1 Introduction

Let \(\Omega \subset \mathbb {R}^{N}\) be a smooth bounded domain, \(s\in (0, 1)\) and \(N>2s\). In this paper we deal with the following fractional elliptic problem with critical growth:

$$\begin{aligned} \left\{ \begin{array}{ll} (-\Delta )^{s} u = \lambda u + u_{+}^{2^{*}_{s}-1} + f(x) &{} \text{ in } \Omega ,\\ u=0 &{} \text{ in } \mathbb {R}^{N}\setminus \Omega , \end{array} \right. \end{aligned}$$
(1.1)

where \(2^{*}_{s}:= \frac{2N}{N-2s}\) is the fractional critical Sobolev exponent, \(\lambda >0\) is a constant, and \(u_{+}:= \max \{u, 0\}\). The leading operator \((-\Delta )^{s}\) is the so-called fractional Laplacian operator defined as follows:

$$\begin{aligned} (-\Delta )^{s}u(x):=C(N, s)\lim _{{{\,\mathrm{\varepsilon }\,}}\rightarrow 0} \int _{\mathbb {R}^{N}\setminus B_{{{\,\mathrm{\varepsilon }\,}}}(x)} \frac{u(x)-u(y)}{|x-y|^{N+2s}}\, dxdy, \quad C(N, s):=\pi ^{-\frac{N}{2}} 2^{2s} \frac{\Gamma (\frac{N+2s}{2})}{-\Gamma (-s)}. \end{aligned}$$

We recall that in recent years a great interest has been devoted to the analysis of fractional and non-local operators of elliptic type, both for their interesting theoretical structure and in view of concrete applications in many fields such as, among the others, optimization, finance, phase transitions, anomalous diffusion, crystal dislocation, conservation laws, fractional quantum mechanics, quasi-geostrophic flows; see for instance [4, 24] for more details on this topic.

When \(s\rightarrow 1\), for u smooth enough, \((-\Delta )^{s}u\rightarrow -\Delta u\) and our problem is strictly related to a class of problems which are known as the Ambrosetti–Prodi type. We recall that Ambrosetti and Prodi [2] in 1972 studied the following boundary value problem:

$$\begin{aligned} \left\{ \begin{array}{ll} -\Delta u= g(u)+f(x) &{} \text{ in } \Omega ,\\ u=0 &{} \text{ on } \partial \Omega , \end{array} \right. \end{aligned}$$
(1.2)

with \(f\in C^{0, \alpha }(\bar{\Omega })\), \(\alpha \in (0, 1)\), \(g\in C^{2}(\mathbb {R})\) is such that \(g(0)=0\), \(g''(t)>0\) for all \(t\in \mathbb {R}\) and

$$\begin{aligned} 0<\lim _{t\rightarrow -\infty }g'(t)<\mu _{1}<\lim _{t\rightarrow \infty }g'(t)<\mu _{2}, \end{aligned}$$

where \(0<\mu _{1}<\mu _{2}\le \dots \le \mu _{n}\le \mu _{n+1}\le \dots\) denote the eigenvalues of \((-\Delta , H^{1}_{0}(\Omega ))\). By using some results on differentiable mappings with singularities, the authors showed that there exists in \(C^{0, \alpha }(\bar{\Omega })\) a closed connected \(C^{1}\) manifold M of codimension 1 such that \(C^{0, \alpha }(\bar{\Omega })\setminus M\) consists exactly of two connected components \(A_{1}\) and \(A_{2}\) with the following properties:

  1. (1)

    if \(f\in A_{1}\), then (1.2) has no solution,

  2. (2)

    if \(f\in A_{2}\), then (1.2) has exactly two solutions,

  3. (3)

    if \(f\in M\), then (1.2) has a unique solution.

Subsequently the the pioneering work by Ambrosetti and Prodi [2], different variants and generalizations of (1.2) have been achieved by many authors; see for instance [1, 5,6,7, 10, 12, 14,15,16,17,18,19,20,21,22, 26]. In particular, Ruf and Srikanth [26] analyzed

$$\begin{aligned} \left\{ \begin{array}{ll} -\Delta u = \lambda u+(u_{+})^{p} + f(x) &{} \text{ in } \Omega ,\\ u=0 &{} \text{ on } \partial \Omega , \end{array} \right. \end{aligned}$$
(1.3)

with \(1<p<2^{*}-1\) if \(N\ge 3\), where \(2^{*}:=\frac{2N}{N-2}\) is the critical exponent, and \(1<p<\infty\) if \(N= 2\), \(\lambda >\mu _{1}, \lambda \ne \mu _{n}\) for all \(n\in \mathbb {N}\), \(f:=t\varphi _{1}+\bar{f}\), \(t\in \mathbb {R}\), \(\varphi _{1}\) is the positive and normalized first eigenfunction of \((-\Delta , H^{1}_{0}(\Omega ))\), and \(\bar{f}\in C^{0}(\Omega )\) is such that \(\int _{\Omega } \bar{f}\varphi _{1}\, dx=0\). By applying the generalized mountain pass theorem [25], the authors proved that there exists a constant \(T=T(\bar{f})\) such that for \(t>T\) there are at least two solutions of (1.3).

On the other hand, Brezis and Nirenberg [9] in 1983 focused on the existence and non-existence of positive solutions for the following nonlinear critical elliptic problem:

$$\begin{aligned} \left\{ \begin{array}{ll} -\Delta u= g(x, u)+|u|^{2^{*}-2}u &{} \text{ in } \Omega ,\\ u=0 &{} \text{ on } \partial \Omega , \end{array} \right. \end{aligned}$$
(1.4)

where \(g(\cdot , 0)=0\) and g is some perturbation of lower order of the critical power. When \(g(x, u)=\lambda u\), with \(\lambda >0\), the authors proved that:

  1. (A)

    if \(N\ge 4\), (1.4) has a positive solution for every \(\lambda \in (0, \mu _{1})\); moreover, (1.4) has no positive solution if \(\lambda \notin (0, \mu _{1})\) and \(\Omega\) is star-shaped,

  2. (B)

    when \(N=3\) and \(\Omega\) is a ball, (1.4) has a positive solution if and only if \(\lambda \in (\frac{\mu _{1}}{4}, \mu _{1})\).

Later, Deng [18] treated the existence of multiple solutions for the following superlinear problem involving the critical exponent:

$$\begin{aligned} \left\{ \begin{array}{ll} -\Delta u= |u|^{2^{*}-2}u+g(x, u)+f(x) &{} \text{ in } \Omega ,\\ u=0 &{} \text{ on } \partial \Omega , \end{array} \right. \end{aligned}$$

where \(g:\bar{\Omega }\times \mathbb {R}\rightarrow \mathbb {R}\in C^{1}\) satisfies \(\lim _{t\rightarrow \pm \infty } \frac{g(x, t)}{|t|^{2^{*}-1}}=0\) uniformly in \(x\in \Omega\), \(f:=t\varphi _{1}+\bar{f}\), \(t\in \mathbb {R}\), and \(\bar{f}\in C^{0, \alpha }(\bar{\Omega })\) is such that \(\int _{\Omega } \bar{f}\varphi _{1}\, dx=0\). Subsequently, motivated by [2, 9, 18, 26], De Figueiredo and Jianfu [16] investigated the existence and the multiplicity of solutions for the following critical Ambrosetti–Prodi type problem:

$$\begin{aligned} \left\{ \begin{array}{ll} -\Delta u =\lambda u+(u_{+})^{2^{*}_{s}-1} + f(x) &{} \text{ in } \Omega ,\\ u=0 &{} \text{ on } \partial \Omega , \end{array} \right. \end{aligned}$$
(1.5)

where \(f:=t \varphi _{1}+\bar{f}\), \(t\in \mathbb {R}\), and \(\bar{f}\in L^{r}(\Omega )\), with \(r>N\), is such that \(\int _{\Omega } \bar{f} \varphi _{1}\, dx=0\). More precisely, they established the following results:

  1. (a)

    if \(0<\lambda <\mu _{1}\), then there exists \(t_{0}=t_{0}(\bar{f})<0\) such that, if \(t<t_{0}\), (1.5) has a negative solution \(u_{t}\),

  2. (b)

    if \(\lambda >\mu _{1}\) and \(\bar{f}\) is such that \(\bar{f}\in \ker (-\Delta -\lambda )^{\perp }\) in the case that \(\lambda\) is an eigenvalue of \((-\Delta , H^{1}_{0}(\Omega ))\), then there exists \(t_{0}= t_{0}(\bar{f})>0\) such that, if \(t>t_{0}\), (1.5) has a negative solution \(u_{t}\),

  3. (c)

    if, in addition to either of the hypotheses above, one assumes that \(\lambda\) is not an eigenvalue of \((-\Delta , H^{1}_{0}(\Omega ))\) and that the dimension \(N>6\), then (1.5) has a second solution,

  4. (d)

    if \(\lambda =\mu _{1}\), there exists \({{\,\mathrm{\varepsilon }\,}}>0\) such that, if \(\Vert f\Vert _{L^{2}(\Omega )}<{{\,\mathrm{\varepsilon }\,}}\), then (1.5) has a solution.

