Hostname: page-component-745bb68f8f-s22k5 Total loading time: 0 Render date: 2025-02-06T20:23:46.102Z Has data issue: false hasContentIssue false

Singular elliptic equations with nonlinearities of subcritical and critical growth

Published online by Cambridge University Press:  27 July 2022

Matheus F. Stapenhorst*
Affiliation:
Departamento de Matemática, Rua Sérgio Buarque de Holanda, Universidade Estadual de Campinas, IMECC, 651 CEP 13083-859, Campinas, SP, Brazil (mstapenhorst@ime.unicamp.br)
Rights & Permissions [Opens in a new window]

Abstract

In this paper, we consider the problem $-\Delta u =-u^{-\beta }\chi _{\{u>0\}} + f(u)$ in $\Omega$ with $u=0$ on $\partial \Omega$, where $0<\beta <1$ and $\Omega$ is a smooth bounded domain in $\mathbb {R}^{N}$, $N\geq 2$. We are able to solve this problem provided $f$ has subcritical growth and satisfy certain hypothesis. We also consider this problem with $f(s)=\lambda s+s^{\frac {N+2}{N-2}}$ and $N\geq 3$. In this case, we are able to obtain a solution for large values of $\lambda$. We replace the singular function $u^{-\beta }$ by a function $g_\epsilon (u)$ which pointwisely converges to $u^{-\beta }$ as $\epsilon \rightarrow 0$. The corresponding energy functional to the perturbed equation $-\Delta u + g_\epsilon (u) = f(u)$ has a critical point $u_\epsilon$ in $H_0^{1}(\Omega )$, which converges to a non-trivial non-negative solution of the original problem as $\epsilon \rightarrow 0$.

Type
Research Article
Copyright
Copyright © The Author(s), 2022. Published by Cambridge University Press on Behalf of The Edinburgh Mathematical Society

1. Introduction

In this paper, we show that the problem

(1)\begin{equation} \left\{ \begin{array}{@{}l} -\Delta u={-}u^{-\beta}\chi_{\{u>0\}} + f(u) \text{ in } \Omega\\ u\geq 0, u\not\equiv 0\text{ in }\Omega\\ u = 0 \text{ on } \partial\Omega, \end{array} \right. \end{equation}

has a non-negative solution when $f$ has subcritical growth. The expression $\chi _{\{u>0\}}$ denotes the characteristic function corresponding to the set $\{x \in \Omega : u(x)>0\}$ and by convention $u^{-\beta }\chi _{\{u>0\}}=0$ if $u=0$. Hereafter, $\Omega \subset \mathbb {R}^{N}$, $N\geq 2$, is a bounded smooth domain, $0<\beta <1$ and $2^{*}=\frac {2N}{N-2}$ for $N\geq 3$.

By a solution of problem (1), we mean a function $u \in H_0^{1}(\Omega )$ such that

\[ u^{-\beta}\chi_{\{u>0\}} \in L^{1}_{loc}(\Omega) \]

and

\[ \int_\Omega \nabla u \nabla \varphi = \int_{\Omega \cap \{u>0\}} \left(\big({-}u^{-\beta} + f(u) \big) \; \varphi\right)\text{ for every }\varphi \in C^{1}_{c}(\Omega). \]

Here, $C_{c}^{1}(\Omega )$ stands for the functions belonging to $C^{1}(\Omega )$ with compact support.

We consider the perturbed problem

(2)\begin{equation} \left\{ \begin{array}{@{}l} -\Delta u +g_{\epsilon}(u)= f(u) \text{ in } \Omega\\ u\geq 0, u\not\equiv 0\text{ in }\Omega\\ u = 0 \text{ on }\partial\Omega, \end{array} \right. \end{equation}

where the perturbation $g_{\epsilon }$ is given by

(3)\begin{equation} g_{\epsilon}(s)=\left\{ \begin{array}{@{}l} \displaystyle\displaystyle\dfrac{s^{q}}{(s+\epsilon)^{q+\beta}} \text{ for } s\geq 0\\ 0\text{ for }s<0, \end{array} \right. \end{equation}

and $0< q<\frac {1}{2}$. We say that $u_{\epsilon }\in H_{0}^{1}(\Omega )$ is a weak solution of problem (2) if

(4)\begin{equation} \displaystyle\int_{\Omega}\nabla u_{\epsilon}\nabla v+\int_{\Omega}g_{\epsilon}(u_{\epsilon})v=\int_{\Omega}f(u_{\epsilon})v\text{ for all }v\in H_{0}^{1}(\Omega). \end{equation}

We define the functional $I_{\epsilon }: H_{0}^{1}(\Omega )\to \mathbb {R}$ associated to problem (2) by

(5)\begin{equation} I_{\epsilon}(u)=\displaystyle\frac{1}{2}\displaystyle\int_{\Omega}|\nabla u|^{2}+\int_{\Omega}G_{\epsilon}(u)-\int_{\Omega}F(u), \end{equation}

where $G_{\epsilon }(s)=\int _{0}^{s}g_{\epsilon }(t)\,{\rm d}t$ and $F(s)=\int _{0}^{s}f(t)\,{\rm d}t$. It turns out that certain solutions of problem (2) converge to a solution of problem (1). Initially, we make the following assumptions on $f$.

(6)\begin{equation} f\text{ is of class } C^{1,\nu}(0,\infty)\cap C[0,\infty)\text{ and }\sup_{s\in[0,1]}s^{1-q_{1}}|f'(s)|<\infty, \end{equation}

for some $0<\nu <1$ and $0< q_{1}<1$, and

(7)\begin{equation} f(s)=0\text{ for }s\leq 0. \end{equation}

We also assume that there exist constants $0<\epsilon _{0},\, \delta <1$ such that

(8)\begin{equation} g_{\epsilon_{0}}(s)\geq f(s)\text{ for all }s\leq\delta, \end{equation}

and that there exists a constant $C>0$ such that

(9)\begin{equation} |f(s)|\leq C(1+s^{p})\text{ for all }s\geq 0, \end{equation}

where $0< p<2^{*}-1$ ($0< p<\infty$ when $N=2$). We also assume that there exists constants $0<\theta <1/2$, $R>0$ and $c>0$ such that

(10)\begin{equation} (1-\theta) f(s)\leq \theta sf'(s)-c\text{ for }s\geq R, \end{equation}

and that there exists $\phi _{0}\in H_{0}^{1}(\Omega )\cap L^{\infty }(\Omega )$ such that

(11)\begin{equation} I_{\epsilon}(\phi_{0})<0\text{ for all }0<\epsilon<\epsilon_{0}. \end{equation}

Condition (11) holds provided

\[ \lim_{s\to\infty}\frac{F(s)}{s^{2}}=\infty. \]

Assumptions (6) and (9) imply that $I_{\epsilon }$ is of class $C^{1}$ and

(12)\begin{equation} I_{\epsilon}'(u)(v)=\displaystyle\int_{\Omega}\nabla u\nabla v+\int_{\Omega}g_{\epsilon}(u)v-\int_{\Omega}f(u)v, \text{ for all } u,v\in H_{0}^{1}(\Omega). \end{equation}

Our first result is as the following

Theorem 1.1 Assume that (6)(11) hold. Then, problem (1) has a non-trivial non-negative solution.

Examples: Let $\lambda >0$ and $\mu \geq 0$ be constants. Conditions (6)(11) hold for the following examples of $f$.

\begin{align*} \bullet f(s)& =\lambda s^{p}\pm \mu s^{q}\text{ with }0< q< p<2^{*}-1\text{ and }p>1;\\ \bullet f(s)& =\lambda s^{p}\pm \mu s^{q}\text{ with } N=3, 0< q< p\text{ and }1< p<5;\\ \bullet f(s)& =\lambda s^{p}\pm \mu s^{q}\text{ with } N=2, 0< q< p<\infty\text{ and }p>1;\\ \bullet f(s)& =\lambda s^{p}\pm \mu s^{q}\log s\text{ with }1< p<2^{*}-1\text{ and }0< q< p;\\ \bullet f(s)& =\lambda s^{p}\log s\pm \mu s^{q}\text{ with }1< p<2^{*}-1\text{ and }0< q< p. \end{align*}

Indeed, condition (10) will hold for $\frac {1}{1+p}<\theta <1/2$ and condition (8) will hold because for each $\lambda >0$ and $0<\tau <1$ there exists $0<\delta _{\tau,\lambda }<1$ and $0<\epsilon _{\tau,\lambda }<1$ such that

\[ g_{\epsilon}(s)\geq \lambda s^{\tau}\text{ for }0\leq s\leq\delta_{\tau,\lambda}\text{ and }0<\epsilon<\epsilon_{\tau,\lambda}, \]

provided $0< q<\tau$ in (3).

Next, we study problem (1) with $N\geq 3$ and $f(s)=\lambda s+s^{2^{*}-1}$, where $\lambda >0$. We prove

Theorem 1.2 Assume that $N\geq 3$ and $f(s)=\lambda s+s^{2^{*}-1}$. Then, there exists $\lambda _{0}>0$ such that problem (1) has a non-trivial non-negative solution for all $\lambda >\lambda _{0}$.

We recall the works of [Reference Ambrosetti, Brezis and Cerami1, Reference Brezis and Nirenberg5], where the authors studied the problem

(13)\begin{equation} \left\{ \begin{array}{@{}l} -\Delta u= \lambda u^{p}+u^{2^{*}-1} \text{ in } \Omega\\ u>0\text{ in }\Omega\\ u = 0 \text{ on }\partial\Omega. \end{array} \right. \end{equation}

In [Reference Brezis and Nirenberg5], it was assumed that $1< p<2^{*}-1$ in (13). They proved that problem (13) has a positive solution for every $\lambda >0$ provided $N\geq 4$. The same result holds if $N=3$ and $3< p<5$. In the case $N=3$ and $1< p\leq 3$, the authors proved in [Reference Brezis and Nirenberg5] that (13) possesses a positive solution provided $\lambda >0$ is sufficiently large. In [Reference Ambrosetti, Brezis and Cerami1], the authors studied problem (13) when $0< p<1$. They showed that there exists a constant $\Lambda _{1}>0$ such that problem (13) has at least two solutions if $0<\lambda <\Lambda _{1}$ and has no solution for $\lambda >\Lambda _{1}$. In [Reference Diaz, Morel and Oswald12], the authors studied the problem $-\Delta u=-u^{-\beta }+f(x)$ in $\Omega$, $u=0$ in $\partial \Omega$, the sub-supersolution method was used and positive solutions were obtained. Equation $-\Delta u+K(x)u^{-\beta }=\lambda u^{p}$ with $0< p<1$ and zero boundary condition was studied in [Reference Shi and Yao27], where $K$ was assumed to be of class $C^{2,\alpha }(\overline {\Omega })$. For more on singular problems with sublinear nonlinearities, see [Reference Papageorgiou and Rădulescu23, Reference Zhang30]. Theorem 1.1 asserts that the problem

(14)\begin{equation} \left\{ \begin{array}{@{}l} -\Delta u={-}u^{-\beta}\chi_{\{u>0\}} + \lambda u^{p}\pm\mu u^{q} \text{ in } \Omega\\ u\geq 0, u\not\equiv 0\text{ in }\Omega\\ u = 0 \text{ on }\partial\Omega, \end{array} \right. \end{equation}

is solvable for each $\lambda >0$ and $\mu \geq 0$, provided $0< q< p<2^{*}-1$ and $p>1$. Problems (13) and (14) are similar in essence, the latter being a singular version of the former. Theorem 1.1 should also be compared with the results of [Reference Dávila and Montenegro10, Reference Montenegro and Silva22], where the authors studied the problem

(15)\begin{equation} \left\{ \begin{array}{@{}l} -\Delta u={-}u^{-\beta}\chi_{\{u>0\}} + \lambda u^{p} \text{ in } \Omega\\ u\geq 0, u\not\equiv 0\text{ in }\Omega\\ u = 0 \text{ on } \partial\Omega, \end{array} \right. \end{equation}

with $\lambda >0$. In [Reference Dávila and Montenegro10], the authors assumed that $p>1$ and they obtained one solution of (15) for each $\lambda >0$. The case $p=1$ was also studied in [Reference Dávila and Montenegro10], and they obtained one solution for $\lambda >\lambda _{1}$, where $\lambda _{1}$ is the first eigenvalue of $-\Delta$. In [Reference Montenegro and Silva22], the authors assumed that $0< p<1$ and they obtained two distinct solutions of (15) for large values of $\lambda$. See also [Reference Dávila and Montenegro9], where the authors obtained sharper regularity results for solutions $u_{\lambda }$ of problem (15) with $0< p<1$. In this work, we consider general nonlinearities $f$ with subcritical growth, and we do not make use of parameters. Observe also that in Theorem 1.1, we make no assumptions on the sign of $f$.

Theorem 1.2 should be compared with the results of [Reference Figueiredo and Montenegro14], where the authors studied the problem

(16)\begin{equation} \left\{ \begin{array}{@{}l} -\Delta u={-}u^{-\beta}\chi_{\{u>0\}} + \lambda u^{p}+u^{\frac{N+2}{N-2}} \text{ in } \Omega\\ u\not\equiv 0\text{ in }\Omega\\ u = 0 \text{ on } \partial\Omega.\end{array} \right. \end{equation}

When $0< p<1$ in (16), the authors obtained a constant $\Lambda _{0}>0$ such that problem (16) has two distinct non-trivial and non-negative solutions for $0<\lambda <\Lambda _{0}$. If $1< p<\frac {N+2}{N-2}$, the authors obtained a constant $\Lambda _{0}^{*}>0$ such that problem (16) admits a solution provided $\lambda >\Lambda _{0}^{*}$. Theorem 1.2 addresses the case $p=1$ in (16).

Problems similar to (1) and (2) arise in the context of heterogeneous catalysis. Consider a reaction $R$ which converts a given gas to useful products, and suppose that $R$ occurs only in the presence of a catalyst that comes in the form of a porous pellet $\Omega$. For the pellet to be useful, the gas must diffuse inside it. In this context, two entities arise: the rate of reaction $k_{R}$ and the rate of diffusion $k_{D}$ of the gas in regions of $\Omega$. If $k_{D}$ is large compared to $k_{R}$, then the reaction occurs throughout $\Omega$ and no free boundary arises. However, when $k_{D}$ is small compared to $k_{R}$, then there are zones within $\Omega$ in which no reaction takes place, these are known as dead cores. The rates of adsorption $k_{a}$ and desorption $k_{d}$ of gas in the surface of the pellet must also be considered, for the equilibrium is reached when $k_{a}$ equals $k_{d}$. Let $A_{1},\, A_{2},\,\ldots A_{S}$ be chemical species involved in the reaction

\[ \sum_{j=1}^{S}\alpha_{j}A_{j}=0, \]

where $\alpha _{j}$ denote the number of molecules of $A_{j}$ being formed ($\alpha _{j}>0$) or consumed ($\alpha _{j}<0$) in $\Omega$. Then, under certain assumptions about the mechanism of the reaction, the concentration $c_{j}=c_{j}(x)$ of $A_{j}$ at $x\in \Omega$ satisfies the following elliptic equation

\[ \left\{ \begin{array}{@{}l} D_{j}\Delta c_{j} +\alpha_{j}\pi_{S}k_{R}=0 \text{ in } \Omega\\ c_{j} = c_{js} \text{ on } \partial\Omega, \end{array} \right. \]

for each $j\in \{1,\,2,\,\ldots,\,S\}$. Here, $D_{j}$ denotes the diffusion coefficient of $A_{j}$ and $\pi _{S}$ the catalytic area per unit volume. At equilibrium, the reaction rate $k_{R}$ can be calculated as a function of the concentrations $c_{j}$. Using a suitable change of variables (see [Reference Aris3], p.168), we get an equation of the form

(17)\begin{equation} \left\{ \begin{array}{@{}l} \Delta u =\lambda^{2} R(u) \text{ in } \Omega\\ u = 1 \text{ on }\partial\Omega, \end{array} \right. \end{equation}

where $\lambda >0$ is a constant called Thiele Modulus, $R:\mathbb {R}\to \mathbb {R}$ is a rational function and $0\leq u\leq 1$ represents a ‘normalized dimensionless concentration’. We see that equations (2) and (17) are similar in essence.

For more applications in catalysis and in other fields of research, such as biochemistry, see [Reference Aris3, Reference Diaz11]. See [Reference Friedman and Phillips15, Reference Phillips26] for studies of the free boundary of solutions of some elliptic equations.

Singular equations are related to phase field models, see [Reference Copetti and Elliott6, Reference Dal Passo, Giacomelli and Novick-Cohen8, Reference Elliot and Garcke13, Reference Gilardi and Rocca16]. For more results on singular elliptic equations, see [Reference Anello and Faraci2, Reference Bai, Gasiński and Papageorgiou4, Reference Hirano, Saccon and Shioji17, Reference Long, Sun and Wu19, Reference Montenegro and Queiroz21, Reference Papageorgiou and Smyrlis24, Reference Perera and Silva25, Reference Wang, Qin and Hu28].

Our paper is organized as follows. In § 2, we give some preliminary results. Next, we study problem (2) by considering two different scenarios; in § 3, we consider the subcritical case and in § 4, we study problem (2) with $f(s)=\lambda s+s^{\frac {N+2}{N-2}}$. In both cases, we show that the associated functional satisfy the assumptions of the Mountain Pass Theorem. We thus obtain solutions of problem (2). These solutions will be shown to be bounded in $H_{0}^{1}(\Omega )$ by a constant that does not depend on $\epsilon$. Such an estimate will be crucial in § 5, where we will establish regularity results for the solutions of problem (2) obtained in § 3 and § 4. In § 6, we prove Theorems 1.1 and 1.2.

2. Preliminary results

First, we show that critical points of the functional $I_{\epsilon }$ defined in (5) must be non-negative.

Lemma 2.1 Assume that (6), (7) and (9) hold. Let $u_{\epsilon }$ be a critical point of the functional $I_{\epsilon }$ defined by (5). Then $u_{\epsilon }\geq 0$ and $u_{\epsilon }$ is a weak solution of problem (2).

