1 Introduction

We prove the existence of solution of the problem

$$\begin{aligned} \left\{ \begin{array}{lll} -\Delta v = \lambda v^{q} + f(v) &{}\hbox {in}&{} \Omega \\ v>0 &{}\hbox {in}&{} \Omega \\ v=0 &{}\hbox {on}&{} \partial \Omega , \end{array} \right. \end{aligned}$$
(1)

where \(\Omega \subset \mathbb {R}^2\) is a bounded domain with smooth boundary, \(\lambda >0\) is a parameter, \(0<q<1\), \(f: [0, \infty ) \rightarrow \mathbb {R}\) is a continuous function, and

$$\begin{aligned} 0\le f(s)s\le C|s|^p\exp (\alpha \,s^2) \end{aligned}$$
(2)

where \(2< p<+\infty \) and \(\alpha >0\).

We state our main result.

Theorem 1.1

Suppose that \(f:[0,\infty ) \rightarrow \mathbb {R}\) is a continuous function satisfying (2). Then there exists \(\lambda ^*>0\) such that for every \(\lambda \in (0,\lambda ^*)\), the problem (1) has a positive weak solution \(u \in H^1_0(\Omega ) \cap H^2(\Omega )\).

Elliptic problems of the type

$$\begin{aligned} \left\{ \begin{array}{lll} -\Delta v = g(x,v) &{}\hbox {in}&{} \Omega \\ v=0 &{}\hbox {on}&{} \partial \Omega , \end{array} \right. \end{aligned}$$
(3)

in \(\Omega \subset \mathbb {R}^2\) where g(xv) is continuous and behaves like \(\exp (\alpha |v|^{2})\) as \(|v| \rightarrow +\infty \) have been studied by many authors, see [6, 1012, 16, 19]. One of the main ingredients is the Trudinger–Moser inequality introduced in [18, 21], namely. Given \(u \in H^1_0(\Omega )\), then

$$\begin{aligned} e^{\sigma |u|^{2}} \in L^1(\Omega )\,\,\,\text{ for } \text{ every }\,\,\,\sigma >0, \end{aligned}$$
(4)

and there exists a positive constant L such that

$$\begin{aligned} \sup _{\Vert u\Vert _{H^1_0(\Omega )}\le 1}\int _{\Omega }e^{\sigma |u|^{2}}{\hbox {d}}x \le L\,\,\,\text{ for } \text{ every }\,\,\,\sigma \le 4\pi . \end{aligned}$$
(5)

We say that g has subcritical growth at \(+\infty \) if for every \(\sigma >0\)

$$\begin{aligned} \lim _{s \rightarrow +\infty }\frac{|g(x,s)|}{e^{\sigma s^2}}=0 \end{aligned}$$

and g has critical growth at \(+\infty \) if there exists \(\sigma _0>0\) such that

$$\begin{aligned} \lim _{s \rightarrow +\infty }\frac{|g(x,s)|}{e^{\sigma s^2}}=0 \, \forall \, \sigma >\sigma _0 \quad \text{ and } \,\, \lim _{s \rightarrow +\infty }\frac{|g(x,s)|}{e^{\sigma s^2}}=+\infty \, \forall \, \sigma <\sigma _0. \end{aligned}$$

The only assumptions we assume are that \(0<q<1\), f is continuous and satisfies the growth assumption (2), and thus the nonlinearity \(g(s)=\lambda s^q + f(s)\) of problem (1) can have subcritical or critical behavior at \(+\infty \).

Most papers treat problem (3) by means of variational methods, and then usually it is assumed that g has subcritical or critical growth and sometimes \(g(s) \ge c |s|^p\), where \(c>0\) is a constant, see [12]. Another common assumption on g is the so-called Ambrosetti–Rabinowitz condition

$$\begin{aligned} \exists R>0 \text{ and } \theta > 2 \text{ such } \text{ that } 0< \theta G(x,s) \le s g(x,s) \,\, \forall |s| \ge R \text{ and } x \in \Omega , \end{aligned}$$

where \(G(s)=\int _0^s g(t){\hbox {d}}t\), see [1012].

Even when the Ambrosetti–Rabinowitz can be dropped, some conditions have to be assumed to give compactness of the Palais-Samle sequences or Cerami sequences, see for instance [16] where they assume

$$\begin{aligned}&g: \overline{\Omega }\times \mathbb {R} \text{ is } \text{ continuous } \text{ and } g(x,0) = 0;\\&\quad \exists t_0 > 0 \text{ and } M > 0 \text{ such } \text{ that } 0 < G(x,s) \le M g(x,s) \,\, \forall |s| \ge t_0 \text{ and } x \in \Omega ;\\&0 < 2 G(x,s) \le s g(x,s) \,\, \forall |s| \ge 0 \text{ and } x \in \Omega . \end{aligned}$$

A problem in \(\Omega = \mathbb {R}^2\) without Ambrosetti–Rabinowitz condition and exponential growth on g different from (2) has been addressed in [15].

We are able to solve (1) under weaker assumptions by using the Galerkin method. For that matter, we approximate f by Lipschitz functions in Sect. 2. We solve the approximating problems (11) in Sect. 3. Section 4 is devoted to prove Theorem 1.1, and in doing so, we show that the solutions \(v_n\) of problem (11) are bounded away from zero and converge to a positive solution of (1).

