F^{{\ast}}(u)\,{\textrm{d}}x\geq{-}\alpha_{k}, \]

for all $u\in H_{0}^{1}(\Omega )$ with $\|u\|\leq t_{k}^{1/2}$. Consequently, it makes sense to define the real number

\[ b_{k}:=\inf_{u\in H_{0}^{1}(\Omega), \|u\|\leq t_{k}^{1/2}}I_{k}(u) =\inf_{0< t\leq t_{k}}\left\{\inf_{u\in H_{0}^{1}(\Omega), \|u\|^{2}=t}I_{k}(u)\right\}=\inf_{0< t\leq t_{k}} g_{k}(t). \]

We deduce from (iii) of Lemma 2.1 that

(3.2)\begin{equation} b_{k}< g_{k}(t_{k-1})={-}\alpha_{k-1}. \end{equation}

To conclude the proof, it suffices to show that $b_k$ is a critical value of $I_k$ with an associated critical point $v_k \in H_{0}^{1}(\Omega )$ of norm $t_{k-1}^{1/2} < \|v_k\| < t_{k}^{1/2}$. To prove it, let $\{u_{n}\}\subset H_{0}^{1}(\Omega )$ be a minimizing sequence: i.e., such that

\[ \lim_{n\to \infty}I_{k}(u_{n})= b_{k} \quad \mbox{and} \quad \|u_{n}\|\leq t_{k}^{1/2}. \]

Hence, up to a subsequence, there exists $v_{k}\in H_{0}^{1}(\Omega )$ and $\tau _k\in [0, t_k]$ such that

\begin{align*} & u_{n}\rightharpoonup v_{k} \ \mbox{in} H_{0}^{1}(\Omega)\quad {\mbox{ and }\quad \|u_{n}\|\to \tau_{k}^{1/2}},\\ & \|v_{k}\|\leq \liminf_{n\to\infty}\|u_{n}\|=\tau_k^{1/2}\leq t_{k}^{1/2}, \end{align*}

and, by the weak continuity of $J$,

\[ \lim_{n\to \infty} \int_{\Omega}F^{{\ast}}(u_{n})\,{\textrm{d}}x =\int_{\Omega}F^{{\ast}}(v_{k})\,{\textrm{d}}x. \]

Observe that

(3.3)\begin{equation} b_k= \lim_{n\to \infty}I_{k}(u_{n})= \frac{1}{2} \lim_{n\to \infty} M_{k}(\|u_{n}\|^{2})- \lim_{n\to \infty} \int_{\Omega}F^{{\ast}}(u_{n})\,{\textrm{d}}x = \frac{1}{2} M_k(\tau_k ) - \int_{\Omega}F^{{\ast}}(v_k)\,{\textrm{d}}x . \end{equation}

Moreover, it is worth to point out that

(3.4)\begin{equation} \tau_k > t_{k-1}, \end{equation}

because in the contrary case we will have $M_{k}(\tau _k)=0$, $\|v_k\|^{2} \leq t_{k-1}$ and we will deduce by (3.3) and the definition of $\alpha _{k-1}$ that

\[ b_{k}={-}\int_{\Omega}F^{{\ast}}(v_{k})\,{\textrm{d}}x \geq{-}\alpha_{k-1}, \]

contradicting (3.2). Therefore, (3.4) holds. Now, we can prove that $\|v_k \|^{2}= \tau _k$. Indeed, we get a contradiction if we assume that $\| v_k\|^{2} < \tau _k$ because this implies by (m) and (3.4) that $M_k(\tau _k)>M_k(\| v_k\|^{2})$ and hence (3.3) gives

\[ b_k > \frac{1}{2} M_k(\| v_k\|^{2} ) - \int_{\Omega}F^{{\ast}}(v_{k})\,{\textrm{d}}x = I_k(v_k), \]

a contradiction with the definition of $b_k$. Therefore, $\{u_n\}$ is weakly convergent to $v_k$ with the norms $\{\| u_n\| \}$ converging also to $\|v_k\|$, which is equivalent to the strong convergence of $\{u_n\}$ to $v_k$. As a consequence,