Proof of Lemma 2.1 Let $u_{\epsilon }^{-}=\max \{-u_{\epsilon },\,0\}$. Taking $v=u_{\epsilon }^{-}$ in (12) and using (7), we obtain

\[ 0=I_{\epsilon}'(u_{\epsilon})(u_{\epsilon}^{-})={-}\|u_{\epsilon}^{-}\|^{2}_{H_{0}^{1}(\Omega)}. \]

Hence, $u_{\epsilon }\geq 0$ and

\[ 0=I_{\epsilon}'(u_{\epsilon})(v)=\int_{\Omega}\nabla u_{\epsilon}\nabla v+\int_{\Omega}g_{\epsilon}(u_{\epsilon})v-\int_{\Omega}f(u_{\epsilon})v, \text{ for all }v\in H_{0}^{1}(\Omega). \]

Hence, (4) holds. This proves Lemma 2.1.

We will need estimates of the perturbation $g_{\epsilon }$ defined in (3). Note that

\[ g_{\epsilon}(s)=\frac{s^{q}}{(s+\epsilon)^{q+\beta}}\geq\frac{s^{q}}{(s+1)^{q+\beta}}=\frac{s^{q-\frac{1}{2}}}{(s+1)^{q+\beta}}s^{\frac{1}{2}}\text{ for }s\geq 0. \]

Hence,

\[ g_{\epsilon}(s)\geq \frac{1}{2^{q+\beta}}s^{q-\frac{1}{2}}s^{\frac{1}{2}}\text{ for }0\leq s<1. \]

Therefore, from the fact that $0< q<\frac {1}{2}$, it follows that, for each $M>0$, there exists $\overline {\delta }=\overline {\delta }(M)<1$ such that

\[ g_{\epsilon}(s)\geq Ms\text{ for }0\leq s<\overline{\delta}<1. \]

We thus obtain

(18)\begin{equation} G_{\epsilon}(s)=\displaystyle\int_{0}^{s}g_{\epsilon}(t)\,{\rm d}t\geq \int_{0}^{s}M t\,{\rm d}t=\displaystyle\frac{M}{2}s^{2}\text{ for }0\leq s<\overline{\delta}<1. \end{equation}

Observe that

\[ g_{\epsilon}'(s)=\frac{q s^{q-1}(s+\epsilon)^{q+\beta}-(q+\beta)s^{q}(s+\epsilon)^{q+\beta-1}}{(s+\epsilon)^{2(q+\beta)}}. \]

Hence,

(19)\begin{equation} sg_{\epsilon}'(s)=\displaystyle\frac{q s^{q}}{(s+\epsilon)^{q+\beta}}-\frac{(q+\beta)s^{q+1}}{(s+\epsilon)^{q+\beta+1}}. \end{equation}

The following lemma will play a crucial role in § 4.

Lemma 2.2

\[ G_{\epsilon}(s)\geq\frac{1}{2}g_{\epsilon}(s)s,\text{ for every }s\geq 0. \]

Proof of Lemma 2.2 Indeed, let $\widetilde {B}_{\epsilon }(s)=G_{\epsilon }(s)-\frac {1}{2}g_{\epsilon }(s)s$. We have that $B_{\epsilon }(0)=0$ and

\[ \widetilde{B}_{\epsilon}'(s)=g_{\epsilon}(s)-\frac{1}{2}g_{\epsilon}(s)-\frac{s}{2}g_{\epsilon}'(s)=\frac{1}{2}\left(g_{\epsilon}(s)-sg_{\epsilon}'(s)\right). \]

Therefore, $\widetilde {B}_{\epsilon }'(s)\geq 0$ if and only if

\[ g_{\epsilon}(s)\geq sg_{\epsilon}'(s). \]

From (19), this inequality will be true if

(20)\begin{equation} \displaystyle\frac{s^{q}}{(s+\epsilon)^{q+\beta}}\geq \frac{q s^{q}}{(s+\epsilon)^{q+\beta}}. \end{equation}

Since $q<1/2$, (20) holds for each $s\geq 0$. This proves Lemma 2.2.

Now, we show that a version of the Ambrosetti–Rabinowitz condition holds. We define

\[ j_{\epsilon}(s)=f(s)-g_{\epsilon}(s)\text{ for }s\in\mathbb{R}. \]

Consequently,

\[ I_{\epsilon}(u)=\frac{1}{2}\int_{\Omega}|\nabla u|^{2}-\int_{\Omega}J_{\epsilon}(u), \]

where $J_{\epsilon }(s)=\int _{0}^{s}j_{\epsilon }(t)\,{\rm d}t$. For simplicity of notation, we denote $J_{\epsilon }$ and $j_{\epsilon }$ merely by $J$ and $j$ respectively.

Lemma 2.3 Suppose that (6) and (10) hold. Let $0<\theta <1/2$ be given by (10). There exists a constant $\overline {R}>0$ such that

\[ J(s)\leq\theta s j(s)\text{ for }s\geq \overline{R}. \]

Proof of Lemma 2.3 Let $B_{\epsilon }(s)=J(s)-\theta s j(s)$. We have

\[ B_{\epsilon}'(s)=(1-\theta)j(s)-\theta sj'(s). \]

Hence,

\[ B_{\epsilon}'(s)={-}(1-\theta)g_{\epsilon}(s)+\theta sg_{\epsilon}'(s)+(1-\theta)f(s)-\theta s f'(s). \]

From (19) we obtain

\[ |sg_{\epsilon}'(s)|\leq q|s|^{-\beta}+(q+\beta)|s|^{-\beta}\to 0\text{ as }s\to\infty. \]

It is also clear that

\[ (1-\theta)g_{\epsilon}(s)\to 0\text{ as }s\to\infty\text{ uniformly for }\epsilon. \]

Hence, for each $0<\tau <1$ there exists $R_{\tau }>0$ that does not depend on $\epsilon$ such that

\[ |(1-\theta)g_{\epsilon}(s)|+|sg_{\epsilon}'(s)|<\tau\text{ for }s\geq R_{\tau}. \]

Therefore,

\[ B_{\epsilon}'(s)<\tau+(1-\theta)f(s)-\theta s f'(s).\text{ for }s\geq R_{\tau}. \]

Consequently, from (10), we get

\[ B_{\epsilon}'(s)<\tau-c\text{ for }s\geq \max\{R,R_{\tau}\}, \]

where $c>0$ and $R>0$ are given by (10). Choosing $\tau =c/2$, we get

(21)\begin{equation} B_{\epsilon}'(s)<{-}\displaystyle\frac{c}{2}\text{ for }s\geq R_{2}, \end{equation}

where

\[ R_{2}=\max\{R,R_{c/2}\}. \]

Note that

\[ B_{\epsilon}(R_{2})\leq C_{1} \]

where

\[ C_{1}=|F(R_{2})|+|\theta R_{2}f(R_{2})|+\theta R_{2}^{1-\beta}. \]

Therefore, (21) implies that there exists a constant $T>0$ such that

\[ B_{\epsilon}(s)\leq{-}\frac{cs}{2}+T\text{ for }s\geq R_{2}. \]

Hence, $B_{\epsilon }(s)\leq 0$ for $s\geq \max \{R_{2},\,2T/c\}$. This proves Lemma 2.3.

Let $\phi _0 \in H^{1}_{0}(\Omega )\cap L^{\infty }(\Omega )$ be given by (11). We have

Lemma 2.4 Assume that (6), (9) and (11) hold. There exist a constant $a_{2}>0$ that does not depend on $\epsilon$ such that

(22)\begin{equation} \sup_{0\leq s\leq 1}I_{\epsilon}(s\phi_{0})< a_{2}\text{ for every }0<\epsilon<\epsilon_{0}. \end{equation}

Proof of Lemma 2.4 We have

\[ I_{\epsilon}(s\phi_{0})\leq\frac{s^{2}}{2}\|\phi_{0}\|_{H_{0}^{1}(\Omega)}+\int_{\Omega} G_{\epsilon}(s\phi_{0})-\int_{\Omega} F(s\phi_{0})\,{\rm d}x,\text{ for every }s\geq 0. \]

Consequently, we get

\[ I_{\epsilon}(s\phi_{0})\leq\frac{s^{2}}{2}\|\phi_{0}\|_{H_{0}^{1}(\Omega)}+\frac{1}{1-\beta}\int_{\Omega}|s\phi_{0}|^{1-\beta}-\int_{\Omega} F(s\phi_{0})\,{\rm d}x,\text{ for every }s\geq 0. \]

We conclude that

\[ \sup_{0\leq s\leq 1}I_{\epsilon,\lambda}(s\phi_{0})< a_{2}, \]

where

\[ a_{2}=\frac{1}{2}\|\phi_{0}\|_{H_{0}^{1}(\Omega)}+|\Omega|\left(\frac{\sup_{\Omega}|\phi_{0}|^{1-\beta}}{1-\beta}+\sup_{0\leq s\leq\sup\phi_{0}}|F(s)|\right). \]

This proves (22). We have proved Lemma 2.4.

3. Existence of solution of the perturbed subcritical problem

Throughout this section, we will assume that $f$ satisfies (9) with $0< p<2^{*}-1$. Our aim is to show that problem (2) has a non-negative non-trivial solution. We recall that given a Banach space $E$ and a functional $\Psi \in C^{1}(E;\mathbb {R})$, we say that a sequence $(u_n)$ in $E$ is a Palais–Smale sequence of $\Psi$ if there exists $c\in \mathbb {R}$ such that $\Psi (u_n)\to c$ and $\|\Psi '(u_{n})\|\to 0$ as $n\to \infty$. We say that $\Psi$ satisfies the Palais–Smale condition if every Palais–Smale sequence of $\Psi$ has a convergent subsequence.

Lemma 3.1 Assume that (6)(10) hold. Fix $0<\epsilon <1$ and let $(u_{n}^{\epsilon })$ be a Palais–Smale sequence for $I_{\epsilon }$ in $H_{0}^{1}(\Omega )$. Assume that there exists a constant $C>0$ that does not depend on $\epsilon$ such that

(23)\begin{equation} |I_{\epsilon}(u_{n}^{\epsilon})|< C\text{ for all }n\in\mathbb{N}. \end{equation}

Then, there exists $D>0$ that does not depend on $\epsilon$ such that

(24)\begin{equation} \|u_{n}^{\epsilon}\|_{H_{0}^{1}(\Omega)}< D\text{ for all }n\in\mathbb{N}. \end{equation}

Furthermore, there exists $u_{\epsilon }\in H_{0}^{1}(\Omega )$ such that, up to a subsequence, $u_{n}^{\epsilon }\to u_{\epsilon }$ strongly in $H_{0}^{1}(\Omega )$. Consequently, $u_{\epsilon }$ is a critical point of $I_{\epsilon }$.

Proof of Lemma 3.1 Throughout this proof, we denote $\|\cdot \|_{H_{0}^{1}(\Omega )}$ by $\|\cdot \|$. Let $(u_{n}^{\epsilon })_{n\in \mathbb {N}}$ be a Palais–Smale sequence for $I_{\epsilon }$ satisfying (23). Consequently,

(25)\begin{equation} \displaystyle\frac{1}{2}\|u_{n}^{\epsilon}\|^{2}\leq C+\displaystyle\int_{\Omega}J(u_{n}^{\epsilon})\,{\rm d}x\text{ for all }n\in\mathbb{N}, \end{equation}

and there is a sequence $\tau _{n}\rightarrow 0$ such that

(26)\begin{equation} \biggl|\displaystyle\int_{\Omega}\nabla u_{n}^{\epsilon}\nabla w\,{\rm d}x-\int_{\Omega}j(u_{n}^{\epsilon})w\,{\rm d}x\biggl|\leq \tau_{n}\|w\| \text{ for each } w\in H_{0}^{1}(\Omega). \end{equation}

Let $0<\theta <1/2$ be given by (10). From Lemma 2.3, there is a constant $\overline {R}>0$ that does not depend on $\epsilon$ such that

\[ J(t)\leq\theta t j(t) \text{ for } t\geq \overline{R}. \]

Since there exists $D_{1}>0$ that does not depend on $\epsilon$ such that

\[ \sup_{0\leq s\leq \overline{R}}\max\{|J(s)|,|sj(s)|\}< D_{1}, \]

we may find a constant $D_{2}>0$ such that

\[ J(u_{n}^{\epsilon})< D_{2}+\theta u_{n}^{\epsilon}j(u_{n}^{\epsilon}). \]

We know from (25) that there is a constant $D_{3}>0$ such that

\[ \frac{1}{2}\|u_{n}^{\epsilon}\|^{2}\leq D_{3}+\theta\int_{\Omega} u_{n}^{\epsilon}j(u_{n}^{\epsilon})\,{\rm d}x. \]

Taking $w=u_{n}^{\epsilon }$ in (26), we also conclude that

\[ \int_{\Omega}j(u_{n}^{\epsilon})u_{n}^{\epsilon}\,{\rm d}x<\|u_{n}^{\epsilon}\|^{2}+\tau_{n}\|u_{n}^{\epsilon}\|. \]

Hence,

\[ \frac{1}{2}\|u_{n}^{\epsilon}\|^{2}< D_{3}+\theta\|u_{n}^{\epsilon}\|^{2}+\tau_{n}\theta\|u_{n}^{\epsilon}\|. \]

Since $0<\theta <\frac {1}{2}$, (24) follows. Consequently, there exists $u_{\epsilon }\in H_{0}^{1}(\Omega )$ such that $u_{n}^{\epsilon }\rightharpoonup u_{\epsilon }$ weakly in $H_{0}^{1}(\Omega )$. Since $J_{\epsilon }$ has subcritical growth at infinity (see [Reference Costa7], Theorem 3.4 and Remark 2.2.1), we conclude that, up to a subsequence, $u_{n}^{\epsilon }\to u_{\epsilon }$ strongly in $H_{0}^{1}(\Omega )$. Since $I_{\epsilon }'(u_{n}^{\epsilon })\to 0$ as $n\to \infty$ and $I_{\epsilon }$ is of class $C^{1}$, we conclude that $I_{\epsilon }'(u_{\epsilon })=0$. This proves the result.

Now, we obtain one solution for problem (2).

Proposition 3.2 Assume that (6)(11) hold and let $a_{2}>0$ be given by Lemma 2.4. Then, there is a non-negative solution $u_{\epsilon }$ of problem (2) and there exist constants $a_{1}>0$ and $D>0$ that do not depend on $\epsilon$ such that

\[ 0< a_{1}\leq I_{\epsilon}(u_{\epsilon})\leq a_{2}, \]

and

\[ \|u_{\epsilon}\|_{H_{0}^{1}(\Omega)}< D. \]

Proof of Proposition 3.2 Let $\delta >0$ and $\epsilon _{0}$ be given by (8). Note that

\begin{align*} I_{\epsilon}(u)& =\frac{1}{2}\int_{\Omega}|\nabla u|^{2}+\int_{\{u\leq\delta\}}(G_{\epsilon}(u)-F(u))\,{\rm d}x\\& \quad +\int_{\{u>\delta\}}(G_{\epsilon}(u)-F(u))\,{\rm d}x\text{ for every }u\in H_{0}^{1}(\Omega). \end{align*}

The fact that $g_{\epsilon }$ is monotone in $\epsilon$ implies that

\[ g_{\epsilon}(s)\geq f(s)\text{ for all }0\leq s<\delta\text{ and }0<\epsilon<\epsilon_{0}. \]

Using the fact that $G_{\epsilon }\geq 0$, we get

\[ I_{\epsilon}(u)\geq\frac{1}{2}\int_{\Omega}|\nabla u|^{2}-\int_{\{u>\delta\}}F(u)\text{ for every }u\in H_{0}^{1}(\Omega). \]

From (9), we have

\[ |F(s)|\leq\int_{0}^{s}|f(t)|\,{\rm d}t\leq C\int_{0}^{s}(1+t^{p})\,{\rm d}t=Cs+\frac{Cs^{1+p}}{1+p}. \]

Consequently,

\[ I_{\epsilon}(u)\geq\frac{1}{2}\int_{\Omega}|\nabla u|^{2}-C\int_{\{u>\delta\}}u-\frac{C}{1+p}\int_{\{u>\delta\}}u^{p+1}\text{ for every }u\in H_{0}^{1}(\Omega). \]

We conclude that there exists $\widetilde {C}>0$ such that

\[ I_{\epsilon}(u)\geq\frac{1}{2}\int_{\Omega}|\nabla u|^{2}-\widetilde{C}\int_{\Omega}|u|^{\sigma}\text{ for every }u\in H_{0}^{1}(\Omega), \]

where $\sigma >2$ is chosen such that $1+p<\sigma <2^{*}$. The Sobolev embedding implies that there is a constant $C_{3}>0$ such that

\[ I_{\epsilon}(u)\geq\frac{1}{2}\|u\|^{2}_{H_{0}^{1}(\Omega)}-C_{3}\|u\|^{\sigma}_{H_{0}^{1}(\Omega)}. \]

Therefore,

\[ I_{\epsilon}(u)\geq\frac{1}{4}\|u\|^{2}_{H_{0}^{1}(\Omega)}\text{ for }\|u\|_{H_{0}^{1}(\Omega)}\leq\rho, \]

where

\[ \rho=\bigg(\frac{1}{4C_{3}}\bigg)^{\frac{1}{\sigma-2}}. \]

Also,

\[ I_{\epsilon}(u)\geq a_{1}\text{ for }\|u\|_{H_{0}^{1}(\Omega)}=\rho, \]

where

\[ a_{1}=\frac{\rho^{2}}{4}. \]

Let $\phi _{0}$ be given by (11) and $\Gamma =\{\gamma \in C([0,\,1],\, H_{0}^{1}(\Omega )):\gamma (0)=0,\, \gamma (1)=\phi _{0}\}$. We know from (23) that $I_{\epsilon }(\phi _{0})<0$. Consequently, we may apply the Mountain Pass Theorem [Reference Willem29], page 12) to conclude that there is a Palais–Smale sequence $(u_{n}^{\epsilon }$ for $I_{\epsilon }$ in $H_{0}^{1}(\Omega )$ and a number

\[ c_{\epsilon}=\inf_{\gamma\in\Gamma}\sup_{s\in[0,1]}I_{\epsilon}(\gamma(s)), \]

such that

\[ \lim_{n\rightarrow\infty}I_{\epsilon}(u_{n}^{\epsilon})=c_{\epsilon}\text{ and }\lim_{n\rightarrow\infty}I_{\epsilon}'(u_{n}^{\epsilon})=0. \]

From Lemma 2.4, we know that $a_{1}\leq c_{\epsilon }\leq a_{2}$. From Lemma 3.1, we conclude that there exist $D>0$ (that does not depend on $\epsilon$) and $u_{\epsilon }\in H_{0}^{1}(\Omega )$ such that, up to a subsequence, $u_{n}^{\epsilon }\to u_{\epsilon }$ strongly in $H_{0}^{1}(\Omega )$ and

\[ \|u_{\epsilon}\|_{H^{1}_{0}(\Omega)}< D. \]

Consequently, $I_{\epsilon }'(u_{\epsilon })=0$ and

\[ a_{1}\leq I_{\epsilon}(u_{\epsilon})< a_{2}. \]

This proves the result.