Problem (1) is also studied in \(\mathbb {R}^2\), see for instance [1, 2, 4, 8, 22]. Problems with nonlinearities with exponential growth are also important in conformal geometry [9, 17].

2 Approximating functions

To prove Theorem 1.1, we approximate f by Lipschitz functions \(f_k:\mathbb {R} \rightarrow \mathbb {R}\) defined by

$$\begin{aligned} f_k(s)=\displaystyle \left\{ \begin{array}{lll} -k\left[ G\left( -k-\frac{1}{k}\right) - G(-k)\right] , &{}\hbox {if}&{} s\le -k\\ -k\left[ G\left( s-\frac{1}{k}\right) - G(s)\right] , &{}\hbox {if}&{} -k\le s \le -\frac{1}{k}\\ k^2s\left[ G\left( -\frac{2}{k}\right) - G\left( -\frac{1}{k}\right) \right] , &{}\hbox {if}&{} -\frac{1}{k}\le s\le 0\\ k^2s\left[ G\left( \frac{2}{k}\right) - G\left( \frac{1}{k}\right) \right] , &{}\hbox {if}&{} 0\le s\le \frac{1}{k}.\\ k\left[ G\left( s+\frac{1}{k}\right) - G(s)\right] , &{}\hbox {if}&{} \frac{1}{k}\le s \le k\\ k\left[ G\left( k+\frac{1}{k}\right) - G(k)\right] , &{}\hbox {if}&{} s\ge k.\\ \end{array} \right. \end{aligned}$$
(6)

where \(G(s)=\int _0^sf(\xi ){\hbox {d}}\xi \).

The following approximation result was proved in [20].

Lemma 2.1

Let \(f:\mathbb {R} \rightarrow \mathbb {R}\) be a continuous function such that \(sf(s)\ge 0\) for every \(s \in \mathbb {R}\). Then there exists a sequence \(f_k:\mathbb {R} \rightarrow \mathbb {R}\) of continuous functions satisfying

  1. (i)

    \(sf_k(s)\ge 0\) for every \(s \in \mathbb {R}\);

  2. (ii)

    \(\forall \, k \in \mathbb {N}\) \(\exists c_k>0\) such that \(|f_k(\xi ) - f_k(\eta )|\le c_k|\xi - \eta |\) for every \(\xi , \eta \in \mathbb {R}\);

  3. (iii)

    \(f_k\) converges uniformly to f in bounded subsets of \(\mathbb {R}\).

The sequence \(f_k\) of the previous lemma has some additional properties.

Lemma 2.2

Let \(f: \mathbb {R} \rightarrow \mathbb {R}\) be a continuous function satisfying (2) for every \(s \in \mathbb {R}\). Then the sequence \(f_k\) of Lemma 2.1 satisfies

  1. (i)

    \(\forall \, k \in \mathbb {N}\), \(0\le sf_k(s) \le C_1|s|^p\exp (4\alpha \,s^2)\) for every \(|s|\ge \frac{1}{k}\);

  2. (ii)

    \(\forall \, k \in \mathbb {N}\), \(0\le sf_k(s) \le C_2|s|^2\exp (4\alpha \,s^2)\) for every \(|s|\le \frac{1}{k}\),

where \(C_1\) and \(C_2\) are positive constants independent of k.

Proof of Lemma 2.2

Everywhere in this proof the constant C is the one of (2).

First step. Suppose that \(-k\le s \le -\frac{1}{k}\).

By the mean value theorem, there exists \(\eta \in \left( s-\frac{1}{k},s \right) \) such that

$$\begin{aligned} f_k(s)=-k\left[ G\left( s-\frac{1}{k}\right) - G(s)\right] =-kG'(\eta )\left( s-\frac{1}{k}-s\right) =f(\eta ) \end{aligned}$$

and

$$\begin{aligned} sf_k(s)=sf(\eta ). \end{aligned}$$

Since \(s-\frac{1}{k}<\eta <s<0\) and \(f(\eta )<0\), we have \(sf(\eta ) \le \eta f(\eta )\). Therefore,

$$\begin{aligned} sf_k(s) \le \eta f(\eta )\le & {} C|\eta |^p\exp (\alpha \,|\eta |^2)\\\le & {} C|s-\frac{1}{k}|^p\exp \left( \alpha \,|s-\frac{1}{k}|^2\right) \\\le & {} C\left( |s| +\frac{1}{k}\right) ^p\exp \left( \alpha \,\left( |s| +\frac{1}{k}\right) ^2\right) \\\le & {} C(2|s|)^p\exp (\alpha \,(2|s|)^2)\\= & {} C2^p|s|^p\exp (4\alpha \,|s|^2). \end{aligned}$$

Second step. Assume \(\frac{1}{k}\le s \le k\).