\[ b_k = I_k(v_k) \]

and $v_k$ is a minimum point of $I_k$ on the ball with radio $\tau _k^{1/2}$ and centered at the origin. To prove that $v_k$ is a critical point of $I_k$ with level $b_k<0$ (by (3.2) and (i) of Lemma 2.1), it remains us to show that $\|v_{k}\|< t_k^{1/2}$. Suppose by contradiction that $\|v_{k}\|=t_{k}^{1/2}$. It follows from (A’) that

\[ b_{k}=I_{k}(v_{k})\geq\frac{1}{2}\int_{t_{k-1}}^{t_{k}}m(s)\,{\textrm{d}}s-\alpha_{k}\geq 0, \]

which contradicts that $b_k<0$ and therefore we have proved that $v_{k}$ is a critical point of $I_{k}$ with $t_{k-1}<\|v_{k}\|^{2}< t_{k}$.

Proof of case (ii) of Theorem 1.2. In this case, we can give exactly the same proof as in the previous case. The hypothesis (mf) is only required due to item (iii) of the Lemma 2.1.

Proof of case (iii) of Theorem 1.2. For every $k=1,\ldots , K$, the solution $v_k$, provided by items (i) and (ii) in Theorem 1.2, corresponds with a critical point of the truncated functional $I_{k}: H_{0}^{1}(\Omega )\to {\rm I}\hskip -0.85mm{\rm R}$ given by (3.1) with $I_{k}(v_{k})<0$. On the other hand, the solution $u_k$ provided in [Reference Santos Júnior and Siciliano13, Theorem 1.1] corresponds with a critical point of the truncated functional $I_{k}$ with $0< I_{k}(u_{k})$. Since $I_{k}(v_{k})<0< I_{k}(u_{k})$, we deduce that $v_{k } \not = u_{k}$ obtaining then 2 K positive solutions of (P).

4. Concentration phenomena of solutions

Throughout this section, we are going to denote by $f_{\lambda }$ a function $f$ which satisfies (f), (mf), (A$'$) and $F(s_{\ast })=\lambda$, where $\lambda$ is a positive parameter. We also denote the dependence on this parameter $\lambda$ of the functional $I_k=I_{k, \lambda }$ given by (3.1). It follows from previous section that problem (Pλ) has $K$ positive solutions $v_{k, \lambda }$, with $I_{k, \lambda }(v_k)<0$, $k=1, \ldots , K$. Before proving the main result of this section, we need the following lemma

Lemma 4.1 Let $b_{k, \lambda }=\inf _{u\in H_{0}^{1}(\Omega ), \|u\|\leq t_{k}^{1/2}}I_{k, \lambda }(u)$, where $I_{k, \lambda }$ is the truncated functional associated with the problem (Pλ) (and given in (3.1)). Then,

\[ \lim_{\lambda\searrow 0}b_{k, \lambda}= {\lim_{\lambda\searrow 0} \int_{\Omega}F^{{\ast}}_\lambda (v_{k, \lambda})\,{\textrm{d}}x =}0,\quad \forall\, k=1, \ldots, K. \]

Proof. It is sufficient to note that

\[ 0>b_{k, \lambda}=I_{k, \lambda}(v_{k, \lambda})>{-}\int_{\Omega}F^{{\ast}}_{\lambda}(v_{k, \lambda})\geq{-}\lambda |\Omega|, \]

where $F^{\ast }_\lambda$ is the primitive of $f^{\ast }_\lambda$. The proof will be finished by passing to the limit as $\lambda \searrow 0$ in the previous inequality.