4. Existence of solution of the perturbed problem when $p=2^{*}-1$

In this section, we study problem (2) with $f(s)=\lambda s+s^{2^{*}-1}$ for $s\geq 0$. We assume that $f(s)=0$ for $s\leq 0$. This function satisfies (6), (7), (8) and (10). The difficulty here is that $f$ no longer satisfies (9), so that Lemma 3.1 does not hold. The associated functional then becomes

(27)\begin{align} I_{\epsilon,\lambda}(u)=\displaystyle\int_{\Omega}|\nabla u|^{2}\,{\rm d}x+\int_{\Omega}G_{\epsilon}(u)-\displaystyle\frac{\lambda}{2}\int_{\Omega}(u^{+})^{2}\,{\rm d}x-\frac{1}{2^{*}}\int_{\Omega}(u^{+})^{2^{*}}\text{ defined for }u\in H_{0}^{1}(\Omega), \end{align}

where $u^{+}=\max \{u,\,0\}$. The functional $I_{\epsilon,\lambda }$ is of class $C^{1}$ and

(28)\begin{align} I_{\epsilon,\lambda}'(u)(v)& =\displaystyle\int_{\Omega}\nabla u\nabla v\,{\rm d}x+\int_{\Omega}g_{\epsilon}(u)v\,{\rm d}x\nonumber\\& \quad -\lambda\int_{\Omega}(u^{+})v\,{\rm d}x-\int_{\Omega}(u^{+})^{2^{*}-1}v\,{\rm d}x\text{ for all }u,v\in H_{0}^{1}(\Omega). \end{align}

The same argument given by Lemma 2.1 implies that critical points of $I_{\epsilon,\lambda }$ are non-negative solutions of problem (2). Observe also that zero is a local minimum of the functional $I_{\epsilon,\lambda }$. Indeed, let $0<\overline {\delta }<1$ be given by (18). Note that

\[ I_{\epsilon,\lambda}(u)\geq\frac{1}{2}\int_{\Omega}|\nabla u|^{2}+\int_{\{u<\overline{\delta}\}}G_{\epsilon}(u)-\frac{\lambda}{2}\int_{\Omega}(u^{+})^{2}-\frac{1}{2^{*}}\int_{\Omega}(u^{+})^{2^{*}}\text{ for every }u\in H_{0}^{1}(\Omega) \]

Choosing $M=\lambda$ in (18), we obtain

\[ I_{\epsilon,\lambda}(u)\geq\frac{1}{2}\int_{\Omega}|\nabla u|^{2}-\frac{\lambda}{2}\int_{\{u>\overline{\delta}\}}u^{2}-\frac{1}{2^{*}}\int_{\Omega}(u^{+})^{2^{*}}\text{for every }u\in H_{0}^{1}(\Omega). \]

Observe that there exists a constant $C_{1}>0$ such that

\[ s^{2}\leq C_{1}s^{2^{*}}\text{ for }s\geq\overline{\delta}. \]

Hence, there exists a constant $C_{2}>0$ such that

\[ I_{\epsilon,\lambda}(u)\geq\frac{1}{2}\|u\|^{2}_{H_{0}^{1}(\Omega)}-C_{2}\int_{\Omega}|u|^{2^{*}}\text{ for every }u\in H_{0}^{1}(\Omega). \]

Consequently, the Sobolev embedding implies that

\[ I_{\epsilon,\lambda}(u)\geq\frac{1}{2}\|u\|^{2}_{H_{0}^{1}(\Omega)}-C_{3}\|u\|^{2^{*}}_{H_{0}^{1}(\Omega)}\text{ for all }u\in H_{0}^{1}(\Omega) \]

We conclude that

\[ I_{\epsilon,\lambda}(u)\geq\frac{1}{4}\|u\|^{2}_{H_{0}^{1}(\Omega)}\text{ for }\|u\|_{H_{0}^{1}(\Omega)}\leq\rho, \]

where

\[ \rho=\bigg(\frac{1}{4C_{3}}\bigg)^{\frac{1}{2^{*}-2}}. \]

Also,

(29)\begin{equation} I_{\epsilon,\lambda}(u)\geq a_{1}\text{ for }\|u\|_{H_{0}^{1}(\Omega)}=\rho, \end{equation}

where

\[ a_{1}=\frac{\rho^{2}}{4}. \]

We now show that there exists $\phi \in H_{0}^{1}(\Omega )\cap L^{\infty }(\Omega )$ such that

\[ I_{\epsilon,\lambda}(\phi)<0. \]

Indeed, let $\phi _1 \in H^{1}_{0}(\Omega )$ be the first eigenfunction of the operator $-\Delta$ with $\|\phi _1\|_{H^{1}_{0}(\Omega )}= 1$. We have

Lemma 4.1 There exist constants $N_{0}>0,$ $a_{2}>0$ and $b_{1}>0$ such that

(30)\begin{equation} I_{\epsilon,\lambda}( N_{0}\phi_{1})<{-}b_{1}<0,\text{ for every }\,0<\epsilon<1, \end{equation}

and

(31)\begin{equation} \sup_{0\leq s\leq 1}I_{\epsilon,\lambda}(sN_{0}\phi_{1})< a_{2}\text{ for every }\lambda>0,\, 0<\epsilon<1. \end{equation}

Moreover, these constants do not depend on $\lambda$.

Proof of Lemma 2.4 For each $t>0$, we have

\[ I_{\epsilon,\lambda}(t\phi_{1})=\frac{t^{2}}{2}+\int_{\Omega} G_{\epsilon}(t\phi_{1})-\frac{\lambda t^{2}}{2}\int_{\Omega}\phi_{1}^{2}\,{\rm d}x-\frac{t^{2^{*}}}{2^{*}}\int \phi_{1}^{2^{*}}\,{\rm d}x. \]

From the fact that $G_{\epsilon }(s)\leq \frac {s^{1-\beta }}{1-\beta }$ for all $s\geq 0$, we get

(32)\begin{equation} I_{\epsilon,\lambda}(t\phi_{1})\leq\displaystyle\frac{t^{2}}{2}+\frac{t^{1-\beta}}{1-\beta}\displaystyle\int_{\Omega} \phi_{1}^{1-\beta}-\frac{t^{2^{*}}}{2^{*}}\int \phi_{1}^{2^{*}}\,{\rm d}x. \end{equation}

Since $2^{*}>2>1-\beta$, inequality (30) then follows by taking $t$ large enough in (32). We also have

\[ I_{\epsilon,\lambda}(sN_{0}\phi_{1})\leq\frac{s^{2}N_{0}^{2}}{2}+\int_{\Omega} G_{\epsilon}(s N_{0}\phi_{1})-\frac{s^{2^{*}}N_{0}^{2^{*}}}{2^{*}}\int \phi_{1}^{2^{*}}\,{\rm d}x,\text{ for every }s\geq 0. \]

Consequently, we get

\[ I_{\epsilon,\lambda}(sN_{0}\phi_{1})\leq \frac{s^{2} N_{0}^{2}}{2}+\frac{s^{1-\beta}N_{0}^{1-\beta}}{1-\beta}\int_{\Omega} \phi_{1}^{1-\beta},\text{ for every }s\geq 0. \]

We conclude that

\[ \sup_{0\leq s\leq 1}I_{\epsilon,\lambda}(sN_{0}\phi_{1})< a_{2}, \]

where

\[ a_{2}=\frac{N_{0}^{2}}{2}+\frac{N_{0}^{1-\beta}}{1-\beta}\int_{\Omega} \phi_{1}^{1-\beta}. \]

This proves (31). We have proved Lemma 4.1.

Lemma 4.1, (29) and the Mountain Pass Theorem imply that there is a sequence $(u_{n}^{\epsilon })$ in $H_{0}^{1}(\Omega )$ and a number

(33)\begin{equation} c_{\epsilon,\lambda}=\inf_{\gamma\in\Gamma}\sup_{s\in[0,1]}I_{\epsilon,\lambda}(\gamma(s)), \end{equation}

such that

(34)\begin{equation} \lim_{n\rightarrow\infty}I_{\epsilon,\lambda}(u_{n}^{\epsilon})=c_{\epsilon,\lambda}\text{ and }\lim_{n\rightarrow\infty}I_{\epsilon,\lambda}'(u_{n}^{\epsilon})=0, \end{equation}

where $\Gamma =\{\gamma \in C([0,\,1],\, H_{0}^{1}(\Omega )):\gamma (0)=0,\, \gamma (1)=N_{0}\phi _{1}\}$. It is clear that the function $f(s)=\lambda s+s^{2^{*}-1}$ satisfies (10). Consequently, the same computations developed in the proof of Lemma 2.3 imply that there exists $0<\theta <1/2$ such that

\[ J(s)\leq\theta sj(s), \]

where $j(s)=\lambda s+s^{2^{*}-1}-g_{\epsilon }(s)$ and $J(s)=\int _{0}^{s}j(t)\,{\rm d}t$. Consequently, by a similar argument given in the proof of Lemma 3.1, we obtain a constant $D>0$ such that

(35)\begin{equation} \|u_{n}^{\epsilon}\|_{H_{0}^{1}(\Omega)}< D\text{ for all }n\in\mathbb{N},0<\epsilon<1. \end{equation}

Furthermore, we have

Lemma 4.2 Let $c_{\lambda,\epsilon }$ be given by (33). Then

(36)\begin{equation} \lim_{\lambda\to\infty}c_{\lambda,\epsilon}=0\text{ uniformly on }0<\epsilon<1. \end{equation}

Proof of Lemma 4.2 Fix $0<\epsilon <1$ and let $t_{\lambda,\epsilon }\geq 0$ be such that

\[ I_{\epsilon,\lambda}(t_{\lambda,\epsilon}\phi_{1})=\max_{0\leq t\leq 1}I_{\epsilon,\lambda}(t N_{0}\phi_{1}). \]

From Lemmas 2.4 and (29), we get

\[ I_{\epsilon,\lambda}(t\phi_{1})>0\text{ for }0< t<\rho\quad\text{ and }I_{\epsilon,\lambda}(N_{0}\phi_{1})<0. \]

Hence, $0< t_{\lambda,\epsilon }<1$. Consequently,

\[ \frac{d}{{\rm d}t}I_{\epsilon,\lambda}(t N_{0}\phi_{1})\biggl|_{t=t_{\lambda,\epsilon}}=0. \]

Equivalently, from (28),

(37)\begin{equation} \begin{aligned} 0=I_{\epsilon,\lambda}'(t_{\lambda,\epsilon}N_{0}\phi_{1})(N_{0}\phi_{1}) & =N_{0}^{2}t_{\lambda,\epsilon}\displaystyle\int_{\Omega}|\nabla\phi_{1}|^{2}+N_{0}^{1+q}t_{\lambda,\epsilon}^{q}\int_{\Omega}\displaystyle\frac{\phi_{1}^{1+q}}{(N_{0}t_{\lambda,\epsilon}\phi_{1}+\epsilon)^{q+\beta}}\\ & \quad -N_{0}^{2}\lambda t_{\lambda,\epsilon}\displaystyle\int_{\Omega}\phi_{1}^{2}-N_{0}^{2^{*}}t_{\lambda,\epsilon}^{2^{*}-1}\int_{\Omega}\phi_{1}^{2^{*}}. \end{aligned} \end{equation}

Fix a sequence $(\lambda _{n})$ in $\mathbb {R}$ such that $\lambda _{n}\to \infty$. Since $0< t_{\lambda _{n},\,\epsilon }<1$, we know that for each $0<\epsilon <1$ there exists an element $0\leq t_{0,\epsilon }\leq 1$ such that

\[ t_{\lambda_{n},\epsilon}\to t_{0,\epsilon}\text{ as }n\to\infty. \]

We will show that $t_{0,\epsilon }=0$. Indeed, from (37) there exists a constant $M_{0}>0$ that does not depend on $\lambda$ nor on $\epsilon$ such that

(38)\begin{equation} \lambda_{n}t^{2}_{\lambda_{n},\epsilon}\displaystyle\int_{\Omega}\phi_{1}^{2}\leq t_{\lambda_{n},\epsilon}^{2}\int_{\Omega}|\nabla\phi_{1}|^{2}+t_{\lambda_{n},\epsilon}^{1-\beta}\int_{\Omega}\phi_{1}^{1-\beta}\leq M_{0}. \end{equation}

Letting $n\to \infty$ in (38), it follows that $t_{\lambda _{n},\,\epsilon }\to 0$ as $n\to \infty$ uniformly on $\epsilon$. Hence, $t_{0,\epsilon }=0$. Consequently,

\[ 0< c_{\lambda_{n},\epsilon}\leq\max_{0\leq t\leq 1}I_{\epsilon,\lambda_{n}}(tN_{0}\phi_{1})=I_{\epsilon,\lambda_{n}}(t_{\lambda_{n},\epsilon}\phi_{1})\leq t_{\lambda_{n},\epsilon}^{2}\int_{\Omega}|\nabla\phi_{1}|^{2}\,{\rm d}x+t_{\lambda_{n},\epsilon}^{1-\beta}\int_{\Omega}\phi_{1}^{1-\beta}\,{\rm d}x. \]

Letting $n\to \infty$, we obtain

\[ \lim_{n\to\infty}c_{\lambda_{n},\epsilon}=0\text{ uniformly on }0<\epsilon<1. \]

Since the sequence $(\lambda _{n})$ was arbitrarily chosen, (36) follows.

Consequently, there exist $\lambda _{0}>0$ and $0<\epsilon _{0}<1$ such that

(39)\begin{equation} c_{\lambda,\epsilon}< \left(\displaystyle\frac{1}{2}-\frac{1}{2^{*}}\right)S^{\frac{N}{2}}\text{ for all }\lambda>\lambda_{0}, 0<\epsilon<\epsilon_{0}, \end{equation}

where

(40)\begin{equation} S=\inf_{u\in H_{0}^{1}(\Omega)\backslash\{0\}}\displaystyle\displaystyle\frac{\displaystyle\int_{\Omega}|\nabla u|^{2}\,{\rm d}x}{\bigg(\displaystyle\int_{\Omega}|u|^{2^{*}}\,{\rm d}x\bigg)^{\frac{2}{2^{*}}}}. \end{equation}

We now obtain the main result of this section.