By the mean value theorem, there exists \(\eta \in \left( s,s+\frac{1}{k}\right) \) such that

$$\begin{aligned} f_k(s)=k\left[ G\left( s+\frac{1}{k}\right) - G(s)\right] =kG'(\eta )\left( s+\frac{1}{k}-s\right) =f(\eta ) \end{aligned}$$

and

$$\begin{aligned} sf_k(s)=sf(\eta ). \end{aligned}$$

Since \(0<s <\eta <s+\frac{1}{k}\) and \(f(\eta )>0\), we have \(sf(\eta ) \le \eta f(\eta )\). Therefore,

$$\begin{aligned} sf_k(s) \le \eta f(\eta )\le & {} C|\eta |^p\exp (\alpha \,|\eta |^2)\\\le & {} C|s+\frac{1}{k}|^p\exp \left( \alpha \,|s+\frac{1}{k}|^2\right) \\\le & {} C(2|s|)^p\exp (\alpha \,(2|s|)^2)\\= & {} C2^p|s|^p\exp (4\alpha \,|s|^2). \end{aligned}$$

Third step. Suppose that \(|s|\ge k\), then

$$\begin{aligned} f_k(s)=\displaystyle \left\{ \begin{array}{lcc} -k\left[ G\left( -k-\frac{1}{k}\right) - G(-k)\right] , &{}\hbox {if}&{} s\le -k\\ k[G\left( k+\frac{1}{k}\right) - G(k)], &{}\hbox {if}&{} s\ge k.\\ \end{array} \right. \end{aligned}$$
(7)

If \(s\le -k\), by the mean value theorem, there exists \(\eta \in \left( -k-\frac{1}{k},-k\right) \) such that

$$\begin{aligned} f_k(s)=k\left[ G\left( -k-\frac{1}{k}\right) - G(-k)\right] =-kG'(\eta )\left( -k-\frac{1}{k}-(-k)\right) =f(\eta ) \end{aligned}$$

and

$$\begin{aligned} sf_k(s)=sf(\eta ). \end{aligned}$$

Since \(-k-\frac{1}{k} <\eta <-k<0\) and \(k<|\eta | < k + \frac{1}{k}\), we conclude that

$$\begin{aligned} sf_k(s)=\frac{s}{\eta }\eta f(\eta )\le & {} \frac{|s|}{|\eta |}C|\eta |^p\exp (\alpha \,|\eta |^2) =C|s||\eta |^{p-1}\exp (\alpha \,|\eta |^2)\nonumber \\\le & {} C|s|\left( k + \frac{1}{k}\right) ^{p-1}\exp \left( \alpha \,\left( k + \frac{1}{k}\right) ^2\right) \nonumber \\\le & {} C|s|\left( |s| + \frac{1}{k}\right) ^{p-1}\exp \left( \alpha \,\left( |s| + \frac{1}{k}\right) ^2\right) \nonumber \\\le & {} C|s|(2|s|)^{p-1}\exp (\alpha \,(2|s|)^2)\nonumber \\\le & {} C2^{p-1}|s|^{p}\exp (4\alpha \,|s|^2). \end{aligned}$$
(8)

If \(s\ge k\), by the mean value theorem, there exists \(\eta \in \left( k,k + \frac{1}{k}\right) \) such that

$$\begin{aligned} f_k(s)=k\left[ G\left( k+\frac{1}{k}\right) - G(k)\right] =kG'(\eta )\left( k+\frac{1}{k}-k\right) =f(\eta ). \end{aligned}$$

By computations similar to conclude (8) one has

$$\begin{aligned} sf_k(s)=sf(\eta )=\frac{s}{\eta }\eta f(\eta ) \le \frac{|s|}{|\eta |}C|\eta |^p \exp (\alpha \,|\eta |^2) \le C2^{p-1}|s|^{p}\exp (4\alpha \,|s|^2). \end{aligned}$$

Fourth step. Assume \(-\frac{1}{k}\le s\le \frac{1}{k}\), then

$$\begin{aligned} f_k(s)=\displaystyle \left\{ \begin{array}{lcc} k^2s\left[ G\left( -\frac{2}{k}\right) - G\left( -\frac{1}{k}\right) \right] , &{}\hbox {if}&{} -\frac{1}{k}\le s\le 0\\ k^2s\left[ G\left( \frac{2}{k}\right) - G\left( \frac{1}{k}\right) \right] , &{}\hbox {if}&{} 0\le s\ge \frac{1}{k}.\\ \end{array} \right. \end{aligned}$$
(9)

If \(-\frac{1}{k}\le s\le 0\), by the mean value theorem, there exists \(\eta \in (-\frac{2}{k},-\frac{1}{k})\) such that

$$\begin{aligned} f_k(s)=k^2s\left[ G\left( -\frac{2}{k}\right) - G\left( -\frac{1}{k}\right) \right] =k^2sG'(\eta )\left( -\frac{2}{k}-\left( -\frac{1}{k}\right) \right) =-ksf(\eta ). \end{aligned}$$

Therefore,

$$\begin{aligned} sf_k(s)= & {} -ks^2f(\eta )=-k\frac{s^2}{\eta }\eta f(\eta ) \le k\frac{s^2}{|\eta |}\eta f(\eta ) \nonumber \\&\le Ck|s|^2|\eta |^{p-1}\exp (\alpha \,|\eta |^2) \le Ck|s|^2(\frac{2}{k})^{p-1}\exp (\alpha \,|\eta |^2)\nonumber \\&\le C2^{p-1}|s|^2\exp \left( \alpha \,\left( \frac{2}{k}\right) ^2\right) \nonumber \\&\le C2^{p-1}|s|^2\exp (4\alpha ) \le C2^{p-1}\exp (4\alpha )|s|^2\exp (4\alpha \,|s|^2). \end{aligned}$$
(10)