To our knowledge, the critical fractional Ambrosetti–Prodi problem (1.1) has never been considered in the literature and the goal of this paper is to fill this gap. We underline that an Ambrosetti–Prodi type result for a subcritical problem driven by the fractional spectral Laplacian operator was studied in [3], and that several ideas used in [3] can be easily adapted to deal with the corresponding subcritical problem involving \((-\Delta )^{s}\).

Before stating our main theorems, we introduce some notations and definitions. Denote by

$$\begin{aligned} H^{s}(\mathbb {R}^{N}):=\left\{ u\in L^{2}(\mathbb {R}^{N}): \frac{u(x)-u(y)}{|x-y|^{\frac{N+2s}{2}}}\in L^{1}(\mathbb {R}^{2N})\right\} \end{aligned}$$

the usual fractional Sobolev space endowed with the norm

$$\begin{aligned} \Vert u\Vert _{H^{s}(\mathbb {R}^{N})}:=\left( \Vert u\Vert ^{2}_{L^{2}(\mathbb {R}^{N})}+[u]^{2}_{s}\right) ^{\frac{1}{2}}, \end{aligned}$$

where

$$\begin{aligned} {[}u]_{s}:=\left( \iint _{\mathbb {R}^{2N}} \frac{|u(x)-u(y)|^{2}}{|x-y|^{N+2s}} \, dxdy \right) ^{\frac{1}{2}}. \end{aligned}$$

Let \(\mathbb {H}_{0}^{s}(\Omega )\) be the closed linear subspace given by

$$\begin{aligned} \mathbb {H}_{0}^{s}(\Omega ):= \left\{ u\in H^{s}(\mathbb {R}^{N}) \, : \, u=0 \text{ a.e. } \text{ in } \mathbb {R}^{N}\setminus \Omega \right\} \end{aligned}$$

equipped with the norm

$$\begin{aligned} \Vert u\Vert _{\mathbb {H}_{0}^{s}(\Omega )}:=\sqrt{\frac{C(N, s)}{2}} [u]_{s}. \end{aligned}$$

It is well-known that \(\mathbb {H}_{0}^{s}(\Omega )\) is a Hilbert space with the inner product

$$\begin{aligned} \langle u, v\rangle _{\mathbb {H}_{0}^{s}(\Omega )} := \frac{C(N,s)}{2}\iint _{\mathbb {R}^{2N}} \frac{(u(x)-u(y))(v(x)-v(y))}{|x-y|^{N+2s}} \, dxdy, \text{ for } u, v\in \mathbb {H}_{0}^{s}(\Omega ), \end{aligned}$$

\(\mathbb {H}_{0}^{s}(\Omega )\) is continuously embedded in \(L^{r}(\mathbb {R}^{N})\) for all \(r\in [1, 2^{*}_{s}]\) and compactly embedded in \(L^{r}(\mathbb {R}^{N})\) for all \(r\in [1, 2^{*}_{s})\); see [24, Lemma 1.29 and Lemma 1.31]. In order to lighten the notation, in what follows we set \(C(N, s)/2:=1\). With

$$\begin{aligned} S_{*}:=\inf \left\{ \frac{[u]_{s}^{2}}{\Vert u\Vert ^{2}_{L^{2^{*}_{s}}(\mathbb {R}^N)}} \,:\, u\in H^{s}(\mathbb {R}^{N})\setminus \{0\} \right\} \end{aligned}$$

we indicate the best constant in the Sobolev embedding \(H^{s}(\mathbb {R}^{N})\subset L^{2^{*}_{s}}(\mathbb {R}^{N})\); see [13, Theorem 1.1].

Thanks to the space \(\mathbb {H}_{0}^{s}(\Omega )\), we can correctly encode the variational formulation of problem (1.1). We say that \(u\in \mathbb {H}_{0}^{s}(\Omega )\) is a weak solution to (1.1) if for every \(\phi \in \mathbb {H}_{0}^{s}(\Omega )\) it holds

$$\begin{aligned} \langle u, \phi \rangle _{\mathbb {H}_{0}^{s}(\Omega )}=\int _{\Omega } (\lambda u+u^{2^{*}_{s}-1}_{+}+f) \phi \, dx. \end{aligned}$$

Denote by \(0<\lambda _{1}<\lambda _{2}\le \dots \le \lambda _{n}\le \lambda _{n+1}\le \dots\) the eigenvalues of \(((-\Delta )^{s}, \mathbb {H}_{0}^{s}(\Omega ))\) with corresponding eigenfunctions \(e_{1}, e_{2}, \dots , e_{n}, e_{n+1}, \dots\) belonging to \(\mathbb {H}_{0}^{s}(\Omega )\). We recall that \(\lambda _{1}\) is simple, \(\lambda _{n}\rightarrow \infty\) as \(n\rightarrow \infty\), \(\{e_{n}\}_{n\in \mathbb {N}}\) is an orthonormal basis of \(L^{2}(\Omega )\) and an orthogonal basis of \(\mathbb {H}_{0}^{s}(\Omega )\), \(e_{n}\in C^{0, \alpha }(\bar{\Omega })\) for some \(\alpha \in (0, 1)\) and \(e_{1}>0\) in \(\Omega\); see [24, Proposition 3.1, Corollary 4.8 and Theorem 5.2].

Our principal results are the following.

Theorem 1

Let \(f:=te_{1}+h\) with \(h\in L^{\infty }(\Omega )\) such that

$$\begin{aligned} \int _{\Omega } h e_{1}\, dx=0. \end{aligned}$$

Suppose \(\lambda \in (\lambda _{k}, \lambda _{k+1})\) for some \(k\in \mathbb {N}\). Then:

  1. (i)

    there exists \(t_{0}= t_{0}(h)>0\) such that, if \(t>t_{0}\), (1.1) has a nonpositive solution \(u_{t}\in \mathbb {H}_{0}^{s}(\Omega )\),

  2. (ii)

    if in addition \(N>6s\), (1.1) has a second solution.

Theorem 2

Let \(\lambda =\lambda _{1}\) in (1.1). Suppose

$$\begin{aligned} \int _{\Omega } f e_{1} \, dx <0, \end{aligned}$$
(1.6)

and

$$\begin{aligned} \Vert f\Vert _{L^{2}(\Omega )}\le M_{1} \quad \text{ and } \quad -\int _{\Omega }f e_{1} \, dx <M_{2}. \end{aligned}$$
(1.7)

where \(M_{1}, M_{2}>0\) are defined as

$$\begin{aligned} M_{1}:= \frac{s}{N+2s} S_{*}^{\frac{N}{4s}} \left( \frac{N}{N+2s}\right) ^{\frac{N-2s}{4s}} \left( 1- \frac{\lambda _{1}}{\lambda _{2}}\right) ^{\frac{N+2s}{4s}} \lambda _{2}^{\frac{1}{2}}, \end{aligned}$$
(1.8)

and

$$\begin{aligned} M_{2}:= \min \left\{ \left( \frac{2s}{N+2s}\right) ^{\frac{N+2s}{2N}}S_{*}^{\frac{N+2s}{4s}}, \left( \frac{2s}{N+2s}\right) ^{\frac{N+2s}{2N}} \Vert e_{1}\Vert _{L^{2^{*}_{s}}(\Omega )} \left( \frac{N}{N+2s}\right) ^{\frac{N+2s}{4s}} \left( 1- \frac{\lambda _{1}}{\lambda _{2}}\right) ^{\frac{N+2s}{4s}} S_{*}^{\frac{N+2s}{4s}}\right\} . \end{aligned}$$
(1.9)

Then (1.1) has a solution.

The proofs of Theorem 1 and Theorem 2 will be obtained by applying suitable variational methods inspired by [16]. More precisely, in order to deduce the existence of a second solution in Theorem 1, we invoke the linking theorem [25] without the Palais–Smale condition. In particular, we use some standard estimates found in [24, 27] to deal with the fractional Brezis–Nirenberg type problem, and we prove some useful lemmas and crucial estimates needed to ensure that the energy functional associated with (1.1) satisfies all geometric assumptions of the linking theorem and that the solution accomplished as weak limit of a Palais–Smale sequence at the linking critical level is not trivial. Clearly, with respect to the local case \(s=1\), a more accurate analysis will be done to overcome some technical difficulties coming from the nonlocal character of \((-\Delta )^{s}\). Finally, we prove Theorem 2 by means of a minimization argument.

The paper is organized as follows. In Sect. 2 we give the proof of Theorem 1 and in Sect. 3 we prove Theorem 2.