Proposition 4.3 Let $a_{1}>0,$ $a_{2}>0$ and $\lambda _{0}>0$ be given by (29), Lemmas 4.1 and (39), respectively. If $\lambda >\lambda _{0},$ then problem (2) has a non-negative solution $u_{\epsilon }$ such that

\[ 0< a_{1}\leq c_{\lambda,\epsilon}=I_{\epsilon,\lambda}(u_{\epsilon})\leq a_{2}, \]

where $c_{\lambda,\epsilon }$ is given by (33). Furthermore, there exists a constant $D>0$ that does not depend on $\epsilon$ such that

\[ \|u_{\epsilon}\|_{H_{0}^{1}(\Omega)}< D\text{ for all }0<\epsilon<1. \]

Proof of Proposition 4.3 Inequality (35) implies that there is $u_{\epsilon }\in H_{0}^{1}(\Omega )$ with $\|u_{\epsilon }\|_{H^{1}_{0}(\Omega )}< D$ such that, up to a subsequence,

(41)\begin{align} u_n^{\epsilon}\rightharpoonup u_{\epsilon} \text{ weakly in } H^{1}_0(\Omega),\quad u_{n}^{\epsilon}\to u_{\epsilon} \text{ in } L^{r}(\Omega)\text{ for all }1\leq r< 2^{*},\quad u_{n}^{\epsilon}\to u_{\epsilon} \text{ a.e in }\Omega. \end{align}

We claim that

(42)\begin{equation} \displaystyle\int_{\Omega}((u_n^{\epsilon})^{+})^{2^{*}}\to\int_{\Omega}((u_{\epsilon})^{+})^{2^{*}}\text{ as }n\to\infty. \end{equation}

To do this, we use the ideas given in [Reference Figueiredo and Montenegro14]. Note that there exist positive measures $\mu,\,\nu$ in $\Omega$ such that

\[ |\nabla (u_n^{\epsilon})^{+}|^{2}\rightharpoonup|\nabla u_{\epsilon}^{+}|^{2}+\mu\text{ and }((u_n^{\epsilon})^{+})^{2^{*}}\rightharpoonup(u_{\epsilon}^{+})^{2^{*}}+\nu \]

Using the concentration-compactness principle due to Lions (cf. [Reference Lions18], Lemma 1.1), we obtain at most a countable set of indexes denoted by $\Lambda$, sequences $x_{i}\in \overline {\Omega }$, $\mu _{i},\,\nu _{i}\in (0,\,\infty )$ such that

\[ \nu=\sum_{i\in\Lambda}\nu_{i}\delta_{x_{i}},\quad\mu\geq\sum_{i\in\Lambda}\mu_{i}\delta_{x_{i}}\quad\text{ and } S\nu_{i}^{\frac{2}{2^{*}}}\leq \mu_{i}, \]

for every $i\in \Lambda$, where $S$ is given by (40). Now, for every $\sigma >0$ and $i\in \Lambda$, we define

\[ \psi_{\sigma,i}(x)=\psi\bigg(\frac{x-x_{i}}{\sigma}\bigg), \]

where $\psi \in C_{c}^{\infty }(\mathbb {R}^{n})$ is a function satisfying

\[ 0\leq\psi\leq 1,\quad\psi\equiv 1\text{ in }B_{1}(0),\quad\psi\equiv 0\text{ in }\mathbb{R}^{n}\setminus B_{2}(0)\quad\text{ and } \|\nabla\psi\|_{L^{\infty}(\mathbb{R}^{n})}\leq 2. \]

Since the function $\psi _{\sigma,i}(u_{n}^{\epsilon })^{+}$ is bounded in $H_{0}^{1}(\Omega )$, we know that $I_{\lambda,\epsilon }'(u_{n})(\psi _{\sigma,i}(u_{n}^{\epsilon })^{+})\to 0$ as $n\to \infty$. Hence,

\begin{align*} & \int_{\Omega}\nabla u_{n}^{\epsilon}\nabla(\psi_{\sigma,i}(u_{n}^{\epsilon})^{+})+\int_{\Omega}g_{\epsilon}(u_{n}^{\epsilon})\psi_{\sigma,i}((u_{n}^{\epsilon})^{+})\\& \quad =\lambda\int_{\Omega}((u_{n}^{\epsilon})^{+})^{2}\psi_{\sigma,i}+\int_{\Omega}((u_{n}^{\epsilon})^{+})^{2^{*}}\psi_{\sigma,i}+o_{n}(1). \end{align*}

Consequently,

(43)\begin{align} \displaystyle\int_{\Omega}|\nabla (u_{n}^{\epsilon})^{+}|^{2}\psi_{\sigma,i}+\int_{\Omega}(u_{n}^{\epsilon})^{+}\nabla u_{n}^{\epsilon}\nabla\psi_{\sigma,i}\leq\lambda\int_{\Omega}((u_{n}^{\epsilon})^{+})^{2}\psi_{\sigma,i}+\int_{\Omega}((u_{n}^{\epsilon})^{+})^{2^{*}}\psi_{\sigma,i}+o_{n}(1). \end{align}

Note that

\[ \lim_{n\to\infty}\int_{\Omega}((u_{n}^{\epsilon})^{+})^{2^{*}}\psi_{\sigma,i}=\int_{\Omega}(u_{\epsilon}^{+})^{2^{*}}\psi_{\sigma,i}+\int_{\Omega}\psi_{\sigma,i}\,d\nu. \]

Hence,

\[ \lim_{\sigma\to 0}\lim_{n\to\infty}\int_{\Omega}((u_{n}^{\epsilon})^{+})^{2^{*}}\psi_{\sigma,i}=\nu_{i}. \]

It is also clear that

\[ \lim_{\sigma\to 0}\lim_{n\to\infty}\int_{\Omega}((u_{n}^{\epsilon})^{+})^{2}\psi_{\sigma,i}=\lim_{\sigma\to 0}\int_{\Omega}(u_{\epsilon}^{+})^{2}\psi_{\sigma,i}=0 \]

and

\[ \lim_{\sigma\to 0}\lim_{n\to\infty}\int_{\Omega}|\nabla (u_{n}^{\epsilon})^{+}|^{2}\psi_{\sigma,i}\geq\mu_{i}. \]

We claim that

(44)\begin{equation} \lim_{\sigma\to 0}\lim_{n\to\infty}\displaystyle\int_{\Omega}(u_{n}^{\epsilon})^{+}\nabla u_{n}^{\epsilon}\nabla\psi_{\sigma,i}=0. \end{equation}

Indeed,

\[ |(u_{n}^{\epsilon})^{+}\nabla u_{n}^{\epsilon}\nabla\psi_{\sigma,i}|\leq\frac{((u_{n}^{\epsilon})^{+})^{2}|\nabla\psi_{\sigma,i}|^{2}}{2}+\frac{|\nabla (u_{n}^{\epsilon})^{+}|^{2}}{2} \]

Therefore,

\[ \lim_{n\to\infty}\int_{\Omega}(u_{n}^{\epsilon})^{+}\nabla u_{n}^{\epsilon}\nabla\psi_{\sigma,i}\leq J_{1,\sigma}+J_{2,\sigma} \]

where

\[ J_{1,\sigma}=\frac{1}{2}\int_{\Omega} (u_{\epsilon}^{+})^{2}|\nabla\psi_{\sigma,i}|^{2}\quad\text{ and }\quad J_{2,\sigma}=\frac{1}{2}\int_{\{\sigma<|x-x_{i}|<2\sigma\}} |\nabla u_{\epsilon}^{+}|^{2}+\int_{\{\sigma<|x-x_{i}|<2\sigma\}}\,d\mu. \]

Using the Lebesgue Differentiation Theorem and the bound on $\nabla \psi$, we obtain

\[ 2\lim_{\sigma\to 0}J_{1,\sigma}\leq\lim_{\sigma\to 0}\frac{4}{\sigma^{2}}\int_{\{\sigma<|x-x_{i}|<2\sigma\}}(u_{\epsilon}^{+})^{2}\leq C_{N}\lim_{\sigma\to 0}\sigma^{N-2}\frac{1}{V(B_{2\sigma}(x_{i}))}\int_{B_{2\sigma}(x_{i})}(u_{\epsilon}^{+})^{2}=0, \]

where $V(B_{2\sigma }(x_{i}))$ denotes the volume of the ball $B_{2\sigma }(x_{i})$. Hence, $\lim _{\sigma \to 0}J_{1,\sigma }=0$. It is also clear that $\lim _{\sigma \to 0}J_{2,\sigma }=0$. This proves (44). Letting $n\to \infty$ and $\sigma \to 0$ in (43), we get

\[ \mu_{i}\leq\nu_{i}\text{ for every }i\in\Lambda. \]

Hence,

\[ \nu_{i}^{\frac{2}{N}}\geq S\text{ for each }i\in\Lambda. \]

Since

\[ I_{\epsilon,\lambda}(u_{n}^{\epsilon})=\frac{1}{2}\int_{\Omega}|\nabla u_{n}^{\epsilon}|^{2}+\int_{\Omega}G_{\epsilon}(u_{n}^{\epsilon})-\frac{\lambda}{2}\int_{\Omega}((u_{n}^{\epsilon})^{+})^{2}-\frac{1}{2^{*}}\int_{\Omega}((u_{n}^{\epsilon})^{+})^{2^{*}}. \]

and

\[ I_{\epsilon,\lambda}'(u_{n}^{\epsilon})(u_{n}^{\epsilon})=\int_{\Omega}\nabla u_{n}^{\epsilon}\nabla u_{n}^{\epsilon}+\int_{\Omega}g_{\epsilon}(u_{n}^{\epsilon})u_{n}^{\epsilon}-\lambda\int_{\Omega}(u_{n}^{\epsilon})^{+}u_{n}^{\epsilon}-\int_{\Omega}((u_{n}^{\epsilon})^{+})^{2^{*}-1}u_{n}^{\epsilon}, \]

it follows that

\[ I_{\epsilon,\lambda}(u_{n}^{\epsilon})-\frac{1}{2}I_{\epsilon,\lambda}'(u_{n}^{\epsilon})(u_{n}^{\epsilon})=\int_{\Omega}\left(G_{\epsilon}(u_{n}^{\epsilon})-\frac{1}{2}g_{\epsilon}(u_{n}^{\epsilon})u_{n}^{\epsilon}\right)+\bigg(\frac{1}{2}-\frac{1}{2^{*}}\bigg)\int_{\Omega}((u_{n}^{\epsilon})^{+})^{2^{*}}. \]

From Lemmas 2.2, (34) and from the definition of $\psi _{\sigma,i}$, we obtain

(45)\begin{equation} c_{\lambda,\epsilon}+o_{n}(1)\geq \left(\displaystyle\frac{1}{2}-\frac{1}{2^{*}}\right)\displaystyle\int_{\Omega}((u_{n}^{\epsilon})^{+})^{2^{*}}\geq\bigg(\frac{1}{2}-\frac{1}{2^{*}}\bigg)\int_{B_{\sigma}(x_{i})}\psi_{\sigma,i}((u_{n}^{\epsilon})^{+})^{2^{*}}\text{ for each }i\in\Lambda. \end{equation}

Note that

\[ \lim_{n\to\infty}\int_{B_{\sigma}(x_{i})}\psi_{\sigma,i}((u_{n}^{\epsilon})^{+})^{2^{*}}=\int_{B_{\sigma}(x_{i})}\psi_{\sigma,i}(u_{\epsilon}^{+})^{2^{*}}+\int_{B_{\sigma}(x_{i})}\psi_{\sigma,i}\,d\nu\geq\nu_{i}\geq S^{\frac{N}{2}}. \]

Hence, letting $n\to \infty$ in (45), we obtain

\[ c_{\lambda,\epsilon}\geq \bigg(\frac{1}{2}-\frac{1}{2^{*}}\bigg)S^{\frac{N}{2}}. \]

This contradicts (39). This proves that $\Lambda =\emptyset$ and therefore (42) holds. We will now show that $u_{n}^{\epsilon }\to u_{\epsilon }$ in $H_{0}^{1}(\Omega )$. Indeed

\begin{align*} I_{\epsilon,\lambda}'(u_{n}^{\epsilon})(u_{n}^{\epsilon})-I_{\epsilon,\lambda}'(u_{n}^{\epsilon})(u^{\epsilon})& =\displaystyle\int_{\Omega}|\nabla u_{n}^{\epsilon}|^{2}-2\int_{\Omega}\nabla u_{n}^{\epsilon}\nabla u_{\epsilon}+\int_{\Omega}|\nabla u_{\epsilon}|^{2}\\ & \quad-L_{1,n}^{\epsilon}+L_{2,n}^{\epsilon}-L_{3,n}^{\epsilon}-L_{4,n}^{\epsilon}, \end{align*}

where

\begin{align*} L_{1,n}^{\epsilon}& =\int_{\Omega}\left(|\nabla u_{\epsilon}|^{2}-\nabla u_{n}^{\epsilon}\nabla u_{\epsilon}\right),\\ L_{2,n}^{\epsilon}& =\int_{\Omega}\left(g_{\epsilon}(u_{n}^{\epsilon})(u_{n}^{\epsilon}-u_{\epsilon})\right),\\ L_{3,n}^{\epsilon}& =\lambda\int_{\Omega}\left((u_{n}^{\epsilon})^{+}(u_{n}^{\epsilon}-u_{\epsilon})\right) \end{align*}

and

\[ L_{4,n}^{\epsilon}=\int_{\Omega}\left(((u_{n}^{\epsilon})^{+})^{2^{*}-1}(u_{n}^{\epsilon}-u_{\epsilon})\right)=\int_{\Omega}((u_{n}^{\epsilon})^{+})^{2^{*}}-\int_{\Omega}((u_{n}^{\epsilon})^{+})^{2^{*}-1}u_{\epsilon} \]

Using the Dominated Convergence Theorem, (41) and (42), we obtain that

\[ \lim_{n\to\infty}\max\{L_{1,n}^{\epsilon},L_{2,n}^{\epsilon},L_{3,n}^{\epsilon}, L_{4,n}^{\epsilon}\}=0. \]

Therefore, it follows from (34) that

\[ \|u_{n}^{\epsilon}-u_{\epsilon}\|^{2}_{H_{0}^{1}(\Omega)}=o_{n}(1). \]

Therefore, $u_{n}^{\epsilon }\to u_{\epsilon }$ strongly in $H_{0}^{1}(\Omega )$. From (34), it follows that $u_{\epsilon }$ is a critical point of $I_{\epsilon,\lambda }$ with

\[ 0< a_{1}< I_{\epsilon,\lambda}(u_{\epsilon})< a_{2}. \]

In particular, we know that $u_{\epsilon }\geq 0$. This proves Proposition 4.3.

5. Regularity results and gradient estimates

We will need the following a priori bound in $L^{\infty }(\Omega )$.

Lemma 5.1 Let $u_{\epsilon,\lambda }\in H_{0}^{1}(\Omega )$ be a non-negative solution of problem (2) with $f(s)=\lambda s+s^{p}$ and assume that there exists a constant $D>0$ independent of $\epsilon$ such that

(46)\begin{equation} \|u_{\epsilon,\lambda}\|_{H_{0}^{1}(\Omega)}\leq D\text{ for each }0<\epsilon<1. \end{equation}

Then the following assertions hold

${\rm (i)}$ If $1< p<2^{*}-1$ then $u_{\epsilon,\lambda }\in L^{\infty }(\Omega )$ and there exists a constant $K_{1}>0$ such that

(47)\begin{equation} \|u_{\epsilon,\lambda}\|_{L^{\infty}(\Omega)}\leq K_{1}\text{ for each }0<\epsilon<1. \end{equation}

${\rm (ii)}$ If $p=2^{*}-1$ and

(48)\begin{equation} \lim_{\lambda\to\infty}I_{\epsilon,\lambda}(u_{\epsilon,\lambda})=0\text{ uniformly on }\epsilon, \end{equation}

then there exists $\widehat {\lambda _{0}}>0$ such that $u_{\epsilon,\lambda }\in L^{\infty }(\Omega )$ for each $\lambda >\widehat {\lambda _{0}}$ and (47) holds.

Proof of Lemma 5.1 For simplicity, we denote $u_{\epsilon,\lambda }$ by $u_{\epsilon }$. For $s\geq 0$, define $h(s)=\lambda s+s^{p}$. From (4), we get

(49)\begin{equation} \displaystyle\int_{\Omega}\nabla u_{\epsilon}\nabla v+\int_{\Omega}g_{\epsilon}(u_{\epsilon})v=\int_{\Omega}h(u_{\epsilon})v\text{ for all }v\in H_{0}^{1}(\Omega). \end{equation}

Note that

\[ \frac{h(s)}{g_{\epsilon}(s)}=\frac{(s+\epsilon)^{q+\beta}}{s^{q}}(\lambda s+s^{p})\leq (s+1)^{q+\beta}(\lambda s^{1-q}+s^{p-q})\to 0\text{ as }s\to 0. \]

Hence, the exists $0<\delta _{\lambda }<1$ that does not depend on $\epsilon$ such that

(50)\begin{equation} \displaystyle\frac{h(s)}{g_{\epsilon}(s)}<\frac{1}{2}\text{ for }s\leq\delta_{\lambda}. \end{equation}

Also,

\[ \frac{h(s)}{s^{p}}=\lambda s^{1-p}+1\text{ for }s\geq\delta_{\lambda}. \]

Therefore, we conclude that

(51)\begin{equation} \displaystyle\frac{h(s)}{s^{p}}\leq 2\text{ for }s\geq A_{\lambda}, \end{equation}

where

\[ A_{\lambda}=\lambda^{\frac{1}{p-1}}. \]

It is also clear that

(52)\begin{equation} h(s)=\lambda s+s^{p}\leq\lambda A_{\lambda}+A_{\lambda}^{p}=C_{\lambda}\text{ for }\delta_{\lambda}\leq s\leq A_{\lambda}. \end{equation}

Using (50), (51) and (52), we obtain

\[ h(s)<\left(\frac{g_{\epsilon}(s)}{2}\right)\chi_{\{0\leq s\leq\delta_{\lambda}\}}+C_{\lambda}\chi_{\{\delta_{\lambda}\leq s\leq A_{\lambda}\}}+2s^{p}\chi_{\{s\geq A_{\lambda}\}}\text{ for }s\geq 0 \]

Hence, from (70), we get

(53)\begin{equation} \displaystyle\int_{\Omega}\nabla u_{\epsilon}\nabla v< C_{\lambda}\int_{\{\delta_{\lambda}\leq u_{\epsilon}\leq A_{\lambda}\}}v+2\int_{\Omega}u_{\epsilon}^{p}v\text{ for all }v\in H_{0}^{1}(\Omega),v\geq 0. \end{equation}

We will now prove assertions $(i)$ and $(ii)$ separately. The proof of $(ii)$ is more intricate, because we need to study the dependence of certain constants on $\lambda$, so that we can let $\lambda \to \infty$.