If \(0\le s \le \frac{1}{k}\), by the mean value theorem, there exists \(\eta \in (\frac{1}{k},\frac{2}{k})\) such that

$$\begin{aligned} f_k(s)=k^2s[G(\frac{2}{k}) - G(\frac{1}{k})]=k^2sG'(\eta )(\frac{2}{k}-\frac{1}{k})=ksf(\eta ). \end{aligned}$$

By similar computations to conclude (10) one obtains

$$\begin{aligned} sf_k(s)=ks^2f(\eta )=k\frac{s^2}{|\eta |}\eta f(\eta ) \le C2^{p-1}\exp (4\alpha )|s|^2\exp (4\alpha \,|s|^2). \end{aligned}$$

The proof of the lemma follows by taking \(C_1=C2^p\) ad \(C_2=C2^{p-1}\), where C is given in (2). \(\square \)

3 Approximate equation

To prove Theorem 1.1, we first show the existence of a solution of the following auxiliary problem

$$\begin{aligned} \left\{ \begin{array}{lcc} -\Delta v = \lambda v^{q} + f_n(v) + \frac{1}{n} &{}\hbox {in}&{} \Omega \\ v>0 &{}\hbox {in}&{} \Omega \\ v=0 &{}\hbox {on}&{} \partial \Omega , \end{array} \right. \end{aligned}$$
(11)

where \(f_n\) are given by Lemma 2.1 and Lemma 2.2.

We will use the Galerkin method together with the following fixed point theorem, see [20] and [14, Theorem5.2.5]. A similar approach was already used in [3].

Lemma 3.1

Let \(F: \mathbb {R}^d \rightarrow \mathbb {R}^d\) be a continuous function such that \(\left\langle F(\xi ),\xi \right\rangle \ge 0\) for every \(\xi \in \mathbb {R}^d\) with \(|\xi |=r\) for some \(r>0\). Then, there exists \(z_0\) in the closed ball \(\overline{B}_r(0)\) such that \(F(z_0)=0\).

The main result in this section is the following.

Lemma 3.2

There exist \(\lambda ^*>0\) and \(n^* \in \mathbb {N}\) such that (11) has a weak nonnegative and nontrivial solution for every \(\lambda \in (0,\lambda ^*)\) and \(n\ge n^*\).

Proof of Lemma 3.2

Let \(\mathcal {B}=\{w_1,w_2,\dots ,w_m,\dots \}\) be an orthonormal basis of \(H_0^1(\Omega )\) and define

$$\begin{aligned} W_m=[w_1,w_2,\dots ,w_m], \end{aligned}$$

to be the space generated by \(\{w_1,w_2,\dots ,w_m\}\). Define the function \(F:\mathbb {R}^m \rightarrow \mathbb {R}^m\) such that \(F(\xi )=(F_1(\xi ),F_2(\xi ),\dots , F_m(\xi ))\), where \(\xi =(\xi _1, \xi _2, ..., \xi _m) \in \mathbb {R}^m\),

$$\begin{aligned} F_j(\xi )=\int _{\Omega }\nabla v\nabla w_j - \lambda \int _{\Omega }(v_+)^{q}w_j - \int _{\Omega }f_n(v_+)w_j - \frac{1}{n}\int _{\Omega }w_j, \,\,\,j=1,2,\dots ,m \end{aligned}$$

and \(v=\sum _{i=1}^m\xi _i w_i\) belongs to \(W_m\). Therefore,

$$\begin{aligned} \left\langle F(\xi ),\xi \right\rangle =\int _{\Omega }|\nabla v|^2 - \lambda \int _{\Omega }(v_+)^{q+1} - \int _{\Omega }f_n(v_+)v_+ - \frac{1}{n}\int _{\Omega }v, \end{aligned}$$
(12)

where \(v_+ = \max \{v,0\}\) and \(v_- = v_+ - v\).

Given \(v \in W_m\), we define

$$\begin{aligned} \Omega ^+_n=\left\{ x \in \Omega : |v(x)|\ge \frac{1}{n}\right\} \end{aligned}$$

and

$$\begin{aligned} \Omega ^-_n=\left\{ x \in \Omega : |v(x)|< \frac{1}{n}\right\} . \end{aligned}$$

Thus, we rewrite (12) as

$$\begin{aligned} \left\langle F(\xi ),\xi \right\rangle = \left\langle F(\xi ),\xi \right\rangle _P + \left\langle F(\xi ),\xi \right\rangle _{N}, \end{aligned}$$

where

$$\begin{aligned} \left\langle F(\xi ),\xi \right\rangle _P=\int _{\Omega ^+_n}|\nabla v|^2 - \lambda \int _{\Omega ^+_n}(v_+)^{q+1} - \int _{\Omega ^+_n}f_n(v_+)v_+ - \frac{1}{n}\int _{\Omega ^+_n}v \end{aligned}$$

and

$$\begin{aligned} \left\langle F(\xi ),\xi \right\rangle _{N}=\int _{\Omega ^-_n}|\nabla v|^2 - \lambda \int _{\Omega ^-_n}(v_+)^{q+1} - \int _{\Omega ^-_n}f_n(v_+)v_+ - \frac{1}{n}\int _{\Omega ^-_n}v. \end{aligned}$$