2 Proof of Theorem 1

First we claim that (1.1) has a nonpositive solution \(u_{t}\). Let us note that all nonpositive solutions to (1.1) satisfy

$$\begin{aligned} \left\{ \begin{array}{ll} (-\Delta )^{s}u = \lambda u+ te_{1}+h &{} \text{ in } \Omega ,\\ u=0 &{} \text{ in } \mathbb {R}^{N}\setminus \Omega . \end{array} \right. \end{aligned}$$

By exploiting the Fredholm alternative, it is easy to check that the problem

$$\begin{aligned} \left\{ \begin{array}{ll} (-\Delta )^{s}u=\lambda u + h &{} \text{ in } \Omega ,\\ u=0 &{} \text{ in } \mathbb {R}^{N}\setminus \Omega , \end{array} \right. \end{aligned}$$

admits a unique solution \(u_{0}\in \mathbb {H}_{0}^{s}(\Omega )\). From \(h\in L^{\infty }(\Omega )\) and [24, Proposition 4.11], we deduce that \(u_{0}\in L^{\infty }(\Omega )\). On the other hand, \(w_{t}:=\beta _{t} e_{1}\), where \(\beta _{t}:=\frac{t}{\lambda _{1}- \lambda }\), is the unique solution of

$$\begin{aligned} \left\{ \begin{array}{ll} (-\Delta )^{s} w_{t}= \lambda w_{t} + t e_{1} &{} \text{ in } \Omega ,\\ w_{t}=0 &{} \text{ in } \mathbb {R}^{N}\setminus \Omega . \end{array} \right. \end{aligned}$$
(2.1)

Set \(u_{t}:=u_{0}+\beta _{t} e_{1}\). Since \(\lambda >\lambda _{1}\), we get \(u_{t}\le 0\) for \(t>0\) sufficiently large. This implies that (i) of Theorem 1 holds. Let us observe that, if \(\lambda >\lambda _{1}\) is an eigenvalue of \(((-\Delta )^{s}, \mathbb {H}_{0}^{s}(\Omega ))\) and \(h\in \ker ((-\Delta )^{s}-\lambda )^{\perp }\), then we can show that there exists \(t_{0}= t_{0}(h)>0\) such that, if \(t>t_{0}\), (1.1) has a nonpositive solution \(u_{t}\in \mathbb {H}_{0}^{s}(\Omega )\).

From now on, we focus on (ii) of Theorem 1 by finding a second solution to (1.1) of the form \(u=v+u_{t}\), where v solves

$$\begin{aligned} \left\{ \begin{array}{ll} (-\Delta )^{s} v= \lambda v + (v+u_{t})_{+}^{2^{*}_{s}-1} &{} \text{ in } \Omega ,\\ v=0 &{} \text{ in } \mathbb {R}^{N}\setminus \Omega . \end{array} \right. \end{aligned}$$
(2.2)

To accomplish this purpose, we seek a critical point of the Euler-Lagrange functional \(\mathcal {J}: \mathbb {H}_{0}^{s}(\Omega )\rightarrow \mathbb {R}\) associated with (2.2), that is

$$\begin{aligned} \mathcal {J}(v):= \frac{1}{2}\Vert v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} - \frac{\lambda }{2} \Vert v\Vert ^{2}_{L^{2}(\Omega )} - \frac{1}{2^{*}_{s}} \Vert (v+u_{t})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )}. \end{aligned}$$

We next verify that \(\mathcal {J}\) fulfills all geometric assumptions of the linking theorem (see [25, Theorem 5.3]). Since \(\mathbb {H}_{0}^{s}(\Omega )\) is a Hilbert space, we can consider the direct sum decomposition \(\mathbb {H}_{0}^{s}(\Omega )=\mathbb {H}^{-}\oplus \mathbb {H}^{+}\), where

$$\begin{aligned} \mathbb {H}^{-}:=\mathrm{span} \{e_{1}, \dots , e_{k}\}, \end{aligned}$$

and

$$\begin{aligned} \mathbb {H}^{+}:=(\mathbb {H}^{-})^{\perp }=\{u\in \mathbb {H}_{0}^{s}(\Omega ): \langle u, e_{j}\rangle _{\mathbb {H}_{0}^{s}(\Omega )}=0 \, \text{ for } \text{ all } j=1,\dots ,k\}. \end{aligned}$$

Then it holds (see [24, Proposition 3.1 and Proposition 3.2])

$$\begin{aligned}&\Vert v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} \ge \lambda _{k+1} \Vert v\Vert ^{2}_{L^{2}(\Omega )} \quad \text{ for } \text{ all } v\in \mathbb {H}^{+}, \end{aligned}$$
(2.3)
$$\begin{aligned}&\Vert v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} \le \lambda _{k} \Vert v\Vert ^{2}_{L^{2}(\Omega )} \quad \text{ for } \text{ all } v\in \mathbb {H}^{-}. \end{aligned}$$
(2.4)

Let \(S_{\varrho }:=\partial B_{\varrho }\cap \mathbb {H}^{+}\) and \(Q:=\{v\in \mathbb {H}_{0}^{s}(\Omega ): v= w+ \zeta \sigma , \, w\in \mathbb {H}^{-}, \, \Vert w\Vert _{\mathbb {H}_{0}^{s}(\Omega )}\le r, \, 0\le \zeta \le R\}\), where \(\sigma \in \mathbb {H}^{+}\), \(0<\varrho <R\) and \(r>0\) will be chosen later so that the following conditions hold:

$$\begin{aligned}&\inf _{v\in S_{\varrho }}\mathcal {J}(v)\ge \alpha >0, \end{aligned}$$
(2.5)
$$\begin{aligned}&\max _{v\in \partial Q}\mathcal {J}(v)\le \alpha _{0}, \text{ with } \alpha _{0}< \alpha , \end{aligned}$$
(2.6)
$$\begin{aligned}&\max _{v\in Q} \mathcal {J}(v)<\frac{s}{N}S_{*}^{\frac{N}{2s}}. \end{aligned}$$
(2.7)

Lemma 1

There exist \(\varrho >0\) and \(\alpha >0\) such that

$$\begin{aligned} \mathcal {J}(v)\ge \alpha \quad \text{ for } \text{ all } v\in S_{\varrho }. \end{aligned}$$

Proof

Let \(v\in \mathbb {H}^{+}\). Since \(u_{t}\le 0\) and \(\lambda \in (\lambda _{k}, \lambda _{k+1})\), it follows from (2.3) and the Sobolev inequality [13, Theorem 1.1] that

$$\begin{aligned} \mathcal {J}(v)&\ge \frac{1}{2}\Vert v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} - \frac{\lambda }{2} \Vert v\Vert ^{2}_{L^{2}(\Omega )} - \frac{1}{2^{*}_{s}} \Vert v_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )}\\&\ge \frac{1}{2} \left( 1- \frac{\lambda }{\lambda _{k+1}}\right) \Vert v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} - \frac{1}{2^{*}_{s}} \Vert v_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )}\\&\ge \frac{1}{2} \left( 1- \frac{\lambda }{\lambda _{k+1}}\right) \Vert v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} - \frac{1}{2^{*}_{s}}S_{*}^{-\frac{N}{N-2s}}\Vert v\Vert ^{2^{*}_{s}}_{\mathbb {H}_{0}^{s}(\Omega )}. \end{aligned}$$

Define

$$\begin{aligned} \omega (\zeta ):=\frac{1}{2} \left( 1- \frac{\lambda }{\lambda _{k+1}}\right) \zeta ^{2} - \frac{1}{2^{*}_{s}}S_{*}^{-\frac{N}{N-2s}} \zeta ^{2^{*}_{s}}, \quad \zeta \ge 0. \end{aligned}$$

Observing that

$$\begin{aligned} \omega '(\zeta )= \zeta \left[ \left( 1- \frac{\lambda }{\lambda _{k+1}}\right) - S_{*}^{-\frac{N}{N-2s}} \zeta ^{2^{*}_{s}-2}\right] , \end{aligned}$$

we deduce that \(\omega (\zeta )\) achieve its maximum at \(\zeta = \varrho\) given by

$$\begin{aligned} \varrho :=\left( 1- \frac{\lambda }{\lambda _{k+1}}\right) ^{\frac{N-2s}{4s}}S_{*}^{\frac{N}{4s}}, \end{aligned}$$

and that

$$\begin{aligned} \alpha := \omega (\varrho )&= \frac{1}{2} \left( 1- \frac{\lambda }{\lambda _{k+1}}\right) ^{\frac{N}{2s}}S_{*}^{\frac{N}{2s}}- \frac{1}{2^{*}_{s}} \left( 1- \frac{\lambda }{\lambda _{k+1}}\right) ^{\frac{N}{2s}}S_{*}^{\frac{N}{2s}} \nonumber \\&=\frac{s}{N}S_{*}^{\frac{N}{2s}} \left( 1- \frac{\lambda }{\lambda _{k+1}}\right) ^{\frac{N}{2s}}. \end{aligned}$$
(2.8)

Therefore, \(\mathcal {J}(v)\ge \alpha\) for all \(v\in S_{\varrho }\). The proof of the lemma is complete.