Proof of (i): From (53), we obtain a constant $C_{\delta,\lambda }>0$ such that

(54)\begin{equation} \displaystyle\int_{\Omega}\nabla u_{\epsilon}\nabla v< C_{\delta,\lambda}\int_{\Omega}u_{\epsilon}^{p}v\text{ for all }v\in H_{0}^{1}(\Omega),v\geq 0. \end{equation}

For $L > 1$, we define,

\begin{align*} u_{L,\epsilon}(x)& =\left\{ \begin{array}{@{}l} u_{\epsilon}(x), \text{if } u_{\epsilon}(x) \leq L \\ L, \text{if } u_{\epsilon}(x) \geq L, \\ \end{array} \right.\\ z_{L,\epsilon} & = u_{L,\epsilon}^{2(\sigma - 1)}u_{\epsilon} \quad \mbox{and} \quad w_{L,\epsilon} = u_{\epsilon} u_{L,\epsilon}^{\sigma - 1}, \end{align*}

with $\sigma > 1$ to be determined later. Note that $z_{L,\epsilon }\in H_{0}^{1}(\Omega )$, $z_{L,\epsilon }\geq 0$ and

\[ \nabla z_{L,\epsilon}=u_{L,\epsilon}^{2(\sigma-1)}\nabla u_{\epsilon}+2(\sigma-1)u_{\epsilon}u_{L,\epsilon}^{2\sigma-3}\nabla u_{L,\epsilon}. \]

Taking $v=z_{L,\epsilon }$ in (54) we obtain

\[ \int_{\Omega}u_{L,\epsilon}^{2(\sigma-1)}|\nabla u_{\epsilon}|^{2}+2(\sigma-1)\int_{\Omega}u_{\epsilon}u_{L,\epsilon}^{2\sigma-3}\nabla u_{\epsilon}\nabla u_{L,\epsilon}< C_{\lambda,\delta}\int_{\Omega}u_{\epsilon}^{p+1}u_{L,\epsilon}^{2(\sigma-1)}. \]

Since $\sigma >1$ and

\[ \int_{\Omega}u_{\epsilon}u_{L,\epsilon}^{2\sigma-3}\nabla u_{\epsilon}\nabla u_{L,\epsilon}=\int_{\{u_{\epsilon}< L\}}u_{\epsilon}^{2(\sigma-1)}|\nabla u_{\epsilon}|^{2}\geq 0, \]

we conclude that

(55)\begin{equation} \displaystyle\int_{\Omega}u_{L,\epsilon}^{2(\sigma-1)}|\nabla u_{\epsilon}|^{2}< C_{\lambda,\delta}\int_{\Omega}u_{\epsilon}^{p+1}u_{L,\epsilon}^{2(\sigma-1)}< C_{\lambda,\delta}\int_{\Omega}u_{\epsilon}^{p-1}u_{\epsilon}^{2\sigma}. \end{equation}

On the other hand, from the Sobolev embedding, we know that there is a constant $C_{1}>0$ such that

\[ \bigg(\int_{\Omega}w_{L,\epsilon}^{p+1}\,{\rm d}x\bigg)^{\frac{2}{p+1}}\leq C_{1} \int_{\Omega}|\nabla w_{L,\epsilon}|^{2}\,{\rm d}x. \]

Since

\[ \nabla w_{L,\epsilon}=u_{L,\epsilon}^{\sigma-1}\nabla u_{\epsilon}+(\sigma-1)u_{\epsilon}u_{L,\epsilon}^{\sigma-2}\nabla u_{L,\epsilon}, \]

it follows that

\begin{align*} & \left(\displaystyle\int_{\Omega}w_{L,\epsilon}^{p+1}\,{\rm d}x\right)^{\frac{2}{p+1}}\leq C_{1}\int_{\Omega}u_{L,\epsilon}^{2(\sigma-1)}|\nabla u_{\epsilon}|^{2}\,{\rm d}x+ C_{1}(\sigma-1)^{2}\int_{\Omega}u_{\epsilon}^{2}u_{L,\epsilon}^{2(\sigma-2)}|\nabla u_{L,\epsilon}|^{2}\\ & \quad+2C_{1}(\sigma-1)\displaystyle\int_{\Omega}u_{\epsilon}u_{L,\epsilon}^{2\sigma-3}\nabla u_{\epsilon}\nabla u_{L,\epsilon}. \end{align*}

From the definition of $u_{L,\epsilon }$, we conclude that

\[ \bigg(\int_{\Omega}w_{L,\epsilon}^{p+1}\,{\rm d}x\bigg)^{\frac{2}{p+1}}\leq C_{1}\sigma^{2}\int_{\Omega}u_{L,\epsilon}^{2(\sigma-1)}|\nabla u_{\epsilon}|^{2}\,{\rm d}x. \]

Using (55), we obtain

(56)\begin{equation} \left(\displaystyle\int_{\Omega}w_{L,\epsilon}^{p+1}\,{\rm d}x\right)^{\frac{2}{p+1}}\leq C_{1}\sigma^{2} C_{\delta,\lambda}\int_{\Omega}u_{\epsilon}^{p-1}u_{\epsilon}^{2\sigma}. \end{equation}

Now, observe that

\[ \left(\int_{\Omega}w_{L,\epsilon}^{p+1}\,{\rm d}x\right)^{\frac{2}{p+1}}=\left(\int_{\Omega}u_{\epsilon}^{p+1}u_{L,\epsilon}^{(p+1)(\sigma-1)}\,{\rm d}x\right)^{\frac{2}{p+1}}\geq\left(\int_{\Omega}u_{L,\epsilon}^{\sigma(p+1)}\,{\rm d}x\right)^{\frac{2}{p+1}}. \]

Hence, there is a constant $\widetilde {C_{\delta,\lambda }}>0$ such that

(57)\begin{equation} \left(\displaystyle\int_{\Omega}u_{L,\epsilon}^{\sigma(p+1)}\,{\rm d}x\right)^{\frac{2}{p+1}}\leq \sigma^{2}\widetilde{C_{\delta,\lambda}}\int_{\Omega}u_{\epsilon}^{p-1}u_{\epsilon}^{2\sigma}. \end{equation}

Let $\alpha _{1},\,\alpha _{2}>1$ be constants such that $\frac {1}{\alpha _{1}}+\frac {1}{\alpha _{2}}=1$ and $p+1<\alpha _{1}(p-1)<2^{*}$. From (57) and Hölder's inequality it follows that

\[ \left(\int_{\Omega}u_{L,\epsilon}^{\sigma(p+1)}\,{\rm d}x\right)^{\frac{2}{p+1}}\leq \sigma^{2} \widetilde{C_{\delta,\lambda}}\left(\int_{\Omega}u_{\epsilon}^{\alpha_{1}(p-1)}\,{\rm d}x\right)^{\frac{1}{\alpha_{1}}}\left(\int_{\Omega}u_{\epsilon}^{2\sigma\alpha_{2}}\,{\rm d}x\right)^{\frac{1}{\alpha_{2}}}. \]

Using (46) and the Sobolev Embedding, we obtain a constant $\widetilde {C}>0$ such that

\[ \int_{\Omega}u_{\epsilon}^{\alpha_{1}(p-1)}\,{\rm d}x\leq \widetilde{C}. \]

Hence, there exists a constant $\widehat {C}>0$ that does not depend on $\sigma$ nor on $\epsilon$ such that

(58)\begin{equation} \left(\displaystyle\int_{\Omega}u_{L,\epsilon}^{\sigma(p+1)}\,{\rm d}x\right)^{\frac{2}{p+1}}\leq \widehat{C}\sigma^{2}\left(\int_{\Omega}u_{\epsilon}^{2\sigma\alpha_{2}}\,{\rm d}x\right)^{\frac{1}{\alpha_{2}}}. \end{equation}

Letting $L\to \infty$ in (58) and using Fatou's Lemma, we conclude that

\[ \left(\int_{\Omega}u_{\epsilon}^{\sigma(p+1)}\,{\rm d}x\right)^{\frac{2}{p+1}}\leq \widehat{C}\sigma^{2}\left(\int_{\Omega}u_{\epsilon}^{2\sigma\alpha_{2}}\,{\rm d}x\right)^{\frac{1}{\alpha_{2}}}\text{ for each }\sigma>1, \]

provided $u_{\epsilon }\in L^{2\sigma \alpha _{2}}(\Omega )$. Equivalently,

(59)\begin{equation} \|u_{\epsilon}\|_{L^{\sigma(p+1)}}\leq C^{\frac{1}{\sigma}}\sigma^{\frac{1}{\sigma}}\|u_{\epsilon}\|_{L^{2\sigma\alpha_{2}}}\text{ for each }\sigma>1, \end{equation}

where $C=\sqrt {\widehat {C}}$. Observe that the choices of $\alpha _{1}$ and $\alpha _{2}$ imply that $\sigma (p+1)>2\sigma \alpha _{2}$. The result now follows from an iterative argument. Indeed, take

\[ \sigma_{1}=\frac{p+1}{2\alpha_{2}}. \]

Using the Sobolev embedding and (46), we obtain a constant $\widetilde {D}>0$ such that

\[ \|u_{\epsilon}\|_{L^{\sigma_{1}(p+1)}(\Omega)}\leq C^{\frac{1}{\sigma_{1}}}\sigma_{1}^{\frac{1}{\sigma_{1}}}\|u_{\epsilon}\|_{L^{p+1}(\Omega)}\leq \widetilde{D}C^{\frac{1}{\sigma_{1}}}\sigma_{1}^{\frac{1}{\sigma_{1}}} \]

Now, take $\sigma _{2}=\sigma _{1}^{2}$ in (59). We get

\[ \|u_{\epsilon}\|_{L^{\sigma_{1}^{2}(p+1)}(\Omega)}\leq C^{\frac{1}{\sigma_{1}^{2}}}\sigma_{2}^{\frac{1}{\sigma_{2}}}\|u_{\epsilon}\|_{L^{\sigma_{1}(p+1)}}\leq\widetilde{D} C^{\frac{1}{\sigma_{1}}+\frac{1}{\sigma_{1}^{2}}}\left(\sigma_{1}^{\frac{1}{\sigma_{1}}}\sigma_{2}^{\frac{1}{\sigma_{2}}}\right) \]

Taking $\sigma _{k}=\sigma _{1}^{k}$ in (59), we get

(60)\begin{equation} \|u_{\epsilon}\|_{L^{\sigma_{1}^{k}(p+1)}(\Omega)}\leq \widetilde{D} C^{\sum_{i=1}^{k}\frac{1}{\sigma_{1}^{i}}}(\Pi_{i=1}^{k}\sigma_{i}^{\frac{1}{\sigma_{i}}}). \end{equation}

It is clear that

\[ \lim_{k\rightarrow\infty}\left(\Pi_{i=1}^{k}\sigma_{i}^{\frac{1}{\sigma_{i}}}\right)=\lim_{k\rightarrow\infty}\left(\Pi_{i=1}^{k}\sigma_{1}^{\frac{i}{\sigma_{1}^{i}}}\right)<\infty\text{ and } \lim_{k\rightarrow\infty}C^{\sum_{i=1}^{k}\frac{1}{\sigma_{1}^{i}}}<\infty. \]

Letting $k\rightarrow \infty$ in (60), it follows that $u_{\epsilon }\in L^{\infty }(\Omega )$ and we obtain a constant $K_{1}>0$ that does not depend on $\epsilon$ such that

\[ \|u_{\epsilon}\|_{L^{\infty}(\Omega)}\leq K_{1}. \]

This proves (47).

Proof of (ii). Suppose that $p=2^{*}-1$. This case is much more complicated. We have

\[ \int_{\Omega}\nabla u_{\epsilon}\nabla v\,{\rm d}x+\int_{\Omega}g_{\epsilon}(u_{\epsilon})v\,{\rm d}x=\int_{\Omega}(\lambda u_{\epsilon}+u_{\epsilon}^{2^{*}-1})v\,{\rm d}x\text{ for all }v\in H_{0}^{1}(\Omega). \]

Consequently, since $g_{\epsilon }\geq 0$, we know that

(61)\begin{equation} \displaystyle\int_{\Omega}\nabla u_{\epsilon}\nabla v\leq\int_{\Omega}u_{\epsilon}(\lambda+u_{\epsilon}^{2^{*}-2})v\text{ for all } v\in H_{0}^{1}(\Omega),v\geq 0, 0<\epsilon<1. \end{equation}

For each $0<\epsilon <1$, we define

\[ a_{\epsilon}(x)=\lambda+u_{\epsilon}(x)^{2^{*}-2}. \]

Observe that

\[ \int_{\Omega}a_{\epsilon}(x)^{N/2}\leq C(N)\left(\lambda^{N/2}|\Omega|+\int_{\Omega}\left(u_{\epsilon}^{\frac{4}{N-2}}\right)^{N/2}\right)=C(N)(\lambda^{N/2}|\Omega|+\|u_{\epsilon}\|_{L^{2^{*}}(\Omega)}^{2^{*}}), \]

where $C(N)$ is a constant that depends only on $N$. From (46) and the Sobolev embedding, we get a constant $C>0$ such that

\[ \|a_{\epsilon}\|_{L^{N/2}(\Omega)}\leq C\text{ for all }0<\epsilon<1. \]

Let $\sigma \geq 0$ be a constant to be fixed later and consider the function $z_{L,\epsilon }=u_{\epsilon }\min \{u_{\epsilon }^{2\sigma },\, L^{2}\}\in H_{0}^{1}(\Omega )$, with $L>0$. Observe that

\[ \nabla u_{\epsilon}\nabla z_{L,\epsilon}=|\nabla u_{\epsilon}|^{2}\min\{u_{\epsilon}^{2\sigma},L^{2}\}+2\sigma u_{\epsilon}^{2\sigma}|\nabla u_{\epsilon}|^{2}\chi_{\{u_{\epsilon}^{\sigma}\leq L\}}. \]

Taking $v=z_{L,\epsilon }$ in (61), we get

(62)\begin{equation} \displaystyle\int_{\Omega}|\nabla u_{\epsilon}|^{2}\min\{u_{\epsilon}^{2\sigma},L^{2}\}+2\sigma\int_{\{u_{\epsilon}^{s}\leq L\}}u_{\epsilon}^{2\sigma}|\nabla u_{\epsilon}|^{2}\leq\int_{\Omega}a_{\epsilon}(x)u_{\epsilon}^{2}\min\{u_{\epsilon}^{2\sigma},L^{2}\}. \end{equation}

Define $w_{L,\epsilon }$ by $w_{L,\epsilon }=u_{\epsilon }\min \{u_{\epsilon }^{\sigma },\,L\}\in H_{0}^{1}(\Omega )$. We have

\[ \nabla w_{L,\epsilon}=\min\{u_{\epsilon}^{\sigma},L\}\nabla u_{\epsilon}+\sigma u_{\epsilon}^{\sigma}\nabla u_{\epsilon}\chi_{\{u_{\epsilon}^{\sigma}\leq L\}}. \]

We thus get

\[ |\nabla w_{L,\epsilon}|^{2}= \sigma^{2} u_{\epsilon}^{2\sigma}|\nabla u_{\epsilon}|^{2}\chi_{\{u_{\epsilon}^{\sigma}\leq L\}}+\min\{u_{\epsilon}^{2\sigma},L^{2}\}|\nabla u_{\epsilon}|^{2}+\sigma u_{\epsilon}^{2\sigma}|\nabla u_{\epsilon}|^{2}\chi_{\{u_{\epsilon}^{\sigma}\leq L\}}. \]

Hence,

\[ |\nabla w_{L,\epsilon}|^{2}=(1+\sigma(\sigma+1))\min\{u_{\epsilon}^{2\sigma},L^{2}\}|\nabla u_{\epsilon}|^{2}\text{ in }\{u_{\epsilon}^{\sigma}\leq L\} \]

and

\[ |\nabla w_{L,\epsilon}|^{2}=\min\{u_{\epsilon}^{2\sigma},L^{2}\}|\nabla u_{\epsilon}|^{2}\text{ in }\{u_{\epsilon}^{\sigma}> L\}. \]

We conclude that

\[ |\nabla w_{L,\epsilon}|^{2}\leq c(\sigma)\min\{u_{\epsilon}^{2\sigma},L^{2}\}|\nabla u_{\epsilon}|^{2}\text{ in }\Omega, \]

where $c(\sigma )=1+\sigma (\sigma +1)$. From (62), we get

\[ \int_{\Omega}|\nabla w_{L,\epsilon}|^{2}\leq c(\sigma)\int_{\Omega}\min\{u_{\epsilon}^{2\sigma},L^{2}\}|\nabla u_{\epsilon}|^{2}\leq c(\sigma)\int_{\Omega}a_{\epsilon}(x)u_{\epsilon}^{2}\min\{u_{\epsilon}^{2\sigma},L^{2}\}. \]

Now, fix $K>0$. We have

\[ \int_{\Omega}|\nabla w_{L,\epsilon}|^{2}\leq c(\sigma)K\int_{\Omega}u_{\epsilon}^{2}\min\{u_{\epsilon}^{2\sigma},L^{2}\}+c(\sigma)\int_{\{a_{\epsilon}\geq K\}}a_{\epsilon}(x)u_{\epsilon}^{2}\min\{u_{\epsilon}^{2\sigma},L^{2}\}. \]

Hence,

\begin{align*} \displaystyle\int_{\Omega}|\nabla w_{L,\epsilon}|^{2}& \leq c(\sigma)K\int_{\Omega}u_{\epsilon}^{2}\min\{u_{\epsilon}^{2\sigma},L^{2}\}\\ & \quad+c(\sigma)\left(\displaystyle\int_{\{a_{\epsilon}\geq K\}}a_{\epsilon}(x)^{\frac{N}{2}}\right)^{\frac{2}{N}}\left(\int_{\Omega}(u_{\epsilon}\min\{u_{\epsilon}^{\sigma},L\})^{\frac{2N}{N-2}}\right)^{\frac{N-2}{N}}. \end{align*}

Consequently,

\[ \int_{\Omega}|\nabla w_{L,\epsilon}|^{2}\leq c(\sigma)K\int_{\Omega}u_{\epsilon}^{2}\min\{u_{\epsilon}^{2\sigma},L^{2}\}+c(\sigma)\left(\int_{\{a_{\epsilon}\geq K\}}a_{\epsilon}(x)^{\frac{N}{2}}\right)^{\frac{2}{N}}\left(\int_{\Omega}w_{L,\epsilon}^{\frac{2N}{N-2}}\right)^{\frac{N-2}{N}}. \]

From the Sobolev embedding Theorem, we get

(63)\begin{align} \displaystyle\int_{\Omega}|\nabla w_{L,\epsilon}|^{2}\leq c(\sigma)K\int_{\Omega}u_{\epsilon}^{2}\min\{u_{\epsilon}^{2\sigma},L^{2}\}+Cc(\sigma)\left(\int_{\{a_{\epsilon}\geq K\}}a_{\epsilon}(x)^{\frac{N}{2}}\right)^{\frac{2}{N}}\int_{\Omega}|\nabla w_{L,\epsilon}|^{2}. \end{align}

Choose $K>0$ such that

\[ \left(\int_{\{a_{\epsilon}\geq K\}}a_{\epsilon}(x)^{\frac{N}{2}}\right)^{\frac{2}{N}}\leq \frac{1}{2C c(\sigma)}. \]

Claim 1: $K$ can be chosen independently of $\epsilon$, provided $\lambda$ is sufficiently large.