Step 1. Since \(0<q<1\), then

$$\begin{aligned} \int _{\Omega ^+_n}(v_+)^{q+1} \le \int _{\Omega }|v|^{q+1} = \Vert v\Vert ^{q+1}_{L^{q+1}(\Omega )}\le C_1\Vert v\Vert ^{q+1}_{H^1_0(\Omega )}. \end{aligned}$$
(13)

By virtue of Lemma 2.2 (i), we get

$$\begin{aligned} \displaystyle \int _{\Omega ^+_n}f_n(v_+)v_+\le & {} \displaystyle C_1\int _{\Omega ^+_n}|v_+|^p\exp (4\alpha |v_+|^2){\hbox {d}}x\nonumber \\\le & {} \displaystyle C_1 \left( \int _{\Omega }|v_+|^{p+1}\right) ^{\frac{p}{p+1}}\left( \int _{\Omega }\exp (4\alpha (p+1)|v_+|^2){\hbox {d}}x\right) ^{\frac{1}{p+1}}\nonumber \\\le & {} \displaystyle C_1\Vert v\Vert _{L^{p+1}(\Omega )}^{p}\left( \int _{\Omega }\exp (4\alpha (p+1)|v|^2){\hbox {d}}x\right) ^{\frac{1}{p+1}}. \end{aligned}$$
(14)

It follows from (13) and (14) that

$$\begin{aligned} \begin{array}{rcl} \left\langle F(\xi ),\xi \right\rangle _P &{}\ge &{}\displaystyle \int _{\Omega ^+_n}|\nabla v|^2 - \lambda C_0\Vert v\Vert ^{q+1}_{H^1_0(\Omega )}\\ &{}-&{} \displaystyle C_1\Vert v\Vert ^{p}_{H^1_0(\Omega )}\left( \int _{\Omega }\exp (4\alpha (p+1)|v|^2){\hbox {d}}x\right) ^{\frac{1}{p+1}} - \frac{C_3}{n}\Vert v\Vert _{H^1_0(\Omega )}, \end{array} \end{aligned}$$
(15)

where \(C_0\), \(C_1\), and \(C_3\) are constants depending only on C, p, and \(|\Omega |\).

Step 2. Since \(0<q<1\), then

$$\begin{aligned} \int _{\Omega ^-_n}(v_+)^{q+1} \le \int _{\Omega ^-_n}|v|^{q+1} \le |\Omega |\frac{1}{n^{q+1}}. \end{aligned}$$
(16)

By virtue of Lemma 2.2 (ii), we get

$$\begin{aligned} \int _{\Omega ^-_n}f_n(v_+)v_+ \le C_2 \int _{\Omega ^-_n}|v_+|^2\exp (4\alpha |v_+|^2){\hbox {d}}x\le C_2 \exp (4\alpha )|\Omega |\frac{1}{n^2}. \end{aligned}$$
(17)

It follows from (16) and (17) that

$$\begin{aligned} \left\langle F(\xi ),\xi \right\rangle _N \ge \int _{\Omega ^-_n}|\nabla v|^2 - \lambda |\Omega |\frac{1}{n^{q+1}} - C_2 \exp (4\alpha )|\Omega |\frac{1}{n^2} - |\Omega |\frac{1}{n^2}. \end{aligned}$$
(18)

Thus, (15) and (18) imply

$$\begin{aligned} \begin{array}{rcl} \displaystyle \left\langle F(\xi ),\xi \right\rangle &{}\ge &{} \displaystyle \Vert v\Vert ^2_{H^1_0(\Omega )} - \lambda C_0\Vert v\Vert ^{q+1}_{H^1_0(\Omega )} - C_1\Vert v\Vert ^{p}_{H^1_0(\Omega )}\left( \int _{\Omega }\exp (4\alpha (p+1)|v|^2){\hbox {d}}x\right) ^{\frac{1}{p+1}} \\ &{} - &{}\displaystyle \frac{C_3}{n}\Vert v\Vert _{H^1_0(\Omega )} - \lambda |\Omega |\frac{1}{n^{q+1}} - C_2 \exp (4\alpha )|\Omega |\frac{1}{n^2} - |\Omega |\frac{1}{n^2}. \end{array} \end{aligned}$$
(19)

Assume now that \(\Vert v\Vert _{H^1_0(\Omega )}=r\) for some \(r>0\) to be chosen later. We have

$$\begin{aligned} \int _{\Omega }\exp (4\alpha (p+1)|v|^2){\hbox {d}}x = \int _{\Omega }\exp \left( 4\alpha (p+1)r^2\left( \frac{v}{\Vert v\Vert _{H^1_0(\Omega )}}\right) ^2\right) {\hbox {d}}x \end{aligned}$$
(20)

and in order to apply the Trudinger–Moser inequality (5), we must have \(4\alpha (p+1)r^2 \le 4\pi \). Consequently,

$$\begin{aligned} r\le \left( \frac{\pi }{\alpha (p+1)}\right) ^{\frac{1}{2}}. \end{aligned}$$