Let \(\eta \in C_{c}^{\infty }(\mathbb {R}^{N})\) be a cut-off function such that \(0\le \eta \le 1\) in \(\mathbb {R}^{N}\), \(\eta =1\) in \(B_{\varsigma }\) and \(\eta =0\) in \(\mathbb {R}^{N}\setminus B_{2\varsigma }\), where \(\varsigma >0\) is such that \(B_{4\varsigma }\subset \Omega\). It is well known (see [13, Theorem 1.1]) that \(S_{*}\) is achieved by

$$\begin{aligned} \tilde{u}(x):= \kappa (\mu ^{2}+ |x-x_{0}|^{2})^{-\frac{N-2s}{2}}, \end{aligned}$$

with \(\kappa \in \mathbb {R}\setminus \{0\}\), \(\mu >0\) and \(x_{0}\in \mathbb {R}^{N}\) fixed constants. As in [27], for every \({{\,\mathrm{\varepsilon }\,}}>0\) we define

$$\begin{aligned}&u^{*}(x):= \tilde{u}\left( \frac{x}{S_{*}^{\frac{1}{2s}}} \right) \frac{1}{\Vert \tilde{u}\Vert _{L^{2^{*}_{s}}(\mathbb {R}^{N})}}, \\&U_{{{\,\mathrm{\varepsilon }\,}}}(x):= {{\,\mathrm{\varepsilon }\,}}^{-\frac{N-2s}{2}} u^{*}\left( \frac{x}{{{\,\mathrm{\varepsilon }\,}}}\right) , \end{aligned}$$

and set

$$\begin{aligned} u_{{{\,\mathrm{\varepsilon }\,}}}(x):= \eta (x) U_{{{\,\mathrm{\varepsilon }\,}}}(x). \end{aligned}$$

We recall the following useful estimates (see [27, Proposition 21 and Proposition 22] and [24, Proposition 15.6]).

Lemma 2

Let \(s\in (0, 1)\) and \(N>2s\). Then we have:

$$\begin{aligned}&{[}u_{{{\,\mathrm{\varepsilon }\,}}}]^{2}_{s} \le S_{*}^{\frac{N}{2s}} + O({{\,\mathrm{\varepsilon }\,}}^{N-2s}), \end{aligned}$$
(2.9)
$$\begin{aligned}&\Vert u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\mathbb {R}^{N})} = S_{*}^{\frac{N}{2s}} + O({{\,\mathrm{\varepsilon }\,}}^{N}) , \end{aligned}$$
(2.10)
$$\begin{aligned}&\Vert u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{L^{2}(\mathbb {R}^{N})} \ge \left\{ \begin{array}{ll} C{{\,\mathrm{\varepsilon }\,}}^{2s} + O({{\,\mathrm{\varepsilon }\,}}^{N-2s}) &{} \text{ if } N>4s,\\ C{{\,\mathrm{\varepsilon }\,}}^{2s} |\log {{\,\mathrm{\varepsilon }\,}}|+ O({{\,\mathrm{\varepsilon }\,}}^{2s}) &{} \text{ if } N=4s,\\ C{{\,\mathrm{\varepsilon }\,}}^{N-2s}+O({{\,\mathrm{\varepsilon }\,}}^{2s}) &{} \text{ if } 2s<N<4s, \end{array} \right. \end{aligned}$$
(2.11)
$$\begin{aligned}&\Vert u_{{{\,\mathrm{\varepsilon }\,}}}\Vert _{L^{1}(\mathbb {R}^{N})} =O( {{\,\mathrm{\varepsilon }\,}}^{\frac{N-2s}{2}}), \end{aligned}$$
(2.12)
$$\begin{aligned}&\Vert u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}-1}_{L^{2^{*}_{s}-1}(\mathbb {R}^{N})} = O({{\,\mathrm{\varepsilon }\,}}^{\frac{N-2s}{2}}). \end{aligned}$$
(2.13)

Let \(P_{-}: \mathbb {H}_{0}^{s}(\Omega )\rightarrow \mathbb {H}^{-}\) and \(P_{+}: \mathbb {H}_{0}^{s}(\Omega )\rightarrow \mathbb {H}^{+}\) be the orthogonal projections of \(\mathbb {H}_{0}^{s}(\Omega )\) onto \(\mathbb {H}^{-}\) and \(\mathbb {H}^{+}\), respectively. Then, we can prove the next lemma.

Lemma 3

Let \(s\in (0, 1)\) and \(N>2s\). Then the following estimates hold:

$$\begin{aligned}&\Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} \le [u_{{{\,\mathrm{\varepsilon }\,}}}]_{s}^{2}\le S_{*}^{\frac{N}{2s}} + O({{\,\mathrm{\varepsilon }\,}}^{N-2s}), \end{aligned}$$
(2.14)
$$\begin{aligned}&\left| \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} - \Vert u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} \right| \le C {{\,\mathrm{\varepsilon }\,}}^{N-2s}, \end{aligned}$$
(2.15)
$$\begin{aligned}&\Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}-1}_{L^{2^{*}_{s}-1}(\Omega )}\le C{{\,\mathrm{\varepsilon }\,}}^{\frac{N-2s}{2}}, \end{aligned}$$
(2.16)
$$\begin{aligned}&\Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert _{L^{1}(\Omega )} \le C{{\,\mathrm{\varepsilon }\,}}^{\frac{N-2s}{2}}, \end{aligned}$$
(2.17)
$$\begin{aligned}&|P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}(x)|\le C {{\,\mathrm{\varepsilon }\,}}^{\frac{N-2s}{2}} \, \text{ for } x\in \Omega . \end{aligned}$$
(2.18)

Proof

We argue as in [11, Lemma 6]. We write \(u_{{{\,\mathrm{\varepsilon }\,}}}= P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}+P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\) and

$$\begin{aligned} P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}= \sum _{j=1}^{k} c_{j} e_{j}, \, \text{ with } c_{j}:=\int _{\Omega } u_{{{\,\mathrm{\varepsilon }\,}}}(x) e_{j}(x) \, dx , \, j=1, \dots , k. \end{aligned}$$

Then, by using (2.12), we can infer

$$\begin{aligned} \Vert P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{L^{2}(\Omega )}&= \sum _{j=1}^{k} c_{j}^{2} = \sum _{j=1}^{k} \left( \int _{\Omega } u_{{{\,\mathrm{\varepsilon }\,}}} e_{j} \, dx \right) ^{2}\nonumber \\&\le \left( \sum _{j=1}^{k} \Vert e_{j}\Vert ^{2}_{L^{\infty }(\Omega )}\right) \Vert u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{L^{1}(\Omega )} \nonumber \\&\le C {{\,\mathrm{\varepsilon }\,}}^{N-2s}. \end{aligned}$$
(2.19)

Analogously, we can prove that, for some constant \(C>0\),

$$\begin{aligned} \Vert P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert _{L^{\infty }(\Omega )} \le C {{\,\mathrm{\varepsilon }\,}}^{\frac{N-2s}{2}}, \end{aligned}$$
(2.20)

that is (2.18) holds. Now, exploiting (2.19), (2.20) and (2.13), we have

$$\begin{aligned} \left| \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} - \Vert u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} \right|&=\left| \int _{\Omega } \left( \int _{0}^{1} \frac{d}{d\tau } |u_{{{\,\mathrm{\varepsilon }\,}}}- \tau P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}|^{2^{*}_{s}} \, d\tau \right) \, dx \right| \nonumber \\&\le 2^{*}_{s} \int _{0}^{1} d\tau \, \int _{\Omega } |u_{{{\,\mathrm{\varepsilon }\,}}}- \tau P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}|^{2^{*}_{s}-2}|u_{{{\,\mathrm{\varepsilon }\,}}}- \tau P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}| |P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}| \, dx \nonumber \\&\le 2^{*}_{s} \, 2^{2^{*}_{s}-1} \int _{0}^{1} d\tau \, \int _{\Omega } \left( |u_{{{\,\mathrm{\varepsilon }\,}}}|^{2^{*}_{s}-1} + \tau ^{2^{*}_{s}-1} |P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}|^{2^{*}_{s}-1}\right) |P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}|\, dx \nonumber \\&\le C \left[ \Vert u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}-1}_{L^{2^{*}_{s}-1}(\Omega )} \Vert P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert _{L^{\infty }(\Omega )} + \Vert P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}}_{L^{2}(\Omega )} \right] \nonumber \\&\le C {{\,\mathrm{\varepsilon }\,}}^{N-2s}, \end{aligned}$$

and thus (2.15) is satisfied. On the other hand, from (2.12), we see that

$$\begin{aligned} \Vert P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert _{L^{1}(\Omega )}&\le \sum _{j=1}^{k} |c_{j}| \Vert e_{j}\Vert _{L^{1}(\Omega )} \nonumber \\&\le \left( \max _{1\le j \le k} \Vert e_{j}\Vert _{L^{1}(\Omega )} \right) \sum _{j=1}^{k} \Vert u_{{{\,\mathrm{\varepsilon }\,}}} e_{j}\Vert _{L^{1}(\Omega )} \nonumber \\&\le \left( \max _{1\le j\le k} \Vert e_{j}\Vert _{L^{1}(\Omega )} \right) \sum _{j=1}^{k} \Vert e_{j}\Vert _{L^{\infty }(\Omega )} \Vert u_{{{\,\mathrm{\varepsilon }\,}}}\Vert _{L^{1}(\Omega )} \le C {{\,\mathrm{\varepsilon }\,}}^{\frac{N-2s}{2}}. \end{aligned}$$
(2.21)