Assuming the claim to be true, we obtain

\[ \int_{\Omega}|\nabla w_{\epsilon}|^{2}\,{\rm d}x\leq Cc(\sigma)K\int_{\Omega}u_{\epsilon}^{2}\min\{u_{\epsilon}^{2\sigma},L^{2}\}. \]

Consequently,

\[ \int_{\{u_{\epsilon}^{\sigma}\leq L\}}|\nabla (u_{\epsilon}^{\sigma+1})|^{2}\,{\rm d}x\leq Cc(\sigma)K\int_{\Omega}(u_{\epsilon}\min\{u_{\epsilon}^{\sigma},L\})^{2}\,{\rm d}x. \]

Suppose that $u_{\epsilon }\in L^{2\sigma +2}(\Omega )$ and let $u_{L,\epsilon }=(u_{\epsilon }\min \{u_{\epsilon }^{\sigma },\,L\})^{2}$. Observe that

\[ \lim_{L\to\infty}u_{L,\epsilon}= u_{\epsilon}^{2+2\sigma}\text{ almost everywhere in }\Omega \]

Furthermore,

\[ L_{1}< L_{2}\text{ implies that }u_{L_{1},\epsilon}< u_{L_{2},\epsilon}\text{ in }\Omega. \]

The Monotone Convergence Theorem implies that

\[ \int_{\{u_{\epsilon}^{\sigma}\leq L\}}|\nabla (u_{\epsilon}^{\sigma+1})|^{2}\,{\rm d}x\leq Cc(\sigma)K\int_{\Omega}u_{\epsilon}^{2+2\sigma}\,{\rm d}x. \]

From Fatou's Lemma, we get

(64)\begin{equation} \displaystyle\int_{\Omega}|\nabla (u_{\epsilon}^{\sigma+1})|^{2}\,{\rm d}x\leq Cc(\sigma)K\int_{\Omega}u_{\epsilon}^{2+2\sigma}\,{\rm d}x\text{ for all }\sigma\geq 0. \end{equation}

Consequently, $u_{\epsilon }^{\sigma +1}\in H_{0}^{1}(\Omega )$ and $u_{\epsilon }\in L^{\frac {(2+2\sigma )N}{N-2}}(\Omega )$. Now let $q>1$. We will show that

(65)\begin{equation} u_{\epsilon}\in L^{q}(\Omega)\text{ and }\|u_{\epsilon}\|_{L^{q}(\Omega)}< C_{q}\text{ for all }0<\epsilon<1. \end{equation}

Indeed, this follows by choosing adequate values for $\sigma$ in (64). Let $\sigma _{0}=0$. From (64), we get

\[ \int_{\Omega}|\nabla u_{\epsilon}|^{2}\,{\rm d}x\leq CK\int_{\Omega}u_{\epsilon}^{2}\,{\rm d}x\leq C_{0}. \]

Let $\sigma _{1}=\frac {N}{N-2}-1$. From (64), we get

\[ \int_{\Omega}|\nabla (u_{\epsilon}^{\frac{N}{N-2}})|^{2}\,{\rm d}x\leq CK\int_{\Omega}u_{\epsilon}^{\frac{2N}{N-2}}\,{\rm d}x\leq C_{1}. \]

Consequently, $u_{\epsilon }^{\frac {N}{N-2}}\in H_{0}^{1}(\Omega )$ and

\[ \int_{\Omega}u_{\epsilon}^{(\frac{N}{N-2})\frac{2N}{N-2}}\leq C_{1}. \]

Let $\sigma _{2}=\frac {4N-4}{(N-2)^{2}}$. From (64), we get

\[ \int_{\Omega}|\nabla (u_{\epsilon}^{\frac{N^{2}}{(N-2)^{2}}})|^{2}\,{\rm d}x\leq CK\int_{\Omega}u_{\epsilon}^{\frac{2N^{2}}{(N-2)^{2}}}\,{\rm d}x\leq C_{2}. \]

Consequently, $u_{\epsilon }^{\frac {N^{2}}{(N-2)^{2}}}\in H_{0}^{1}(\Omega )$ and

\[ \int_{\Omega}u_{\epsilon}^{(\frac{N^{2}}{(N-2)^{2}})\frac{2N}{N-2}}\leq C_{2}. \]

Assertion (65) then follows by choosing

\[ \sigma_{i}+1=(\sigma_{i-1}+1)\frac{N}{N-2}. \]

and iterating up until $s_{M}>q$ for some $M\in \mathbb {N}$. Now let $w_{\epsilon }$ be the solution of the non-singular problem

(66)\begin{equation} \left\{ \begin{array}{@{}l} -\Delta w=\lambda u_{\epsilon}+u_{\epsilon}^{2^{*}-1} \text{ in } \Omega\\ w = 0 \text{ on }\partial\Omega, \end{array} \right. \end{equation}

Assertion (65) and elliptic regularity theory implies that $w_{\epsilon }\in W^{2,q}(\Omega )$ and

\[ \|w_{\epsilon}\|_{ W^{2,q}(\Omega)}\leq C|\lambda u_{\epsilon}+u_{\epsilon}^{2^{*}-1}|_{L^{q}(\Omega)}=C_{q}, \]

where $C_{q}$ does not depend on $\epsilon$. Consequently, the Sobolev embedding assures that $w_{\epsilon }\in C^{1}(\overline {\Omega })$ and

\[ \|w_{\epsilon}\|_{C^{1}(\Omega)}\leq C\text{ for all }0<\epsilon<1. \]

Observe that

\[ \int_{\Omega}\nabla w_{\epsilon}\nabla v=\int_{\Omega}(\lambda u_{\epsilon}+u_{\epsilon}^{2^{*}-1})v\text{ for all }v\in H_{0}^{1}(\Omega), 0<\epsilon<1. \]

Consequently,

\[ \int_{\Omega}\nabla (u_{\epsilon}-w_{\epsilon})\nabla v\leq 0\text{ for all }v\in H_{0}^{1}(\Omega), v\geq 0, 0<\epsilon<1. \]

The weak maximum principle implies

\[ \sup_{\Omega} (u_{\epsilon}-w_{\epsilon})=0. \]

Hence,

\[ u_{\epsilon}\leq w_{\epsilon}. \]

Consequently,

\[ \|u_{\epsilon}\|_{L^{\infty}(\Omega)}< C\text{ for all }0<\epsilon<1. \]

This proves the result. We need to only show that the claim holds. Indeed, from (46), there exists an element $u\in H_{0}^{1}(\Omega )$ such that, up to a subsequence,

(67)\begin{equation} \left\{ \begin{aligned} & u_{\epsilon}\rightharpoonup u \text{ weakly in } H^{1}_0(\Omega),\\ & u_{\epsilon}\to u \text{ in } L^{r}(\Omega) \text{ for } 1< r<2^{*},\\ & u_{\epsilon}\to u \text{ a.e in } \Omega. \end{aligned} \right. \end{equation}

We first show that

(68)\begin{equation} \displaystyle\int_{\Omega}u_{\epsilon}^{2^{*}}\to\int_{\Omega}u^{2^{*}}\text{ as }\epsilon\to 0. \end{equation}

Again using the Concentration-compactness principle of Lions, we get positive measures $\mu,\,\nu$ in $\Omega$ such that

\[ |\nabla u_{\epsilon}|^{2}\rightharpoonup|\nabla u|^{2}+\mu\text{ and }u_{\epsilon}^{2^{*}}\rightharpoonup u^{2^{*}}+\nu. \]

Furthermore, there is at most a countable set of indexes denoted by $\Lambda$, sequences $x_{i}\in \overline {\Omega }$, $\mu _{i},\,\nu _{i}\in (0,\,\infty )$ such that

\[ \nu=\sum_{i\in\Lambda}\nu_{i}\delta_{x_{i}},\quad\mu\geq\sum_{i\in\Lambda}\mu_{i}\delta_{x_{i}}\quad\text{ and } S\nu_{i}^{\frac{2}{2^{*}}}\leq \mu_{i}, \]

for every $i\in \Lambda$, where $S$ is given by (40). Now, for every $\sigma >0$ and $i\in \Lambda$, we define

\[ \psi_{\sigma,i}(x)=\psi\left(\frac{x-x_{i}}{\sigma}\right), \]

where $\psi \in C_{c}^{\infty }(\mathbb {R}^{n})$ is a function satisfying

\[ 0\leq\psi\leq 1,\quad\psi\equiv 1\text{ in }B_{1}(0),\quad\psi\equiv 0\text{ in }\mathbb{R}^{n}\setminus B_{2}(0)\quad\text{ and } \|\nabla\psi\|_{L^{\infty}(\mathbb{R}^{n})}\leq 2. \]

Proceeding as in the proof of Proposition 4.3 and using the hypothesis $\lim _{\lambda \to \infty }I_{\epsilon }(u_{\epsilon,\lambda })=0$, we conclude that $\Lambda =\emptyset$, provided $\lambda >0$ is sufficiently large, thus proving (68). Consequently, $a_{\epsilon }^{\frac {N}{2}}$ converges in $L^{1}(\Omega )$ to $a$, where

\[ a(x)=(\lambda+u(x)^{2^{*}-2})^{\frac{N}{2}}. \]

We now show that for each $\delta ^{*}>0$ there exists $\eta >0$ and $\epsilon _{0}>0$ such that

(69)\begin{equation} \displaystyle\int_{B}a_{\epsilon}(x)^{\frac{N}{2}}<\delta^{*}\text{ for all sets } B\subset\Omega\text{ with }|B|<\eta\text{ and }0<\epsilon<\epsilon_{0}. \end{equation}

Since $a\in L^{1}(\Omega )$, there exists $\eta >0$ such that

\[ \int_{B}a(x)<\delta^{*}/2\text{ for all sets } B\subset\Omega\text{ with }|B|<\eta. \]

We write

\[ \int_{B}a_{\epsilon}(x)^{\frac{N}{2}}=\int_{B}(a_{\epsilon}(x)^{\frac{N}{2}}-a(x))+\int_{B}a(x). \]

and we choose $\epsilon _{0}>0$ such that

\[ \int_{\Omega}|a_{\epsilon}(x)^{\frac{N}{2}}-a(x)|\leq\delta^{*}/2\text{ for all }0<\epsilon<\epsilon_{0}. \]

Consequently,

\[ \int_{B}a_{\epsilon}(x)^{\frac{N}{2}}\leq\delta^{*}\text{ for all sets } B\subset\Omega\text{ with }|B|<\eta\text{ and }0<\epsilon<\epsilon_{0}. \]

This proves (69). We now finally prove Claim 1. Indeed, we choose

\[ \delta^{*}=\left(\frac{1}{2 Cc(\sigma)}\right)^{\frac{N}{2}} \]

and we choose $K>0$ such that

\[ \left|\left\{u> \frac{(K-\lambda)^{\frac{1}{2^{*}-2}}}{2}\right\}\right|<\eta. \]

Observe that

\[ \{a_{\epsilon}\geq K\}=\{u_{\epsilon}(x)\geq (K-\lambda)^{\frac{1}{2^{*}-2}}\}. \]

The choice of $K$ implies that

\[ |\{a_{\epsilon}\geq K\}|\leq \left|\left\{u> \frac{(K-\lambda)^{\frac{1}{2^{*}-2}}}{2}\right\}\right|<\eta\text{ for sufficiently small }\epsilon. \]

Consequently, from (69) and the choice of $\delta ^{*}$, we get

\[ \left(\int_{\{a_{\epsilon}\geq K\}}a_{\epsilon}(x)^{\frac{N}{2}}\right)^{\frac{2}{N}}\leq \frac{1}{2C c(\sigma)}\text{ for sufficiently small }\epsilon. \]

This proves Claim 1 and the result.

We also have

Lemma 5.2 Let $u_{\epsilon }\in H_{0}^{1}(\Omega )$ be a non-negative solution of problem (2) and assume that there exists a constant $D>0$ independent of $\epsilon$ such that

\[ \|u_{\epsilon}\|_{H_{0}^{1}(\Omega)}\leq D\text{ for each }0<\epsilon<1. \]

If $f$ satisfies (8) and (9) for $0< p<\frac {N+2}{N-2}$, then $u_{\epsilon }\in L^{\infty }(\Omega )$ and there exists a constant $K_{2}>0$ such that

\[ \|u_{\epsilon}\|_{L^{\infty}(\Omega)}\leq K_{2}\text{ for each }0<\epsilon<1. \]

Proof of Lemma 5.2 From (4), we get

(70)\begin{equation} \displaystyle\int_{\Omega}\nabla u_{\epsilon}\nabla v+\int_{\Omega}g_{\epsilon}(u_{\epsilon})v-\int_{\Omega}f(u_{\epsilon})v=0\text{ for all }v\in H_{0}^{1}(\Omega),v\geq 0 \end{equation}

From (8), we get

\[ \int_{\Omega}\nabla u_{\epsilon}\nabla v+\int_{\Omega\cap\{u_{\epsilon}\geq\delta\}}g_{\epsilon}(u_{\epsilon})v-\int_{\Omega\cap\{u_{\epsilon}\geq\delta\}}f(u_{\epsilon})v\leq 0\text{ for all }v\in H_{0}^{1}(\Omega),v\geq 0. \]

Consequently,

\[ \int_{\Omega}\nabla u_{\epsilon}\nabla v\leq\int_{\Omega\cap\{u_{\epsilon}\geq\delta\}}f(u_{\epsilon})v\text{ for all }v\in H_{0}^{1}(\Omega),v\geq 0. \]

From (9), we get

\[ \int_{\Omega}\nabla u_{\epsilon}\nabla v\leq\int_{\Omega\cap\{u_{\epsilon}\geq\delta\}}C(1+u_{\epsilon}^{p})v\text{ for all }v\in H_{0}^{1}(\Omega),v\geq 0. \]

Consequently,

\[ \int_{\Omega}\nabla u_{\epsilon}\nabla v\leq C\int_{\Omega\cap\{u_{\epsilon}\geq\delta\}}\left(\frac{1}{\delta^{p}}+1\right)u_{\epsilon}^{p}v\text{ for all }v\in H_{0}^{1}(\Omega),v\geq 0. \]

Consequently, there exists $\widetilde {C}>0$ and $1<\widetilde {p}<2^{*}-1$ such that

\[ \int_{\Omega}\nabla u_{\epsilon}\nabla v\leq \widetilde{C}\int_{\Omega}u_{\epsilon}^{\widetilde{p}}v\text{ for all }v\in H_{0}^{1}(\Omega),v\geq 0. \]

The proof then follows as in item $(i)$ of Lemma 5.1.

Now, we obtain gradient estimates for solutions $u_{\epsilon }$ of problem (2).

Lemma 5.3 Assume that $f$ satisfies (6). For each $0<\epsilon <1$, let $u_\epsilon \in H_{0}^{1}(\Omega )\cap L^{\infty }(\Omega )$ be a non-negative solution of problem (2) and assume that there exists a constant $T>0$ such that

(71)\begin{equation} \sup_{0<\epsilon<1}\|u_{\epsilon}\|_{L^{\infty}(\Omega)}< T<\infty. \end{equation}

Let $\psi$ be such that

\begin{align*} \psi \in C^{2}(\overline{\Omega}), \ \psi > 0 \text{ in } \Omega,\ \psi = 0 \text{ on } \partial \Omega \text{ and } \displaystyle\frac{|\nabla \psi|^{2}}\psi \text{ is bounded in } \Omega. \end{align*}

Then, there exist constants $M>0$ and $\epsilon _{0}>0$ such that

\[ {\psi(x) |\nabla u_\epsilon (x) |^{2}}\leq M (u_{\epsilon}(x)^{1-\beta}+u_{\epsilon}(x)) \text{ for every } x\in \Omega,\quad 0<\epsilon<\epsilon_{0}. \]

Proof of Lemma 5.3 From (6), we obtain constants $C_{1}>0$ and $0< t_{0}<1$ such that

(72)\begin{equation} |\widetilde{f}'(s)|\leq C_{1} s^{q_{1}-1}\text{ for }0\leq s\leq t_{0}. \end{equation}

From (71) we obtain that $\Delta u_\epsilon$ is bounded in $L^{\infty }(\Omega )$. Thus, by standard elliptic regularity, $u_\epsilon$ belongs to $C^{1,\nu }(\overline \Omega )$. We define

\[ \overline{h}_\epsilon(u) = g_{\epsilon}(u) - f(u). \]

We shall denote $u_\epsilon$ simply by $u$. Define the functions

\[ Z(u)=u^{1-\beta}+u+a,\qquad w = \frac{|\nabla u|^{2}}{Z(u)}, \qquad v = w \psi, \]

where $a>0$ is small. Note that $v$ is $C^{2}$ at all points $x\in \Omega$ such that $u(x)>0$. Indeed, let $x\in \Omega$ be one such point. By continuity, there must exist an open ball $B\subset \Omega$ centred at $x$ such that $u>0$ in $\overline {B}$. Consequently, we know that $g_{\epsilon }(u)\in C^{1,\nu }(B)$ and $f(u)\in C^{1,\nu }(B)$. Hence, $\overline {h}_{\epsilon }(u)\in C^{1,\nu }(B)$. Since $u$ satisfies the equation $-\Delta u+\overline {h}_{\epsilon }(u)=0$ in $B$, we conclude that $u\in C^{3}(B)$, implying that $Z(u)$ and $w$ are $C^{2}$ in $B$.