Then

$$\begin{aligned} \sup _{\Vert v\Vert _{H^1_0(\Omega )}\le 1}\int _{\Omega }\exp \left( 4\alpha (p+1)r^2\left( \frac{v}{\Vert v\Vert _{H^1_0(\Omega )}}\right) ^2\right) {\hbox {d}}x \le L. \end{aligned}$$

Hence,

$$\begin{aligned} \left\langle F(\xi ),\xi \right\rangle \ge r^2 - \lambda C_0 r^{q+1} - C_1 L^{\frac{1}{p+1}}r^p {-} \frac{C_3}{n}r - \lambda |\Omega |\frac{1}{n^{q+1}} {-} C_2 \exp (4\alpha )|\Omega |\frac{1}{n^2} - |\Omega |\frac{1}{n^2}. \end{aligned}$$

We need to choose r such that

$$\begin{aligned} r^2 - C_1 L^{\frac{1}{p+1}}r^p\ge \frac{r^2}{2}; \end{aligned}$$

in other words,

$$\begin{aligned} r\le \frac{1}{(2 C_1 L^{\frac{1}{p+1}})^{\frac{1}{p-2}}}; \end{aligned}$$

thus, let \(r=\min \left\{ \frac{1}{2(2C_1 L^{\frac{1}{p+1}})^{\frac{1}{p-2}}}, \left( \frac{\pi }{\alpha (p+1)}\right) ^{\frac{1}{2}}\right\} \), and hence

$$\begin{aligned} \left\langle F(\xi ),\xi \right\rangle \ge \frac{r^2}{2} - \lambda C_0 r^{q+1} - \frac{C_3}{n}r - \lambda |\Omega |\frac{1}{n^{q+1}} - C_2 \exp (4\alpha )|\Omega |\frac{1}{n^2} - |\Omega |\frac{1}{n^2}. \end{aligned}$$

Now, defining \(\rho =\frac{r^2}{2} - \lambda C_0 r^{q+1}\), we choose \(\lambda ^*>0\) such that \(\rho >0\) for \(\lambda < \lambda ^*\). Therefore, we choose

$$\begin{aligned} \lambda ^*=\frac{r^{1-q}}{4C_0}. \end{aligned}$$

Now we choose \(n^* \in \mathbb {N}\) such that

$$\begin{aligned} \frac{C_3}{n}r + \lambda |\Omega |\frac{1}{n^{q+1}} + C_2 \exp (4\alpha )|\Omega |\frac{1}{n^2} + |\Omega |\frac{1}{n^2} <\frac{\rho }{2}, \end{aligned}$$

for every \(n\ge n^*\). Let \(\xi \in \mathbb {R}^m\), such that \(|\xi |=r\), then for \(\lambda < \lambda ^*\) and \(n\ge n^*\) we obtain

$$\begin{aligned} \left\langle F(\xi ),\xi \right\rangle \ge \frac{\rho }{2}>0. \end{aligned}$$
(21)

For every \(n \in \mathbb {N}\), \(f_n\) is a Lipschitz function, and then by Lemma 3.1 for every \(m \in \mathbb {N}\) there exists \(y \in \mathbb {R}^m\) with \(|y|\le r\) such that \(F(y)=0\), that is, there exists \(v_m \in W_m\) verifying

$$\begin{aligned} \Vert v_m\Vert _{H^1_0(\Omega )}\le r \,\, \text{ for } \text{ every } \,\, m \in \mathbb {N} \end{aligned}$$

and such that

$$\begin{aligned} \int _{\Omega }\nabla v_m\nabla w = \lambda \int _{\Omega }(v_{m+})^{q}w + \int _{\Omega }f_n(v_{m+})w + \frac{1}{n}\int _{\Omega }w, \,\,\, \forall \, w \in W_m. \end{aligned}$$
(22)

Since \(W_m \subset H^1_0(\Omega )\) \(\forall \, m \in \mathbb {N}\) and r does not depend on m, then \((v_m)\) is a bounded sequence in \(H^1_0(\Omega )\). Then, for some subsequence, there exists \(v \in H^1_0(\Omega )\) such that

$$\begin{aligned} v_m \rightharpoonup v \,\,\, \text{ weakly } \text{ in } \,\,\, H^1_0(\Omega ) \end{aligned}$$
(23)

and

$$\begin{aligned} v_m \rightarrow v \,\,\, \text{ in } \,\,\, L^2(\Omega )\,\,\, \text{ and } \text{ a.e. } \text{ in }\,\,\,\Omega . \end{aligned}$$
(24)

Let \(k \in \mathbb {N}\), then for every \(m\ge k\) we obtain

$$\begin{aligned} \int _{\Omega }\nabla v_m\nabla w_k = \lambda \int _{\Omega }(v_{m+})^{q}w_k + \int _{\Omega }f_n(v_{m+})w_k + \frac{1}{n}\int _{\Omega }w_k, \,\,\, \forall \, w_k \in W_k. \end{aligned}$$
(25)

It follows from (23) that

$$\begin{aligned} \int _{\Omega }\nabla v_m\nabla w_k \rightarrow \int _{\Omega }\nabla v\nabla w_k \,\,\,\text{ as }\,\,\, m \rightarrow \infty \end{aligned}$$
(26)

and by (24) one obtains

$$\begin{aligned} \int _{\Omega }f_n(v_{m+})w_k \rightarrow \int _{\Omega }f_n(v_{+})w_k \,\,\,\text{ as }\,\,\, m \rightarrow \infty . \end{aligned}$$
(27)