Using \(u_{{{\,\mathrm{\varepsilon }\,}}}= P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}+P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\), \((a+b)^{p}\le 2^{p-1}(a^{p}+b^{p})\) for \(a, b\ge 0\) and \(p\ge 1\), and relations (2.13), (2.20), (2.21), we obtain

$$\begin{aligned} \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}-1}_{L^{2^{*}_{s}-1}(\Omega )}&\le 2^{2^{*}_{s}-1} \left( \Vert u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}-1}_{L^{2^{*}_{s}-1}(\Omega )} + \Vert P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}-1}_{L^{2^{*}_{s}-1}(\Omega )}\right) \\&\le 2^{2^{*}_{s}-1} \left( C {{\,\mathrm{\varepsilon }\,}}^{\frac{N-2s}{2}} + \Vert P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert _{L^{\infty }(\Omega )}^{\frac{4s}{N-2s}} \Vert P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert _{L^{1}(\Omega )} \right) \\&\le 2^{2^{*}_{s}-1} \left( C {{\,\mathrm{\varepsilon }\,}}^{\frac{N-2s}{2}} + C{{\,\mathrm{\varepsilon }\,}}^{2s} {{\,\mathrm{\varepsilon }\,}}^{\frac{N-2s}{2}}\right) \\&\le C {{\,\mathrm{\varepsilon }\,}}^{\frac{N-2s}{2}}. \end{aligned}$$

Hence, (2.16) is verified. Combining (2.12) and (2.21), we get (2.17). Finally, the estimate (2.14) is a consequence of (2.9). This completes the proof of the lemma.

Fix \(K>0\) and define the set \(\Omega _{{{\,\mathrm{\varepsilon }\,}}, K}:= \{x\in \Omega : P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}(x)>K\}\). By (2.18) we can deduce

$$\begin{aligned} P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}(0)= u_{{{\,\mathrm{\varepsilon }\,}}}(0)- P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}(0)\ge \frac{C_{0}}{\Vert \tilde{u}\Vert _{L^{2^{*}_{s}}(\mathbb {R}^{N})}}{{\,\mathrm{\varepsilon }\,}}^{-\frac{N-2s}{2}}- \Vert P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert _{L^{\infty }(\Omega )} \ge C{{\,\mathrm{\varepsilon }\,}}^{-\frac{N-2s}{2}}, \end{aligned}$$

which implies that \(P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}(0)\rightarrow \infty\) as \({{\,\mathrm{\varepsilon }\,}}\rightarrow 0\). By the continuity of \(P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\), there exists \(\nu >0\) such that \(B_{\nu }\subset \Omega _{{{\,\mathrm{\varepsilon }\,}}, K}\). Therefore, we have the result below.

Lemma 4

Let \(s\in (0, 1)\) and \(N>2s\). Then the following estimates hold:

$$\begin{aligned}&\Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega _{{{\,\mathrm{\varepsilon }\,}}, K})} = \Vert u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} + O({{\,\mathrm{\varepsilon }\,}}^{N-2s}), \\&\Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}-1}_{L^{2^{*}_{s}-1}(\Omega _{{{\,\mathrm{\varepsilon }\,}}, K})} = \Vert u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}-1}_{L^{2^{*}_{s}-1}(\Omega )} + O({{\,\mathrm{\varepsilon }\,}}^{\frac{N+2s}{2}}), \\&\Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert _{L^{1}(\Omega _{{{\,\mathrm{\varepsilon }\,}}, K})} = \Vert u_{{{\,\mathrm{\varepsilon }\,}}}\Vert _{L^{1}(\Omega )} + O({{\,\mathrm{\varepsilon }\,}}^{N}). \end{aligned}$$

From the fundamental theorem of calculus, we derive the following lemma (see [16, Lemma 2.5]).

Lemma 5

Let \(u, v \in L^{p}(\Omega )\) with \(p\in [2, 2^{*}_{s}]\). If \(\tilde{\Omega }\subset \Omega\) and \(u+v>0\) in \(\tilde{\Omega }\), then there exists a constant \(C>0\), depending on p, such that

$$\begin{aligned} \left| \Vert u+v\Vert ^{p}_{L^{p}(\tilde{\Omega })} - \Vert u\Vert ^{p}_{L^{p}(\tilde{\Omega })} -\Vert v\Vert ^{p}_{L^{p}(\tilde{\Omega })} \right| \le C \int _{\tilde{\Omega }} (|u|^{p-1} |v| + |u| |v|^{p-1})\, dx. \end{aligned}$$

Lemma 6

Let \(A, B, C, \nu >0\), and let

$$\begin{aligned} G_{{{\,\mathrm{\varepsilon }\,}}}(\zeta ):= \frac{1}{2} \zeta ^{2}A - \frac{1}{2^{*}_{s}} \zeta ^{2^{*}_{s}} B + \zeta ^{2^{*}_{s}}{{\,\mathrm{\varepsilon }\,}}^{\nu } C, \, \text{ for } \zeta >0. \end{aligned}$$

Then, \(G_{{{\,\mathrm{\varepsilon }\,}}}\) achieve its maximum at

$$\begin{aligned} \zeta _{{{\,\mathrm{\varepsilon }\,}}}:= \left( \frac{A}{B- 2^{*}_{s}{{\,\mathrm{\varepsilon }\,}}^{\nu } C} \right) ^{\frac{1}{2^{*}_{s}-2}}, \end{aligned}$$

and it holds

$$\begin{aligned} G_{{{\,\mathrm{\varepsilon }\,}}}(\zeta )\le G_{{{\,\mathrm{\varepsilon }\,}}}(\zeta _{{{\,\mathrm{\varepsilon }\,}}})= \frac{s}{N}\frac{A^{\frac{N}{2s}}}{B^{\frac{N-2s}{2s}}}+ O({{\,\mathrm{\varepsilon }\,}}^{\nu }). \end{aligned}$$

Proof

It is easy to prove that \(G_{{{\,\mathrm{\varepsilon }\,}}}\) achieves its maximum at \(\zeta _{{{\,\mathrm{\varepsilon }\,}}}\) and thus \(G'_{{{\,\mathrm{\varepsilon }\,}}}(\zeta _{{{\,\mathrm{\varepsilon }\,}}})=0\), that is

$$\begin{aligned} \zeta _{{{\,\mathrm{\varepsilon }\,}}}A- \zeta _{{{\,\mathrm{\varepsilon }\,}}}^{2^{*}_{s}-1}B+ 2^{*}_{s}C{{\,\mathrm{\varepsilon }\,}}^{\nu } \zeta _{{{\,\mathrm{\varepsilon }\,}}}^{2^{*}_{s}-1}=0, \end{aligned}$$
(2.22)

which yields

$$\begin{aligned} \zeta _{{{\,\mathrm{\varepsilon }\,}}}= \left( \frac{A}{B- 2^{*}_{s}{{\,\mathrm{\varepsilon }\,}}^{\nu } C} \right) ^{\frac{1}{2^{*}_{s}-2}}\ge \left( \frac{A}{B} \right) ^{\frac{1}{2^{*}_{s}-2}}=:\zeta _{0}. \end{aligned}$$

Let \(\beta _{{{\,\mathrm{\varepsilon }\,}}}:=\frac{\zeta _{{{\,\mathrm{\varepsilon }\,}}}- \zeta _{0}}{\zeta _{0}}\). Then \(\zeta _{{{\,\mathrm{\varepsilon }\,}}}\rightarrow \zeta _{0}\) and \(\beta _{{{\,\mathrm{\varepsilon }\,}}}\rightarrow 0\) as \({{\,\mathrm{\varepsilon }\,}}\rightarrow 0\). Moreover, by using (2.22) and \(\zeta _{{{\,\mathrm{\varepsilon }\,}}}= (1+ \beta _{{{\,\mathrm{\varepsilon }\,}}}) \zeta _{0}\), we get

$$\begin{aligned} (1+ \beta _{{{\,\mathrm{\varepsilon }\,}}}) \zeta _{0} A - (1+\beta _{{{\,\mathrm{\varepsilon }\,}}})^{2^{*}_{s}-1} \zeta _{0}^{2^{*}_{s}-1} B + 2^{*}_{s} C {{\,\mathrm{\varepsilon }\,}}^{\nu }(1+ \beta _{{{\,\mathrm{\varepsilon }\,}}})^{2^{*}_{s}-1} \zeta _{0}^{2^{*}_{s}-1}=0. \end{aligned}$$
(2.23)

Since

$$\begin{aligned} \zeta _{0}A= \zeta _{0}^{2^{*}_{s}-1}B= \left( \frac{A^{2^{*}_{s}-1}}{B}\right) ^{\frac{1}{2^{*}_{s}-2}}, \end{aligned}$$

we can see that (2.23) becomes

$$\begin{aligned} \left( \frac{A^{2^{*}_{s}-1}}{B}\right) ^{\frac{1}{2^{*}_{s}-2}} \left[ (1+ \beta _{{{\,\mathrm{\varepsilon }\,}}}) - (1+ \beta _{{{\,\mathrm{\varepsilon }\,}}})^{2^{*}_{s}-1}\right] + 2^{*}_{s} C {{\,\mathrm{\varepsilon }\,}}^{\nu }(1+ \beta _{{{\,\mathrm{\varepsilon }\,}}})^{2^{*}_{s}-1} \zeta _{0}^{2^{*}_{s}-1}=0. \end{aligned}$$