The function $v$ is continuous in $\overline \Omega$, hence it attains its maximum at some point $x_0 \in \overline \Omega$. Thus, we obtain

\[ v(x_0) > 0 . \]

Note that $x_0 \in \Omega$, because $v=0$ on $\partial \Omega$. Furthermore, $u(x_{0})>0$, since otherwise $x_{0}$ would be a critical point of $u$ and $w(x_{0})=0$. Hence,

\[ \nabla v(x_0) = 0 \]

and

(73)\begin{equation} \Delta v (x_0) \leq 0. \end{equation}

The computations already carried out in [Reference Lorca and Montenegro20, Reference Montenegro and Silva22] lead to the following expression evaluated at $x_0$

(74)\begin{equation} \begin{aligned} \Delta v & \geq \displaystyle\frac{1}{Z(u)}\left[\psi w^{2}\left(\frac{1}{2} Z'(u)^{2} - Z(u) Z''(u)\right)\right.\\ & \quad+ w ( 2 \psi Z(u)\overline{h}_\epsilon'(u) - \psi \overline{h}_\epsilon (u) Z'(u) - K_0 Z(u) )\\ & \quad\left.- K_{0} Z'(u) Z(u)^{1/2} \psi^{1/2} w^{3/2}\vphantom{\left(\frac{1}{2} Z'(u)^{2} - Z(u) Z''(u)\right)}\right], \end{aligned} \end{equation}

where

\[ K_0 = \max \left( \sup_\Omega \left({\frac{|\nabla \psi |}{\psi^{1/2}}}\right), \sup_\Omega \left(\Delta \psi - 2 \frac{|\nabla \psi|^{2}}\psi\right) \right)>0. \]

We will show that if $v(x_0)$ is large enough then the right-hand side of (74) must be positive, which would contradict (73).

We will establish the following estimates uniformly for every $\epsilon$ sufficiently small.

(75)\begin{align} Z'(u) Z(u)^{1/2} & \leq C \left({ \displaystyle\frac{1}{2}} Z'(u)^{2} - Z''(u) Z(u)\right), \end{align}
(76)\begin{align} Z(u) |\overline{h}_\epsilon'(u)| & \leq C \left( { \frac{1}{2}} Z'(u)^{2} - Z''(u) Z(u)\right), \end{align}
(77)\begin{align} Z'(u) |\overline{h}_\epsilon(u)| & \leq C \left( { \frac{1}{2}} Z'(u)^{2} - Z''(u) Z(u)\right), \end{align}
(78)\begin{align} Z(u) & \leq C \left({ \frac{1}{2}} Z'(u)^{2} - Z''(u) Z(u)\right) , \end{align}

for every $0 \leq u \leq T$. The constant $C$ depends only on $T$, but not on $\epsilon$ nor on $a$.

Assuming for a moment that (75)(78) are true. Inequality (74) implies that

\begin{align*} \Delta v & \geq \displaystyle\frac{\frac 12 Z'(u)^{2} - Z''(u) Z(u)}{Z(u)} ( \psi w^{2} - C (w + \psi^{1/2} w^{3/2} ) ) \\ & = \displaystyle\frac{\frac 12 Z'(u)^{2} - Z''(u) Z(u)}{Z(u) \psi} ( v^{2} - C (v + v^{3/2} ) ). \end{align*}

Since $\Delta v(x_{0})\leq 0$ and

\[ \frac{\frac 12 Z'(u)^{2} - Z''(u) Z(u)}{Z(u)\psi}>0\text{ in }\Omega, \]

we conclude that

\[ v(x_{0})^{2}-C(v(x_{0})+v(x_{0})^{3/2})\leq 0. \]

Consequently, there exists $M>0$ that does not depend on $a$ such that

\[ \sup_{\Omega}v=v(x_{0})< M. \]

Consequently,

\[ |\nabla u(x)|^{2}\psi(x)\leq M(u(x)^{1-\beta}+u(x)+a)\text{ for all }x\in\Omega. \]

The result then follows by letting $a\to 0$.

We prove now the relations (75)(78). In the course of this proof, $C$, $\widetilde {C}$, $C_{i}$, $i\in \{1,\,2,\,3,\,\ldots \}$ denote various positive constants independent of $\epsilon$ and $a$, we obtain gradient

\[ Z(u)=u^{1-\beta}+u+a, \]
\[ Z'(u)=(1-\beta)u^{-\beta}+1,\hspace{0.5cm} Z''(u)={-}\beta(1-\beta)u^{-\beta-1}. \]

Hence,

(79)\begin{equation} \displaystyle\frac{1}{2} Z'(u)^{2} - Z''(u) Z(u) \geq \frac{(1-\beta)^{2}}{2}(u^{{-}2\beta}+1)+a\beta(1-\beta)u^{{-}1-\beta}. \text{ for } u >0 . \end{equation}

We first prove (78). Indeed, there is a constant $C>0$ such that

\[ Z(u)=u^{1-\beta}+u+a\leq C\text{ for } 0\leq u\leq T. \]

Hence, (78) follows from (79).

We now prove (77). Note that there exists a constant $\widetilde {C}>0$ such that

\begin{align*} Z'(u) |\overline{h}_\epsilon(u)|& \leq((1-\beta)u^{-\beta}+1)(g_{\epsilon}(u)+|f(u)|)\\ & \leq (1-\beta)u^{{-}2\beta}+(1-\beta)u^{-\beta}\sup_{0\leq s\leq T}|f(s)|+u^{-\beta}+\sup_{0\leq s\leq T}|f(s)|\\ & \leq \widetilde{C}(1+u^{{-}2\beta}). \end{align*}

Inequality (77) then follows from (79).

Now, we prove (76). Note that

\[ \overline{h}_{\epsilon}'(u)=\frac{u^{q-1}}{(u+\epsilon)^{q+\beta+1}}\left(q\epsilon-\beta u\right)-f'(u). \]

We split the proof of (76) in three cases.

Case I. Suppose that $0< u<\min \{\tfrac {q\epsilon }{2\beta },\,t_{0}\}$, where $0< t_{0}<1$ is given by (72). We define

\[ \omega_{\epsilon}(u)=\frac{u^{q-1}}{(u+\epsilon)^{q+\beta+1}}\left(q\epsilon-\beta u\right)-C_{1} u^{q_{1}-1}, \]

where $C_{1}>0$ is given by (72). We claim that there exists $\epsilon _{0}>0$ such that $\omega _{\epsilon }(u)>0$ for each $0<\epsilon <\epsilon _{0}$. Indeed, assume by contradiction that $\omega _{\epsilon }(u)<0$ for some $0< u<\frac {q\epsilon }{2\beta }$. We then have

\[ q\epsilon u^{q-1}<\beta u^{q}+C_{1} u^{q_{1}-1}(u+\epsilon)^{q+\beta+1}<\beta u^{q}+C_{1} u^{q_{1}-1}\epsilon^{q+\beta+1}\left(1+\frac{q}{2\beta}\right)^{q+\beta+1}. \]

Now take $\epsilon _{0}>0$ such that

\[ C_{1}\epsilon^{q+\beta+1}\left(1+\frac{q}{2\beta}\right)^{q+\beta+1}<\frac{\epsilon q}{2}\text{ for }0<\epsilon<\epsilon_{0}. \]

We may assume that $0< q< q_{1}$. Consequently,

\[ q\epsilon u^{q-1}<\beta u^{q}+\frac{\epsilon qu^{q_{1}-1}}{2}<\beta u^{q}+\frac{\epsilon qu^{q-1}}{2}. \]

Hence,

\[ \frac{q\epsilon u^{q-1}}{2}<\beta u^{q}, \]

which implies that

\[ u>\frac{q\epsilon}{2\beta}. \]

This contradicts our initial assumption. The claim is proven. Since

\[ \frac{q\epsilon u^{q-1}}{(u+\epsilon)^{q+\beta+1}}\leq q\frac{u^{q}}{u(u+\epsilon)^{q}}\frac{\epsilon}{(u+\epsilon)^{\beta+1}}\leq\frac{q}{u^{\beta+1}}, \]

we obtain

\[ |\overline{h}_{\epsilon}'(u)|=\overline{h}_{\epsilon}'(u)\leq\frac{q+C_{1}u^{q_{1}+\beta}}{u^{\beta+1}}\text{ for }0< u<\min\left\{\frac{q\epsilon}{2\beta},t_{0}\right\}. \]

Hence,

\[ |\overline{h}_{\epsilon}'(u)|\leq\frac{2q}{u^{\beta+1}}\text{ for }0< u<\min\left\{\frac{q\epsilon}{2\beta},t_{0},t_{1}\right\}, \]

where $t_{1}>0$ is chosen such that

\[ C_{1}u^{q_{1}+\beta}< q\text{ for }0\leq u\leq t_{1}. \]

Therefore,

\[ Z(u)|\overline{h}_{\epsilon}'(u)|\leq (u^{1-\beta}+u+a)\left(\displaystyle\frac{2q}{u^{\beta+1}}\right)\leq \frac{2q}{u^{2\beta}}+\frac{2qa}{u^{\beta+1}}\text{ for }0\leq u\leq \min\left\{\frac{q\epsilon}{2\beta},t_{0},t_{1}\right\}. \]

Comparing with (79), it follows that there exists a constant $C>0$ that does not depend on $a$ such that

(80)\begin{align} Z(u)|\overline{h}_{\epsilon}'(u)|\leq C(\displaystyle\frac{1}{2} Z'(u)^{2} - Z''(u) Z(u))\text{ for }0< u<\min\left\{\frac{q\epsilon}{2\beta},t_{0},t_{1}\right\},\quad 0<\epsilon<\epsilon_{0}. \end{align}

Case II. Suppose that $\frac {\epsilon q}{2\beta }\leq u\leq \min \{t_{0},\,t_{1}\}$. We have

\[ |\overline{h}_{\epsilon}'(u)|\leq\frac{u^{q-1}|q\epsilon-\beta u|}{(u+\epsilon)^{q+\beta+1}}+C_{1} u^{q_{1}-1}\text{ for }\frac{q\epsilon}{2\beta}\leq u\leq t_{0}. \]

Note that $|q\epsilon -\beta u|\leq \beta u$ if $2\beta u\geq q\epsilon$. We then obtain

\[ |\overline{h}_{\epsilon}'(u)|\leq\frac{\beta u^{q}+C_{1} u^{q+q_{1}+\beta}\left(1+\frac{2\beta}{q}\right)^{q+\beta+1}}{(u+\epsilon)^{q+\beta+1}}\text{ for }\frac{q\epsilon}{2\beta}\leq u\leq t_{0}. \]

Now, observe that there exists $0< t_{2}<\min \{t_{0},\,t_{1}\}$ that does not depend on $\epsilon$ such that

\[ C_{1} u^{q_{1}+\beta}\left(1+\frac{2\beta}{q}\right)^{q+\beta+1}<\beta\text{ for }0\leq u\leq t_{2}. \]

Therefore,

\[ |\overline{h}_{\epsilon}'(u)|\leq\frac{2\beta}{u^{\beta+1}}\text{ for }\frac{q\epsilon}{2\beta}\leq u< t_{2}. \]

Comparing with (79), we obtain

(81)\begin{equation} Z(u)|\overline{h}_{\epsilon}'(u)|\leq C\left(\displaystyle\frac{1}{2} Z'(u)^{2} - Z''(u) Z(u)\right)\text{ for }\frac{q\epsilon}{2\beta}\leq u< t_{2}. \end{equation}

Case III. Assume that $t_{2}\leq u\leq T$. Since there exists a constant $C>0$ such that $|\overline {h}_{\epsilon }'(u)|\leq C$ for $t_{2}\leq u\leq T$, it follows from (78) that

(82)\begin{equation} Z(u)|\overline{h}_{\epsilon}'(u)|\leq C\left(\displaystyle\frac{1}{2} Z'(u)^{2} - Z''(u) Z(u)\right)\text{ for }t_{2}\leq u\leq T. \end{equation}

Hence, (76) follows from (80), (81) and (82).

We now prove (75). Observe that

\[ Z'(u) Z(u)^{1/2}=((1-\beta)u^{-\beta}+1)\sqrt{u^{1-\beta}+u+a}. \]

Hence,

\[ Z'(u) Z(u)^{1/2}\leq \sqrt{3T}((1-\beta) u^{-\beta}+1). \]

When $0\leq u\leq 1$ we know that $u^{2}\leq u$. Hence $u^{-\beta }\leq u^{-2\beta }$. Therefore, from (79), there exist constants $C_{3}>0$ and $C_{4}>C_{3}$ such that

(83)\begin{equation} Z'(u) Z(u)^{1/2}\leq C_{3} (u^{{-}2\beta}+1)\leq C_{4}\left(\displaystyle\frac{1}{2} Z'(u)^{2} - Z''(u) Z(u)\right)\text{ for }0\leq u\leq 1. \end{equation}

If $1\leq u\leq T$, we know that there exists a constant $C_{5}>0$ such that $Z'(u) Z(u)^{1/2}\leq C_{5}$. Hence, from (79), there exists a constant $C_{6}>0$ such that

(84)\begin{equation} Z'(u) Z(u)^{1/2}\leq C_{6}\left(\displaystyle\frac{1}{2} Z'(u)^{2} - Z''(u) Z(u)\right)\text{ for }1\leq u\leq T. \end{equation}

Inequality (75) then follows from (83) and (84). We have proved Lemma 5.3.

Consequently, we obtain

Corollary 5.4 For each $0<\epsilon <1,$ let $u_\epsilon$ be the solution of problem (2) obtained in Propositions 3.2 and 4.3. Let $\psi$ be as in the hypothesis of Lemma 5.3. Then there exist constants $M>0$ and $\epsilon _{0}>0$ such that

\[ {\psi(x) |\nabla u_\epsilon (x) |^{2}}\leq M (u_{\epsilon}(x)^{1-\beta}+u_{\epsilon}(x)) \text{ for every } x\in \Omega,\quad 0<\epsilon<\epsilon_{0}. \]

Proof of Corollary 5.4 From Propositions 3.2 and 4.3, we know that there is a constant $D>0$ such that

\[ \|u_{\epsilon}\|_{H_{0}^{1}(\Omega)}< D\text{ for each } 0<\epsilon<1. \]

From Lemmas 4.2, 5.1 and 5.2 we conclude that the solutions $u_{\epsilon }$ of (2) are bounded in $L^{\infty }(\Omega )$ by constant $K_{1}>0$ and $K_{2}>0$ independent of $\epsilon$. Corollary 5.4 then follows by Lemma 5.3.

6. The limit of approximate solutions

Now we will study the convergence as $\epsilon \to 0$ of the solutions $u_{\epsilon }$ of problem (2) obtained in Propositions 3.2 and 4.3. First, we obtain the existence of a non-trivial limit $u$. Next, we prove that $u$ is a solution of problem (1).

Lemma 6.1 Let $(\epsilon _{n})$ be a sequence in $(0,\,1)$ such that $\epsilon _{n}\to 0$ as $n\to \infty$. Let $(u_{\epsilon _{n}}^{1})$ and $(u_{\epsilon _{n}}^{2})$ be the sequences of solutions obtained in Propositions 3.2 and 4.3 respectively. Then there exist non-trivial functions $u_{1}\in H_{0}^{1}(\Omega )$ and $u_{2}\in H_{0}^{1}(\Omega )$ such that, up to a subsequence, $u_{\epsilon _{n}}^{i}\rightharpoonup u_{i}$ weakly in $H_{0}^{1}(\Omega ),$ where $i\in \{1,\,2\}$.

Proof of Lemma 6.1 From Propositions 3.2 and 4.3, we know that there exist constants $D_{i}>0$ such that

\[ \|u_{\epsilon_{n}}^{i}\|_{H_{0}^{1}(\Omega)}< D_{i}\text{ for each }n\in\mathbb{N},i\in\{1,2\} \]

Hence, there exist functions $u_{i}\in H_{0}^{1}(\Omega )$ such that

(85)\begin{equation} \left\{ \begin{aligned} & u_{\epsilon_{n}}^{i}\rightharpoonup u_{i} \text{ weakly in } H^{1}_0(\Omega);\\ & u_{\epsilon_{n}}^{i}\to u_{i} \text{ in } L^{r}(\Omega) \text{ for every } 1\leq r<2^{*};\\ & u_{\epsilon_{n}}^{i}\to u_{i} \text{ a.e in } \Omega; \end{aligned} \right. \end{equation}

Lemmas 5.1 and 5.2 imply that $u_{\epsilon _{n}}^{i}\in L^{\infty }(\Omega )$ with $\|u_{\epsilon _{n}}^{i}\|_{L^{\infty }(\Omega )}< K_{i}$ for all $n\in \mathbb {N}$. Consequently, the Dominated Convergence Theorem implies that

\[ u_{\epsilon_{n}}^{i}\to u_{i} \text{ in } L^{r}(\Omega) \text{ for every } r\geq 1. \]

We prove the result for $i=1$ and denote $(u_{\epsilon _{n}}^{1})$ and $u_{1}$ merely by $(u_{\epsilon _{n}})$ and $u$ respectively. From Proposition 3.2, we have

\[ 0< a_{1}\leq I_{\epsilon_{n}}(u_{\epsilon_{n}})=\frac{1}{2}\int_{\Omega}|\nabla u_{\epsilon_{n}}|^{2}+\int_{\Omega}G_{\epsilon_{n}}(u_{\epsilon_{n}})-\int_{\Omega}F(u_{\epsilon_{n}}). \]

Since $u_{\epsilon _{n}}$ is a non-negative critical point of $I_{\epsilon _{n}}$, we have

\[ \|u_{\epsilon_{n}}\|_{H_{0}^{1}(\Omega)}^{2}+\int_{\Omega}g_{\epsilon_{n}}(u_{\epsilon_{n}})u_{\epsilon_{n}}=\int_{\Omega}u_{\epsilon_{n}}f(u_{\epsilon_{n}}). \]

Hence,

(86)\begin{align} I_{\epsilon_{n}}(u_{\epsilon_{n}}) = \displaystyle\int_{\Omega}\left(G_{\epsilon_{n}}(u_{\epsilon_{n}})-\displaystyle\frac{1}{2}g_{\epsilon_{n}}(u_{\epsilon_{n}})u_{\epsilon_{n}}\right)\,{\rm d}x-\int_{\Omega}\left(F(u_{\epsilon_{n}})-\frac{1}{2}u_{\epsilon_{n}}f(u_{\epsilon_{n}})\right)\,{\rm d}x>a_{1}.\end{align}

The Dominated Convergence Theorem implies that

\begin{align*} & \lim_{n\rightarrow\infty}\int_{\Omega}g_{\epsilon_{n}}(u_{\epsilon_{n}})u_{\epsilon_{n}}\,{\rm d}x=\int_{\Omega}u^{1-\beta}\,{\rm d}x,\\ & \lim_{n\rightarrow\infty}\int_{\Omega}G_{\epsilon_{n}}(u_{\epsilon_{n}})\,{\rm d}x=\frac{1}{1-\beta}\int_{\Omega}u^{1-\beta}\,{\rm d}x,\\ \end{align*}

and

\[ \int_{\Omega}\left(F(u_{\epsilon_{n}})-\frac{1}{2}u_{\epsilon_{n}}f(u_{\epsilon_{n}})\right)\,{\rm d}x\to\int_{\Omega}\left(F(u)-\frac{1}{2}uf(u)\right)\,{\rm d}x. \]

Taking the above claims into account and letting $n\rightarrow \infty$ in (86), we obtain

\[ \int_{\Omega}\left(\frac{u^{1-\beta}}{1-\beta}-\frac{u^{1-\beta}}{2}\right)\,{\rm d}x -\int_{\Omega}\left(F(u)-\frac{1}{2}uf(u)\right)\,{\rm d}x\geq a_{1}. \]

We proved that $u$ is non-trivial. The proof for $i=2$ is analogous.