Indeed, by Lemma 2.1 (ii) it follows that \(|f_n(v_{m+}) - f_n(v_{+})|\le c_n|v_{m+}-v_{+}|\); hence,

$$\begin{aligned} \left| \int _{\Omega }f_n(v_{m+})w_k - \int _{\Omega }f_n(v_{+})w_k\right| \le c_n\Vert w_k\Vert _{L^2(\Omega )}\Vert v_{m}-v\Vert _{L^2(\Omega )} \,\,\,\text{ as }\,\,\, m \rightarrow \infty \end{aligned}$$

and then (24) implies (27). By (23), (27), and Sobolev compact imbedding, letting \(m \rightarrow \infty \) one has

$$\begin{aligned} \lambda \int _{\Omega }(v_{m+})^{q}w_k + \int _{\Omega }f_n(v_{m+})w_k + \frac{1}{n}\int _{\Omega }w_k \rightarrow \lambda \int _{\Omega }(v_{+})^{q}w_k + \int _{\Omega }f_n(v_{+})w_k + \frac{1}{n}\int _{\Omega }w_k. \end{aligned}$$
(28)

By (25), (26), and (28)

$$\begin{aligned} \int _{\Omega }\nabla v\nabla w_k = \lambda \int _{\Omega }(v_{+})^{q}w_k + \int _{\Omega }f_n(v_{+})w_k + \frac{1}{n}\int _{\Omega }w_k, \,\,\, \forall \, w_k \in W_k. \end{aligned}$$
(29)

Since \([W_k]_{k \in \mathbb {N}}\) is dense in \(H^1_0(\Omega )\), we conclude that

$$\begin{aligned} \int _{\Omega }\nabla v\nabla w = \lambda \int _{\Omega }(v_{+})^{q}w + \int _{\Omega }f_n(v_{+})w + \frac{1}{n}\int _{\Omega }w, \,\,\, \forall \, w \in H^1_0(\Omega ). \end{aligned}$$
(30)

Furthermore, \(v\ge 0\) in \(\Omega \). In fact, since \(v_- \in H^1_0(\Omega )\), then from (30) we obtain

$$\begin{aligned} \int _{\Omega }\nabla v\nabla v_- = \lambda \int _{\Omega }(v_{+})^{q}v_- + \int _{\Omega }f_n(v_{+})v_- + \frac{1}{n}\int _{\Omega }v_-. \end{aligned}$$

Hence,

$$\begin{aligned} - \Vert v_-\Vert ^2_{H^1_0(\Omega )} = \int _{\Omega }\nabla v\nabla v_- = \int _{\Omega }f_n(v_{+})v_- + \frac{1}{n}\int _{\Omega }v_- \ge 0, \end{aligned}$$

then \(v_- \equiv 0\) a.e. in \(\Omega \). \(\square \)

4 Proof of the main result

In this section, we prove Theorem 1.1. We will use the unique solution \(\widetilde{w}\) of the problem

$$\begin{aligned} \left\{ \begin{array}{lcc} -\Delta \widetilde{w} = \widetilde{w}^{q} &{}\hbox {in}&{} \Omega \\ \widetilde{w}>0 &{}\hbox {in}&{} \Omega \\ \widetilde{w}=0 &{}\hbox {on}&{} \partial \Omega , \end{array} \right. \end{aligned}$$
(31)

for \(0<q<1\), see for instance [7]. The solution \(\widetilde{w}\) allows us to bound from below the solutions \(v_n\) of (11).

The following lemma of [20, Theorem1.1] is used to show that \(v_n\) converges to a solution v of (1).

Lemma 4.1

Let \(\Omega \) be a bounded open set in \(\mathbb {R}^N\), \(u_k : \Omega \rightarrow \mathbb {R}\) be a sequence function, and \(g_k : \mathbb {R} \rightarrow \mathbb {R}\) be a sequence of functions such that \(g_k(u_k)\) are measurable in \(\Omega \) for every \(k \in \mathbb {N}\). Assume that \(g_k(u_k) \rightarrow v\) a.e. in \(\Omega \) and \(\int _{\Omega }|g_k(u_k)u_k|{\hbox {d}}x<C\) for a constant C independent of k. And suppose that for every \(B \subset \mathbb {R}\), B bounded, there is a constant \(C_B\) depending only on B such that \(|g_k(x)|\le C_B\), for all \(x \in B\) and \(k \in \mathbb {N}\). Then \(v \in L^1(\Omega )\) and \(g_k(u_k) \rightarrow v\) in \(L^1(\Omega )\).

Proof of Theorem 1.1

By Lemma 3.2, equation (11) has a weak solution \(v_n \in H^1_0(\Omega )\) for each \(n \in \mathbb {N}\). Since \(0<q<1\) and \(f_n\) is Lipschitz, then \(\lambda v_n^{q} + f_n(v_n) + \frac{1}{n} \in L^p(\Omega )\) with \(p > 2\). Hence, \(v_n \in C^{1,\alpha }(\overline{\Omega })\) with \(0<\alpha <1\), see [13]. Therefore, \(v_n \in H^1_0(\Omega )\cap C^{1,\alpha }(\overline{\Omega })\).