Recalling that \((1+x)^{\alpha }=1+ \alpha x + o(x)\) as \(x\rightarrow 0\), we deduce that

$$\begin{aligned} \left( \frac{A^{2^{*}_{s}-1}}{B}\right) ^{\frac{1}{2^{*}_{s}-2}} \left[ \frac{4s}{N-2s} \beta _{{{\,\mathrm{\varepsilon }\,}}} + o(\beta _{{{\,\mathrm{\varepsilon }\,}}})\right] = 2^{*}_{s} C {{\,\mathrm{\varepsilon }\,}}^{\nu }(1+ \beta _{{{\,\mathrm{\varepsilon }\,}}})^{2^{*}_{s}-1} \zeta _{0}^{2^{*}_{s}-1} \end{aligned}$$

which combined with \(\beta _{{{\,\mathrm{\varepsilon }\,}}}\rightarrow 0\) as \({{\,\mathrm{\varepsilon }\,}}\rightarrow 0\), implies that \(\beta _{{{\,\mathrm{\varepsilon }\,}}}=O({{\,\mathrm{\varepsilon }\,}}^{\nu })\). Finally, as \({{\,\mathrm{\varepsilon }\,}}\rightarrow 0\),

$$\begin{aligned} G_{{{\,\mathrm{\varepsilon }\,}}}(\zeta )&\le G_{{{\,\mathrm{\varepsilon }\,}}}(\zeta _{{{\,\mathrm{\varepsilon }\,}}}) = G_{{{\,\mathrm{\varepsilon }\,}}}(\zeta _{0} (1+ \beta _{{{\,\mathrm{\varepsilon }\,}}})) \\&= \frac{1}{2} \zeta _{0}^{2}(1+ \beta _{{{\,\mathrm{\varepsilon }\,}}})^{2}A - \frac{1}{2^{*}_{s}} \zeta _{0}^{2^{*}_{s}}(1+ \beta _{{{\,\mathrm{\varepsilon }\,}}})^{2^{*}_{s}} B + \zeta _{0}^{2^{*}_{s}}(1+ \beta _{{{\,\mathrm{\varepsilon }\,}}})^{2^{*}_{s}}{{\,\mathrm{\varepsilon }\,}}^{\nu } C \\&= \frac{s}{N} \frac{A^{\frac{N}{2s}}}{B^{\frac{N-2s}{2s}}} + O({{\,\mathrm{\varepsilon }\,}}^{\nu }). \end{aligned}$$

The proof of Lemma 6 is now complete.

Our aim is to select Q and \(\rho\) so that (2.5), (2.6) and (2.7) are satisfied. We also pick \(\sigma\) as function of \({{\,\mathrm{\varepsilon }\,}}\), namely \(\sigma :=P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\).

Lemma 7

There exist \(r_{0}>0\), \(R_{0}>0\) and \({{\,\mathrm{\varepsilon }\,}}_{0}>0\) such that, for \(r> r_{0}\), \(R> R_{0}\) and \({{\,\mathrm{\varepsilon }\,}}\in (0, {{\,\mathrm{\varepsilon }\,}}_{0})\), we have

$$\begin{aligned} \max _{v\in \partial Q}\mathcal {J}(v)\le \alpha _{0}, \end{aligned}$$

for some \(\alpha _{0}<\alpha\), where \(\alpha >0\) is given in (2.8).

Proof

Let us observe that \(\partial Q= \mathcal {Q}_{1}\cup \mathcal {Q}_{2} \cup \mathcal {Q}_{3}\), where

$$\begin{aligned}&\mathcal {Q}_{1}:= \bar{B}_{r}\cap \mathbb {H}^{-}, \\&\mathcal {Q}_{2}:= \{v\in \mathbb {H}_{0}^{s}(\Omega ): v= w+ \zeta P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}, \, w\in \mathbb {H}^{-}, \, \Vert w\Vert _{\mathbb {H}_{0}^{s}(\Omega )}=r, \, 0< \zeta < R \}, \\&\mathcal {Q}_{3}:= \{v \in \mathbb {H}_{0}^{s}(\Omega ): v= w+ R P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}, \, w\in \bar{B}_{r}\cap \mathbb {H}^{-}\}. \end{aligned}$$

We will study \(\mathcal {J}|_{\mathcal {Q}_{i}}\) for \(i=1,2,3\). We recall that \(\lambda \in (\lambda _{k}, \lambda _{k+1})\). For \(v\in \mathcal {Q}_{1}\), taking into account (2.4), we can see that

$$\begin{aligned} \mathcal {J}(v)\le \frac{1}{2} \left( 1- \frac{\lambda }{\lambda _{k}}\right) \Vert v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} - \frac{1}{2^{*}_{s}} \Vert (v+u_{t})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )}\le 0. \end{aligned}$$

Let \(v\in \mathcal {Q}_{2}\). Define

$$\begin{aligned} \delta ^{2}:=\sup _{0<{{\,\mathrm{\varepsilon }\,}}\le 1} \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )}, \end{aligned}$$
(2.24)

and take

$$\begin{aligned} \tau _{0}\in \left( 0, \frac{\sqrt{2\alpha }}{\delta }\right) . \end{aligned}$$
(2.25)

Fix \(0< \zeta \le \tau _{0}\). From (2.4), (2.24) and (2.25), we deduce that

$$\begin{aligned} \mathcal {J}(v)&\le \frac{1}{2}\left( 1- \frac{\lambda }{\lambda _{k}}\right) r^{2}+ \frac{\zeta ^{2}}{2} \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} - \frac{1}{2^{*}_{s}} \Vert (w+ \zeta P_{+}u_{{{\,\mathrm{\varepsilon }\,}}} + u_{t})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )}\\&\le \frac{\zeta ^{2}}{2} \delta ^{2} \le \frac{\tau ^{2}_{0}}{2}\delta ^{2}<\alpha . \end{aligned}$$

Now, let \(\zeta > \tau _{0}\). Put

$$\begin{aligned} L:= \sup \left\{ \left\| \frac{w+u_{t}}{\xi } \right\| _{L^{\infty }(\Omega )} : \tau _{0}< \xi < R, \, \Vert w\Vert _{\mathbb {H}_{0}^{s}(\Omega )}= r, \, w \in \mathbb {H}^{-} \right\} . \end{aligned}$$

Since \(P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}(0)\rightarrow \infty\) as \({{\,\mathrm{\varepsilon }\,}}\rightarrow 0\), there exists \({{\,\mathrm{\varepsilon }\,}}'_{0}>0\) such that, for all \(0<{{\,\mathrm{\varepsilon }\,}}<{{\,\mathrm{\varepsilon }\,}}'_{0}\),

$$\begin{aligned} \Omega _{{{\,\mathrm{\varepsilon }\,}}}:=\Omega _{{{\,\mathrm{\varepsilon }\,}}, L}= \{x\in \Omega : P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}(x)>L\}\ne \emptyset . \end{aligned}$$

Applying Lemma 5 with \(u=P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\), \(v=\frac{w+u_{t}}{\zeta }\), \(p=2^{*}_{s}\) and \(\tilde{\Omega }= \Omega _{{{\,\mathrm{\varepsilon }\,}}}\), we get

$$\begin{aligned} \left\| \left( P_{+}u_{{{\,\mathrm{\varepsilon }\,}}} + \frac{w+u_{t}}{\zeta }\right) _{+}\right\| ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )}&\ge \left\| \left( P_{+}u_{{{\,\mathrm{\varepsilon }\,}}} + \frac{w+u_{t}}{\zeta }\right) _{+}\right\| ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega _{{{\,\mathrm{\varepsilon }\,}}})} \nonumber \\&\ge \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega _{{{\,\mathrm{\varepsilon }\,}}})} + \left\| \frac{w+u_{t}}{\zeta }\right\| ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega _{{{\,\mathrm{\varepsilon }\,}}})} \nonumber \\&\quad - C \int _{\Omega _{{{\,\mathrm{\varepsilon }\,}}}} \left[ |P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}|^{2^{*}_{s}-1}\left| \frac{w+u_{t}}{\zeta }\right| + |P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}| \left| \frac{w+u_{t}}{\zeta }\right| ^{2^{*}_{s}-1} \right] dx \nonumber \\&\ge \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega _{{{\,\mathrm{\varepsilon }\,}}})} + \left\| \frac{w+u_{t}}{\zeta }\right\| ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega _{{{\,\mathrm{\varepsilon }\,}}})} \nonumber \\&\quad -C\left( \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}-1}_{L^{2^{*}_{s}-1}(\Omega _{{{\,\mathrm{\varepsilon }\,}}})} + \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert _{L^{1}(\Omega _{{{\,\mathrm{\varepsilon }\,}}})}\right) . \end{aligned}$$
(2.26)