We now show that the functions $u_{1}$ and $u_{2}$ defined in Lemma 6.1 satisfy the following property.

Lemma 6.2 Let $u_{1}$ and $u_{2}$ be the functions given by Lemma 6.1. The function $u_{i}^{-\beta }\chi _{\{u_{i}>0\}}$ belongs to $L^{1}_{loc}(\Omega )$ for $i\in \{1,\,2\}$.

Proof of Lemma 6.2 We again prove the result for $i=1$. The proof for $i=2$ is analogous. Let $(u_{\epsilon _{n}})$ and $u$ be given by (85) with $i=1$. Let $V \subset \Omega$ be a open set such that $\overline {V}\subset \Omega$. Take $\zeta \in C^{1}_c(\Omega )$ such that $0 \leq \zeta \leq 1$ and $\zeta \equiv 1$ in $V$. Since $u_{\epsilon _{n}}$ is a critical point of $I_{\epsilon _{n}},$ we obtain

\[ \int_{\{u_{\epsilon_{n}}< 1-\epsilon_{n} \}} g_{\epsilon_{n}} (u_{\epsilon_{n}} ) \zeta = \int_\Omega f(u_{\epsilon_{n}}) \zeta - \int_\Omega \nabla u_{\epsilon_{n}}\nabla \zeta - \int_{\{u_{\epsilon_{n}}\geq 1-\epsilon_{n} \}} g_{\epsilon_{n}} (u_{\epsilon_{n}} ) \zeta. \]

Corollary 5.4 implies that $u_{\epsilon _{n}}\to u$ uniformly in compact subsets of $\Omega$. Since $u_{\epsilon _{n}}\rightharpoonup u$ weakly in $H_{0}^{1}(\Omega )$, we get

(87)\begin{equation} \displaystyle\int_{\{u_{\epsilon_{n}}< 1-\epsilon_{n} \}} g_{\epsilon_{n}} (u_{\epsilon_{n}} ) \zeta \rightarrow \int_\Omega f(u) \zeta - \int_\Omega \nabla u \nabla \zeta - \int_{\{u \geq 1\}} u^{-\beta} \zeta \text{ as } \epsilon\to 0 \end{equation}

Define the set $\Omega _\rho = \{ x \in \Omega : u(x) \geq \rho \}$ for $\rho >0$. It follows from (87) that there exists a constant $C>0$ that does not depend on $n$ nor on $\rho$ such that

\[ \int_{V \cap \Omega_\rho} \frac{u_{\epsilon_{n}}^{q}}{(u_{\epsilon_{n}}+\epsilon_{n})^{q+\beta}} \chi_{\{u_{\epsilon_{n}} <1-\epsilon_{n} \}}\zeta \leq \int_{\{u_{\epsilon_{n}}< 1-\epsilon_{n} \}} g_{\epsilon_{n}}(u_{\epsilon_{n}}) \zeta < C\text{ for all }n\in\mathbb{N},\quad \rho>0, \]

Letting $n\to \infty$ and using Fatou's Lemma, we then get

\[ \int_V u^{-\beta} \chi_{\Omega_\rho} < C. \]

Letting $\rho \rightarrow 0$ and applying Fatou's Lemma again, we conclude that

\[ \int_V u^{-\beta} \chi_{\{u>0\}} < \infty. \]

Since $V$ was arbitrarily chosen, Lemma 6.2 is proved.

Proof of Theorem 1.1 The proof of this result is very similar to the one given in [Reference Figueiredo and Montenegro14], but for the sake of completeness, we give the proof with details. We will show that the sequences $(u_{\epsilon _{n}}^{1})$ given by Lemma 6.1 converge to a solution $u_{1}$ of (1) as $n\to \infty$. In doing so, we obtain a solution $u_{1}$ of (1) which is non-trivial. The non-triviality of $u_{1}$ is guaranteed by Lemma 6.1. From now on, we denote $u_{\epsilon _{n}}$ and $u_{1}$ merely by $u_{\epsilon }$ and $u$ respectively. Let $\varphi \in C_c^{1}(\Omega )$. From Proposition 3.2, we have

(88)\begin{equation} \displaystyle\int_\Omega \nabla u_\epsilon\nabla \varphi = \int_\Omega ({-}g_\epsilon (u_\epsilon ) + f(u_{\epsilon})) \varphi. \end{equation}

Let $\eta \in C^{\infty }(\mathbb {R})$, $0 \leq \eta \leq 1$, $\eta (s)=0$ for $s \leq 1/2$, $\eta (s)=1$ for $s \geq 1$. For $m >0$ the function $\varrho := \varphi \eta (u_\epsilon /m)$ belongs to $C_c^{1}(\Omega )$.

From Corollary 5.4, we know that $|\nabla u_\epsilon |$ is locally bounded independent on $0<\epsilon <\epsilon _{0}$. It then follows from (47) and the Arzelà-Ascoli Theorem that $u_\epsilon \to u$ in $C^{0}_{loc}(\Omega )$, and the set $\Omega _+ = \{ x \in \Omega : u(x) > 0 \}$ is open. Let $\tilde \Omega$ be an open set such that $\overline {support ( \varphi ) } \subset \tilde \Omega$ and $\overline {\tilde \Omega } \subset \Omega$. Let $\Omega _0 = \Omega _+ \cap \tilde \Omega$. For every $m >0$, there is an $\epsilon _1 > 0$ such that

(89)\begin{equation} u_\epsilon (x) \leq m/2 \text{ for every } x \in \tilde \Omega \setminus \Omega_0\text{ and } 0 < \epsilon\leq \epsilon _1. \end{equation}

Replacing $\varphi$ by $\varrho$ in (88), we obtain

(90)\begin{equation} \displaystyle\int_\Omega \nabla u_\epsilon\nabla ( \varphi \eta(u_\epsilon /m) ) = \int_{\tilde \Omega} ({-}g_\epsilon (u_\epsilon ) + f(u_{\epsilon})) \varphi \eta(u_\epsilon /m). \end{equation}

We break the previous integral as

\[ X_\epsilon:= \int_{\Omega_0} ({-}g_\epsilon (u_\epsilon ) +f(u_{\epsilon})) \varphi \eta(u_\epsilon /m) \]

and

\[ Y_\epsilon:= \int_{\tilde{\Omega} \setminus \Omega_0} ({-}g_\epsilon (u_\epsilon ) + f(u_{\epsilon})) \varphi \eta(u_\epsilon /m). \]

Clearly, $Y_\epsilon =0$, whenever $0<\epsilon \leq \epsilon _1$ by (89) and the definition of $\eta$. From (47), the Dominated Convergence Theorem and from the fact that $u_{\epsilon }\to u$ uniformly in $\Omega _0$, we get

\[ X_\epsilon\to \int_{\Omega_0} ({-}u^{-\beta} + f(u)) \varphi \eta(u/m) \text{ as } \epsilon\to 0. \]

We take the limit in $m$ to conclude that

(91)\begin{equation} \displaystyle\int_{\Omega_0} ({-}u^{-\beta} + f(u)) \varphi \eta(u/m) \rightarrow \int_{\Omega_0} ({-}u^{-\beta} + f(u)) \varphi \text{ as } m \rightarrow 0, \end{equation}

since $\eta (u/m) \leq 1$ and $u^{-\beta }\chi _{\Omega ^{+}} + f(u) \in L^{1}(\tilde \Omega )$, according to Lemma 6.2.

What follows next is identical to [Reference Lorca and Montenegro20]. We proceed with the integral on the left side of (90). We have

(92)\begin{equation} \displaystyle\int_\Omega \nabla u_\epsilon\nabla ( \varphi \eta(u_\epsilon /m) ) = \int_{\tilde \Omega} (\nabla u_\epsilon\nabla \varphi) \eta (u_\epsilon /m) + W_\epsilon, \end{equation}

where

\[ W_\epsilon= \int_{\tilde \Omega} \frac{|\nabla u_\epsilon |^{2}}{m} \eta' (u_\epsilon /m) \varphi. \]

Consequently,

\[ \int_{\tilde \Omega} (\nabla u_\epsilon\nabla \varphi) \eta (u_\epsilon /m) \rightarrow \int_{\tilde \Omega} (\nabla u \nabla \varphi) \eta (u/m) \text{ as } \epsilon\to 0, \]

since $u_\epsilon \rightharpoonup u$ weakly in $H_0^{1}(\Omega )$ and $u_\epsilon \to u$ uniformly in $\tilde \Omega$. Hence, by the Dominated Convergence Theorem,

(93)\begin{equation} \displaystyle\int_{\tilde \Omega} (\nabla u \nabla \varphi) \eta (u/m) \rightarrow \int_{\tilde \Omega} \nabla u \nabla \varphi \text{ as } m \rightarrow 0. \end{equation}

Now we only need to show that

(94)\begin{equation} W_\epsilon\rightarrow 0\text{ as } \epsilon\to 0 \quad (\text{ and then as } m \rightarrow 0). \end{equation}

Let $Z_{0}(u_{\epsilon })=u_{\epsilon }^{1-\beta }+u_{\epsilon }$. The estimate $|\nabla u_\epsilon |^{2} \leq M Z_{0}(u_\epsilon )$ in $\tilde \Omega$ provided by Corollary 5.4 yields

\begin{align*} \limsup_{\epsilon \rightarrow 0}|W_\epsilon | & \leq \frac{M}{m} \lim_{\epsilon \rightarrow 0} \int_{\tilde \Omega \cap \{\frac{m}{2}\leq u_\epsilon\leq m \}} Z_{0}(u_\epsilon ) | \eta' (u_\epsilon /m) \varphi |\\ & \leq M \lim_{\epsilon \rightarrow 0} \int_{\tilde \Omega \cap \{\frac{m}{2} \leq u_\epsilon\leq m \}} \frac{Z_{0}(u_\epsilon ) | \eta' (u_\epsilon /m) \varphi|}{u_{\epsilon}} . \end{align*}

Consequently

\begin{align*} \limsup_{\epsilon \rightarrow 0}|W_\epsilon |& \leq M \sup|\eta'|\sup|\varphi| \lim_{\epsilon \rightarrow 0} \int_{\tilde \Omega \cap \{\frac{m}{2} \leq u_\epsilon\leq m \}} \frac{Z_{0}(u_\epsilon )}{u_\epsilon }\\& \leq M\sup|\eta'|\sup|\varphi| \int_{\tilde \Omega \cap \{\frac{m}{2} \leq u \leq m \}} (1+u^{-\beta}) \chi_{\{u>0\}}, \end{align*}

for every $m>0$.

The claim follows by letting $m\to 0$ and by using Lemma 6.2.

As a immediate consequence of (90), (91),(92), (93) and (94), we have

\[ \int_\Omega \nabla u \nabla \varphi = \int_{\Omega \cap \{u>0\}} \big({-}u^{-\beta} + f(u) \big)\varphi \]

for every $\varphi \in C_c^{1}(\Omega )$. This concludes the proof of Theorem 1.1.

Proof of Theorem 1.2 The proof of this result is entirely analogous to the proof of Theorem 1.1. We only need to use Proposition 4.3 instead of Proposition 3.2 and consider $f(s)=\lambda s+s^{2^{*}-1}$.

Acknowledgements

M.F.S.has been partially supported by CAPES. The author thanks the anonymous referees for their valuable suggestions.

Competing interest declaration

The author declares none.

References

Ambrosetti, A., Brezis, H. and Cerami, G., Combined effects of concave and convex nonlinearities in some elliptic problems, J. Funct. Anal. 122 (1994), 519543.CrossRefGoogle Scholar
Anello, G. and Faraci, F., On a resonant elliptic problem with singular terms, Nonlinear Anal. 195 (2020), 111818.CrossRefGoogle Scholar
Aris, R., The mathematical theory of diffusion and reaction in permeable catalysts (Clarendon Press, 1975).Google Scholar
Bai, Y., Gasiński, L. and Papageorgiou, N. S., Positive solutions for nonlinear singular superlinear elliptic equations, Positivity 23 (2019), 761778.CrossRefGoogle Scholar
Brezis, H. and Nirenberg, L., Positive solutions of nonlinear elliptic equations involving critical Sobolev exponents, Comm. Pure Appl. Math. 36 (1983), 437477.CrossRefGoogle Scholar
Copetti, M. I. M. and Elliott, C. M., Numerical analysis of the Cahn-Hilliard equation with a logarithmic free energy, Numer. Math. 63 (1992), 3965.CrossRefGoogle Scholar
Costa, D. G., An invitation to variational methods in differential equations (Birkhäuser, 2007).CrossRefGoogle Scholar
Dal Passo, R., Giacomelli, L. and Novick-Cohen, A., Existence for an Allen-Cahn/Cahn-Hilliard system with degenerate mobility, Interfaces Free Bound 1 (1999), 199226.CrossRefGoogle Scholar
Dávila, J. and Montenegro, M., Positive versus free boundary solutions to a singular elliptic equation, J. Anal. Math. 90 (2003), 303335.CrossRefGoogle Scholar
Dávila, J. and Montenegro, M., Concentration for an elliptic equation with singular nonlinearity, J. Math. Pures Appl. 97 (2012), 545578.CrossRefGoogle Scholar
Diaz, J. I., Nonlinear partial differential equations and free boundaries, Volume I. Elliptic equations, Research Notes in Mathematics, Vol. 106 (Boston, Pitman, 1985).Google Scholar
Diaz, J. I., Morel, J. M. and Oswald, L., An elliptic equation with singular nonlinearity, Commun. Partial Differ. Equ. 12 (1987), 13331344.CrossRefGoogle Scholar
Elliot, C. M. and Garcke, H., On the Cahn-Hilliard equation with degenerate mobility, SIAM J. Math. Anal. 27 (1996), 404423.CrossRefGoogle Scholar
Figueiredo, G. and Montenegro, M., A class of elliptic equations with singular and critical nonlinearities, Acta Appl. Math. 143 (2016), 6389.CrossRefGoogle Scholar
Friedman, A. and Phillips, D., The free boundary of a semilinear elliptic equation, Trans. AMS 282 (1984), 153182.CrossRefGoogle Scholar
Gilardi, G. and Rocca, E., Well-posedness and long-time behaviour for a singular phase field system of conserved type, IMA J. Appl. Math. 72 (2007), 498530.CrossRefGoogle Scholar
Hirano, N., Saccon, C. and Shioji, N., Existence of multiple positive solutions for singular elliptic problems with concave and convex nonlinearities, Adv. Differ. Equ. 9 (2004), 197220.Google Scholar
Lions, P. L., The concentration-compactness principle in the calculus of variations. The limit case, Rev. Mat. Iberoamericana 1 (1985), 145201.CrossRefGoogle Scholar
Long, Y. M., Sun, Y. J. and Wu, S. P., Combined effects to singular and superlinear nonlinearities in some singular boundary value problems, J. Differ. Equ. 176 (2003), 511531.Google Scholar
Lorca, S. and Montenegro, M., Free boundary solutions to a log-singular elliptic equation, Asymptot. Anal. 82 (2013), 91107.Google Scholar
Montenegro, M. and Queiroz, O., Existence and regularity to an elliptic equation with logarithmic nonlinearity, J. Differ. Equ. 246 (2009), 482511.CrossRefGoogle Scholar
Montenegro, M. and Silva, E. A. B., Two solutions for a singular elliptic equation by variational methods, Ann. Scuola Normale Sup. Pisa 11 (2012), 143165.Google Scholar
Papageorgiou, N. S. and Rădulescu, V., Combined effects of singular and sublinear nonlinearities in some elliptic problems, Nonlinear Anal. 109 (2014), 236244.CrossRefGoogle Scholar
Papageorgiou, N. S. and Smyrlis, G., A bifurcation-type theorem for singular nonlinear elliptic equations, Methods Appl. Anal. 22 (2015), 147170.CrossRefGoogle Scholar
Perera, K. and Silva, E. A. B., Existence and multiplicity of positive solutions for singular quasilinear problems, J. Math. Anal. Appl. 323 (2006), 12381252.CrossRefGoogle Scholar
Phillips, D., A minimization problem and the regularity of solutions in the presence of a free boundary, Indiana Univ. Math. J. 32 (1983), 117.CrossRefGoogle Scholar
Shi, J. and Yao, M., On a singular nonlinear semilinear elliptic problem, Proc. Roy. Soc. Edinburgh Sect. A 128 (1998), 13891401.CrossRefGoogle Scholar
Wang, X., Qin, X. and Hu, G., Existence of weak positive solution for a singular elliptic problem with supercritical nonlinearity, Anal. Math. Phys. 8 (2018), 4355.CrossRefGoogle Scholar
Willem, M., Minimax theorems (Birkhäuser, 1996).CrossRefGoogle Scholar
Zhang, Z., On a Dirichlet problem with a singular nonlinearity, J. Math. Anal. Appl. 194 (1995), 103113.CrossRefGoogle Scholar