We have by (23) that

$$\begin{aligned} v_m \rightharpoonup v_n \,\,\, \text{ weakly } \text{ in } \,\,\, H^1_0(\Omega ) \,\, \text{ as } \,\, m \rightarrow \infty . \end{aligned}$$
(32)

Therefore,

$$\begin{aligned} \Vert v_n\Vert _{H^1_0(\Omega )} \le \liminf _{m \rightarrow \infty }\Vert v_m\Vert _{H^1_0(\Omega )}\le r, \,\, \forall \, n \in \mathbb {N}, \end{aligned}$$

and r does not depend on n. Thus, there exists \(v \in H^1_0(\Omega )\) such that

$$\begin{aligned} v_n \rightharpoonup v \,\,\, \text{ weakly } \text{ in } \,\,\, H^1_0(\Omega ) \,\, \text{ as } \,\, n \rightarrow \infty . \end{aligned}$$
(33)

By Sobolev compact imbedding for \(1 \le s < +\infty \),

$$\begin{aligned} v_n \rightarrow v \,\,\text{ in }\,\,L^s(\Omega ) \,\, \text{ and } \,\, \text{ a.e. } \text{ in } \,\, \Omega . \end{aligned}$$

Note that

$$\begin{aligned} \left\{ \begin{array}{lcc} -\Delta v_n \ge \lambda v_n^{q} , &{}\hbox {in}&{} \Omega \\ v_n>0 &{}\hbox {in}&{} \Omega \\ v_n=0 &{}\hbox {on}&{} \partial \Omega . \end{array} \right. \end{aligned}$$
(34)

By rescaling, thus \(w_n=\lambda ^{\frac{1}{q-1}}v_n\) and we obtain

$$\begin{aligned} -\Delta \left( \frac{w_n}{\lambda ^{\frac{1}{q-1}}}\right) \ge \lambda \left( \frac{w_n}{\lambda ^{\frac{1}{q-1}}}\right) ^q \end{aligned}$$

implying

$$\begin{aligned} -\Delta \,w_n\ge w_n^q. \end{aligned}$$
(35)

By Lemma 3.3 of [5], it follows that \(w_n \ge \widetilde{w}\) \(\forall \, n \in \mathbb {N}\), that is,

$$\begin{aligned} v_n \ge \lambda ^{\frac{1}{1-q}}\widetilde{w} \,\,\text{ a.e. } \text{ in }\,\,\Omega , \,\, \forall \, n \in \mathbb {N}. \end{aligned}$$
(36)

Letting \(n \rightarrow +\infty \) in (36), we obtain

$$\begin{aligned} v \ge \lambda ^{\frac{1}{1-q}}\widetilde{w} \,\,\text{ a.e. } \text{ in }\,\,\Omega \end{aligned}$$

showing that \(v>0\) in \(\Omega \).

We prove now that v is a solution of (1). Since

$$\begin{aligned} v_n \rightarrow v\,\, \text{ a.e. } \text{ in }\,\, \Omega , \end{aligned}$$

we have

$$\begin{aligned} f_n(v_n(x)) \rightarrow f(v(x))\,\, \text{ a.e. } \text{ in }\,\, \Omega , \end{aligned}$$
(37)

by the uniform convergence of Lemma 2.1 (iii).

Recall from (30) that

$$\begin{aligned} \int _{\Omega }\nabla v_n\nabla w = \lambda \int _{\Omega }(v_n)^{q}w + \int _{\Omega }f_n(v_n)w + \frac{1}{n}\int _{\Omega }w, \,\,\, \forall \, w \in H^1_0(\Omega ). \end{aligned}$$
(38)

Taking \(w=v_n\) in (38) and since \(v_n\) is bounded in \(H_0^1(\Omega )\), we obtain

$$\begin{aligned} \int _{\Omega }f_n(v_n)v_n{\hbox {d}}x \le C, \end{aligned}$$
(39)

for every \(n \in \mathbb {N}\), where \(C>0\) is a constant independent of n. By (37), (39), and by the expression of \(f_n\) defined in (6), the assumptions of Lemma 4.1 are satisfied, implying

$$\begin{aligned} f_n(v_n) \rightarrow f(v)\,\, \text{ in }\,\, L^1(\Omega ). \end{aligned}$$
(40)

It follows from (4) that \(e^{v^2} \in L^1(\Omega )\), and in view of (2) and Hölder inequality, we conclude that \(f(v) \in L^2(\Omega )\).

By (38), we have

$$\begin{aligned} \int _{\Omega }\nabla v\nabla w = \lambda \int _{\Omega }v^{q}w + \int _{\Omega }f(v)w, \,\,\, \forall \, w \in H^1_0(\Omega ). \end{aligned}$$
(41)

Since \(f(v) \in L^2(\Omega )\) and \(\lambda \,v^{q} \in L^2(\Omega )\), we conclude from (41) that \(v \in H^2(\Omega )\) and

$$\begin{aligned} -\Delta v= \lambda \,v^{q} + f(v). \end{aligned}$$

The proof of the theorem is complete. \(\square \)