Then, by using Lemmas 2, 3, 4 and (2.26), we can see that, for \({{\,\mathrm{\varepsilon }\,}}>0\) small enough,

$$\begin{aligned} \mathcal {J}(v)&\le \frac{1}{2}\left( 1- \frac{\lambda }{\lambda _{k}}\right) r^{2} + \frac{\zeta ^{2}}{2} \left( \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} -\lambda \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{L^{2}(\Omega )}\right) \\&\quad -\frac{\zeta ^{2^{*}_{s}}}{2^{*}_{s}} \left[ \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega _{{{\,\mathrm{\varepsilon }\,}}})} + \left\| \frac{w+u_{t}}{\zeta }\right\| ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega _{{{\,\mathrm{\varepsilon }\,}}})} - C\left( \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}-1}_{L^{2^{*}_{s}-1}(\Omega _{{{\,\mathrm{\varepsilon }\,}}})} + \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert _{L^{1}(\Omega _{{{\,\mathrm{\varepsilon }\,}}})}\right) \right] \\&\le \frac{1}{2}\left( 1- \frac{\lambda }{\lambda _{k}}\right) r^{2} + \frac{\zeta ^{2}}{2} \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} -\frac{\zeta ^{2^{*}_{s}}}{2^{*}_{s}} \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega _{{{\,\mathrm{\varepsilon }\,}}})} \\&\quad + C \zeta ^{2^{*}_{s}} \left( \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}-1}_{L^{2^{*}_{s}-1}(\Omega _{{{\,\mathrm{\varepsilon }\,}}})} + \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert _{L^{1}(\Omega _{{{\,\mathrm{\varepsilon }\,}}})}\right) \\&\le \frac{1}{2}\left( 1- \frac{\lambda }{\lambda _{k}}\right) r^{2} + \frac{\zeta ^{2}}{2}\left( S_{*}^{\frac{N}{2s}}+O({{\,\mathrm{\varepsilon }\,}}^{N-2s})\right) - \frac{\zeta ^{2^{*}_{s}}}{2^{*}_{s}} \left( S_{*}^{\frac{N}{2s}} +O({{\,\mathrm{\varepsilon }\,}}^{N})+O({{\,\mathrm{\varepsilon }\,}}^{N-2s})\right) \\&\quad +C_{1}\zeta ^{2^{*}_{s}} \left( {{\,\mathrm{\varepsilon }\,}}^{\frac{N-2s}{2}}+{{\,\mathrm{\varepsilon }\,}}^{\frac{N+2s}{2}}+{{\,\mathrm{\varepsilon }\,}}^{N}\right) \\&\le \frac{1}{2}\left( 1- \frac{\lambda }{\lambda _{k}}\right) r^{2} + \frac{\zeta ^{2}}{2} S_{*}^{\frac{N}{2s}} - \frac{\zeta ^{2^{*}_{s}}}{2^{*}_{s}} S_{*}^{\frac{N}{2s}} + C_{2} \zeta ^{2^{*}_{s}} {{\,\mathrm{\varepsilon }\,}}^{\frac{N-2s}{2}}\\&=: \frac{1}{2}\left( 1- \frac{\lambda }{\lambda _{k}}\right) r^{2} + G_{{{\,\mathrm{\varepsilon }\,}}}(\zeta ), \end{aligned}$$

for some \(C_{1}, C_{2}>0\). Invoking Lemma 6, we get

$$\begin{aligned} \mathcal {J}(v)&\le \frac{1}{2}\left( 1- \frac{\lambda }{\lambda _{k}}\right) r^{2} +\frac{s}{N}S_{*}^{\frac{N}{2s}} + O({{\,\mathrm{\varepsilon }\,}}^{\frac{N-2s}{2}}). \end{aligned}$$

Because

$$\begin{aligned} \frac{1}{2}\left( 1- \frac{\lambda }{\lambda _{k}}\right) r^{2}\rightarrow -\infty \text{ as } r\rightarrow \infty , \end{aligned}$$

there exists \(r>0\) such that \(\mathcal {J}(v)<0\) for \(v\in \mathcal {Q}_{2}\). In this way we can obtain an \(r_{0}\) satisfying the thesis.

Now, fix \(v\in \mathcal {Q}_{3}\). Then, \(v= w + R P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\) with \(w\in \bar{B}_{r}\cap \mathbb {H}^{-}\), and it holds

$$\begin{aligned} \mathcal {J}(v)\le \frac{1}{2} \left( 1-\frac{\lambda }{\lambda _{k}}\right) \Vert w\Vert _{\mathbb {H}_{0}^{s}(\Omega )}^{2} + \frac{R^{2}}{2} \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} - \frac{R^{2^{*}_{s}}}{2^{*}_{s}} \left\| \left( P_{+} u_{{{\,\mathrm{\varepsilon }\,}}} + \frac{u_{t}+ w}{R}\right) _{+}\right\| ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )}. \end{aligned}$$

Since \(w\in L^{\infty }(\Omega )\) (note that \(\bar{B}_{r}\cap \mathbb {H}^{-}\) is finite dimensional) and \(u_{t} \in L^{\infty }(\Omega )\) (recall that \(u_{t}=u_{0}+\beta _{t}e_{1}\) and \(u_{0}, e_{1}\in L^{\infty }(\Omega )\)), there exists \(k>0\) such that \(\Vert w+u_{t}\Vert _{L^{\infty }(\Omega )}\le k\). On the other hand, \(P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}(0)\rightarrow \infty\) as \({{\,\mathrm{\varepsilon }\,}}\rightarrow 0\), so we can find \(0<{{\,\mathrm{\varepsilon }\,}}''_{0}<{{\,\mathrm{\varepsilon }\,}}'_{0}\) such that \(P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}(0)>2k\) provided \(0<{{\,\mathrm{\varepsilon }\,}}<{{\,\mathrm{\varepsilon }\,}}''_{0}\). Fix \({{\,\mathrm{\varepsilon }\,}}\in (0, {{\,\mathrm{\varepsilon }\,}}''_{0})\). Then there are \(R_{1}=R_{1}({{\,\mathrm{\varepsilon }\,}})>0\) and \(\eta = \eta ({{\,\mathrm{\varepsilon }\,}})>0\) such that

$$\begin{aligned} \left| \left\{ x\in \Omega \, : \, P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}(x) + \frac{w(x)+u_{t}(x)}{R}>1\right\} \right| \ge \eta>0 \quad \text{ for } \text{ any } R>R_{1}. \end{aligned}$$

Consequently, by (2.14), for \(R>R_{1}\),

$$\begin{aligned} \mathcal {J}(v)&\le \frac{1}{2} \left( 1-\frac{\lambda }{\lambda _{k}}\right) \Vert w\Vert _{\mathbb {H}_{0}^{s}(\Omega )}^{2} + \frac{R^{2}}{2} \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} - \frac{R^{2^{*}_{s}}}{2^{*}_{s}} \eta \\&\le \frac{R^{2}}{2} \left( S_{*}^{\frac{N}{2s}}+O({{\,\mathrm{\varepsilon }\,}}^{N-2s})\right) -\frac{R^{2^{*}_{s}}}{2^{*}_{s}} \eta , \end{aligned}$$

so there exist \({{\,\mathrm{\varepsilon }\,}}_{0}, R_{0}>0\) such that, for \({{\,\mathrm{\varepsilon }\,}}\in (0, {{\,\mathrm{\varepsilon }\,}}_{0})\) and \(R>R_{0}\), we get \(\mathcal {J}(v)\le 0\) for \(v\in \mathcal {Q}_{3}\). The proof of Lemma 7 is now complete.

Lemma 8

Let \(s\in (0, 1)\) and \(N>6s\). Then the following estimate holds:

$$\begin{aligned} \max _{u\in Q} \mathcal {J}(u) < \frac{s}{N} S_{*}^{\frac{N}{2s}}. \end{aligned}$$

Proof

Let \({{\,\mathrm{\varepsilon }\,}}\in (0, {{\,\mathrm{\varepsilon }\,}}_{0})\) be such that the geometry of the linking theorem holds. Take \(w+ \zeta P_{+}u_{{{\,\mathrm{\varepsilon }\,}}} \in Q\). Then,

$$\begin{aligned} \mathcal {J}(w+ \zeta P_{+}u_{{{\,\mathrm{\varepsilon }\,}}})&= \frac{1}{2}\Vert w\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} +\frac{\zeta ^{2}}{2}\Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} -\frac{\lambda }{2} \Vert w\Vert ^{2}_{L^{2}(\Omega )}-\frac{\lambda }{2} \zeta ^{2} \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{L^{2}(\Omega )} \\&\quad - \frac{1}{2^{*}_{s}} \Vert (w+\zeta P_{+}u_{{{\,\mathrm{\varepsilon }\,}}} +u_{t})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )}. \end{aligned}$$

Let \(\delta\) and \(\tau _{0}\) be defined as in (2.24) and (2.25) respectively. Fix \(0< \zeta \le \tau _{0}\). Arguing as in the proof of Lemma 7 and bearing in mind (2.8), we can see that

$$\begin{aligned} \mathcal {J}(w+\zeta P_{+}u_{{{\,\mathrm{\varepsilon }\,}}})\le \frac{\zeta ^{2}}{2} \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )}\le \frac{\tau ^{2}_{0}}{2} \delta ^{2}<\alpha <\frac{s}{N} S_{*}^{\frac{N}{2s}}. \end{aligned}$$