Proof of case (iv) of Theorem 1.2. Assume by contradiction that for some fixed $k=1, \ldots , K$ there exists a sequence of positive numbers such that $\lambda _{n}\searrow 0$ and

\[ \|v_{k, \lambda_{n}}\|\to \tau_{k-1}^{1/2}\ \mbox{as} n\to \infty, \]

for some $t_{k-1}< \tau _{k-1}\leq t_{k}$. Since

\[ t_{k-1}<\|v_{k, \lambda_{n}}\|^{2}< t_{k}, \quad \forall\, n\in{\rm I}\hskip -0.85mm{\rm N}, \]

there exists $v_{k, 0}\in H_{0}^{1}(\Omega )$ such that, up to a subsequence, as $n$ tends to infinity, we obtain

\[ v_{k, \lambda_{n}}\rightharpoonup v_{k, 0}\ \mbox{in } H_{0}^{1}(\Omega). \]

By Lemma 4.1, we deduce that

\begin{align*} 0& =\lim_{n\to\infty}b_{k, \lambda_{n}}=\lim_{n\to\infty}I_{k, \lambda}(v_{k, \lambda_{n}}) = \lim_{n\to\infty}\left[\frac{1}{2}M_{k}(\|v_{k, \lambda_{n}}\|^{2})-\int_{\Omega}F^{{\ast}}_{\lambda_{n}}(v_{k, \lambda_{n}})\,{\textrm{d}}x\right]\\ & = \frac{1}{2}M_{k}(\tau_{k-1}) \end{align*}

and taking into account that by hypothesis (m), we have $M_{k}(\tau _{k-1})> M_{k}(t_{k-1})=0$, contradicting the previous inequality.

Acknowledgements

D. Arcoya was supported by MINECO-FEDER Grant MTM2015-68210-P and Junta de Andalucía FQM-116. J. R. Santos J. was partially supported by CNPq-Proc. 302698/2015-9 and CAPES-Proc. 88881.120045/2016-01, Brazil. A. Suárez was supported by MINECO/ FEDER Grant MTM2015-69875-P and Junta de Andalucía FQM-131. Finally, we would like to acknowledge the anonymous referee for her/his comments which have improved the present paper.

References

Alves, C. O., Corrêa, F. J. S. A. and Ma, T. F., Positive solutions for a quasilinear elliptic equation of Kirchhoff type, Comput. Math. Appl. 49 (2005), 8593.CrossRefGoogle Scholar
Ambrosetti, A. and Arcoya, D., Remarks on non homogeneous elliptic Kirchhoff equations, Nonlinear Differ. Equ. Appl. 23 (2016), Art. 57.CrossRefGoogle Scholar
Ambrosetti, A. and Arcoya, D., Positive solutions of elliptic Kirchhoff equations, Adv. Nonlinear Stud. 17 (2017), 316.CrossRefGoogle Scholar
Ambrosetti, A. and Prodi, G., A primer of nonlinear analysis (Cambridge University Press, Cambridge, 1993).Google Scholar
Brown, K. J. and Budin, H., On the existence of positive solutions for a class of semilinear elliptic boundary value problems, SIAM J. Math. Anal. 10 (1979), 875883.CrossRefGoogle Scholar
Clément, P. and Sweers, G., Existence and multiplicity results for a semilinear eigenvalue problem, Annali della Scuola Normale Superiore di Pisa, (4) 14 (1987), 97121.Google Scholar
He, X. and Zou, W., Infinitely many positive solutions for Kirchhoff-type problems, Nonlinear Anal. 70 (2009), 14071414.CrossRefGoogle Scholar
Hess, P., On multiple positive solutions of nonlinear elliptic eigenvalue problems, Comm. Partial Diff. Equ. 6 (1981), 951961.CrossRefGoogle Scholar
Kirchhoff, G., Mechanik (Teubner, Leipzig, 1883).Google Scholar
Ma, T. F. and Muñoz Rivera, J. E., Positive solutions for a nonlinear nonlocal elliptic transmission problem, Appl. Math. Lett. 16 (2003), 243248.CrossRefGoogle Scholar
Perera, K. and Zhang, Z., Nontrivial solutions of Kirchhoff-type problems via the Yang index, J. Differ. Equ. 221 (2006), 246255.CrossRefGoogle Scholar
Perera, K. and Zhang, Z., Sign changing solutions of Kirchhoff type problems via invariant sets of descent flow, J. Math. Anal. Appl. 317 (2006), 456463.Google Scholar
Santos Júnior, J. R. and Siciliano, G., Positive solutions for a Kirchhoff problem with vanishing nonlocal term, J. Differ. Equ. 265 (2018), 20342043.CrossRefGoogle Scholar