Let \(\zeta > \tau _{0}\). As in the proof of Lemma 7, from (2.26), Lemma 2 and Lemma 4, we derive

$$\begin{aligned} \mathcal {J}(w+ \zeta P_{+}u_{{{\,\mathrm{\varepsilon }\,}}})&\le \frac{\zeta ^{2}}{2} \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )}- \frac{\lambda }{2} \zeta ^{2} \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{L^{2}(\Omega )}\\&\quad - \frac{\zeta ^{2^{*}_{s}}}{2^{*}_{s}} \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega _{{{\,\mathrm{\varepsilon }\,}}})} + C \zeta ^{2^{*}_{s}} {{\,\mathrm{\varepsilon }\,}}^{\frac{N-2s}{2}}=: G_{{{\,\mathrm{\varepsilon }\,}}}(\zeta ). \end{aligned}$$

By Lemma 6 we can infer

$$\begin{aligned} \mathcal {J}(w+ \zeta P_{+}u_{{{\,\mathrm{\varepsilon }\,}}})\le \frac{s}{N} \left( \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )}- \lambda \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{L^{2}(\Omega )} \right) ^{\frac{N}{2s}} \left( \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega _{{{\,\mathrm{\varepsilon }\,}}})}\right) ^{-\frac{N-2s}{2s}} + O({{\,\mathrm{\varepsilon }\,}}^{\frac{N-2s}{2}}). \end{aligned}$$

Using \(N>6s\), the fact that (2.11) and (2.18) imply

$$\begin{aligned} \Vert P_{+}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert _{L^{2}(\Omega )}^{2}=\Vert u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{L^{2} (\Omega )}-\Vert P_{-}u_{{{\,\mathrm{\varepsilon }\,}}}\Vert ^{2}_{L^{2}(\Omega )}\ge C{{\,\mathrm{\varepsilon }\,}}^{2s}+O({{\,\mathrm{\varepsilon }\,}}^{N-2s}), \end{aligned}$$

and applying Lemmas 2, 3 and 4, we have

$$\begin{aligned} \mathcal {J}(w+ \zeta P_{+}u_{{{\,\mathrm{\varepsilon }\,}}})&\le \frac{s}{N} \frac{\left[ S_{*}^{\frac{N}{2s}}+O({{\,\mathrm{\varepsilon }\,}}^{N-2s})- \lambda C {{\,\mathrm{\varepsilon }\,}}^{2s}+O({{\,\mathrm{\varepsilon }\,}}^{N-2s})\right] ^{\frac{N}{2s}}}{\left[ S_{*}^{\frac{N}{2s}}+O({{\,\mathrm{\varepsilon }\,}}^{N}) +O({{\,\mathrm{\varepsilon }\,}}^{N-2s})\right] ^{\frac{N-2s}{2s}}}+O({{\,\mathrm{\varepsilon }\,}}^{\frac{N-2s}{2}})\\&\le \frac{s}{N} S_{*}^{\frac{N}{2s}} - \lambda O({{\,\mathrm{\varepsilon }\,}}^{2s}) + O({{\,\mathrm{\varepsilon }\,}}^{\frac{N-2s}{2}}). \end{aligned}$$

Since \(2s<\frac{N-2s}{2}\) and \(\lambda >0\), the thesis follows by taking \({{\,\mathrm{\varepsilon }\,}}>0\) sufficiently small.

Now we are ready to complete the proof of Theorem 1.

Proof

In view of Lemmas 1, 7 and 8 with \({{\,\mathrm{\varepsilon }\,}}>0\) sufficiently small, it follows from the linking theorem without the Palais–Smale condition (see [23, Theorem 4.3] and [15, Theorem 5.1]) that there exists a sequence \(\{v_{n}\}_{n\in \mathbb {N}}\subset \mathbb {H}_{0}^{s}(\Omega )\) such that

$$\begin{aligned} \mathcal {J}(v_{n})\rightarrow c \quad \text{ and } \quad \mathcal {J}'(v_{n})\rightarrow 0 \text{ in } (\mathbb {H}_{0}^{s}(\Omega ))^{-1}, \end{aligned}$$

that is, for all \(\phi \in \mathbb {H}_{0}^{s}(\Omega )\),

$$\begin{aligned}&\mathcal {J}(v_{n})= \frac{1}{2} \Vert v_{n}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} - \frac{\lambda }{2} \Vert v_{n}\Vert ^{2}_{L^{2}(\Omega )} - \frac{1}{2^{*}_{s}} \Vert (v_{n}+ u_{t})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} = c+o(1), \end{aligned}$$
(2.27)
$$\begin{aligned}&\langle \mathcal {J}'(v_{n}), \phi \rangle = \langle v_{n}, \phi \rangle _{\mathbb {H}_{0}^{s}(\Omega )} - \lambda \int _{\Omega } v_{n}\phi \, dx - \int _{\Omega } (v_{n}+u_{t})_{+}^{2^{*}_{s}-1}\phi \, dx= o(1) \Vert \phi \Vert _{\mathbb {H}_{0}^{s}(\Omega )}, \end{aligned}$$
(2.28)

where c is the linking critical level given by

$$\begin{aligned} c:=\inf _{\gamma \in \Gamma } \sup _{u\in Q} \mathcal {J}(\gamma (u)), \, \text{ with } \Gamma :=\{\gamma \in C^{0}(Q, \mathbb {H}_{0}^{s}(\Omega )): \gamma =\mathrm{Id} \text{ on } \partial Q \}. \end{aligned}$$

Now we prove that \(\{v_{n}\}_{n\in \mathbb {N}}\subset \mathbb {H}_{0}^{s}(\Omega )\) is bounded. Let us observe that, since \(u_{t}\le 0\), we have

$$\begin{aligned} \mathcal {J}(v_{n})-\frac{1}{2}\langle \mathcal {J}'(v_{n}), v_{n} \rangle&=-\frac{1}{2^{*}_{s}} \int _{\Omega } (v_{n}+u_{t})_{+}^{2^{*}_{s}} dx + \frac{1}{2} \int _{\Omega } (v_{n}+u_{t})_{+}^{2^{*}_{s}-1} v_{n} \, dx \nonumber \\&=-\frac{1}{2^{*}_{s}} \int _{\Omega } (v_{n}+u_{t})_{+}^{2^{*}_{s}} dx + \frac{1}{2} \int _{\Omega } (v_{n}+u_{t})_{+}^{2^{*}_{s}-1} (v_{n}+u_{t}-u_{t}) \, dx \nonumber \\&\ge \left( \frac{1}{2} -\frac{1}{2^{*}_{s}}\right) \int _{\Omega } (v_{n}+u_{t})_{+}^{2^{*}_{s}} dx = \frac{s}{N} \Vert (v_{n}+u_{t})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )}. \end{aligned}$$
(2.29)

On the other hand, from (2.27) and (2.28), we obtain

$$\begin{aligned} \mathcal {J}(v_{n})-\frac{1}{2}\langle \mathcal {J}'(v_{n}), v_{n} \rangle \le c+ {{\,\mathrm{\varepsilon }\,}}_{n}\Vert v_{n}\Vert _{\mathbb {H}_{0}^{s}(\Omega )} +o(1). \end{aligned}$$
(2.30)

Putting together (2.29) and (2.30), we get

$$\begin{aligned} \frac{s}{N} \Vert (v_{n}+u_{t})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} \le c+ {{\,\mathrm{\varepsilon }\,}}_{n}\Vert v_{n}\Vert _{\mathbb {H}_{0}^{s}(\Omega )} +o(1). \end{aligned}$$
(2.31)

Write \(v_{n}= v_{n}^{-}+v_{n}^{+}\), with \(v_{n}^{\pm }\in \mathbb {H}^{\pm }\). Then, by using (2.3), we can see that

$$\begin{aligned} \langle \mathcal {J}'(v_{n}), v_{n}^{+}\rangle&=\Vert v_{n}^{+}\Vert _{\mathbb {H}_{0}^{s}(\Omega )}^{2} -\lambda \Vert v_{n}^{+}\Vert ^{2}_{L^{2}(\Omega )} -\int _{\Omega } (v_{n}+u_{t})_{+}^{2^{*}_{s}-1} v_{n}^{+}\, dx \nonumber \\&\ge \left( 1- \frac{\lambda }{\lambda _{k+1}}\right) \Vert v_{n}^{+}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} -\int _{\Omega } (v_{n}+u_{t})_{+}^{2^{*}_{s}-1} v_{n}^{+}\, dx. \end{aligned}$$
(2.32)

From (2.28), (2.31), (2.32), the Hölder inequality, the Young inequality with \(\eta >0\) (that is \(ab\le \eta a^{2}+C_{\eta }b^{2}\) for \(a, b\ge 0\)) and the Sobolev inequality [13, Theorem 1.1], we have

$$\begin{aligned} \left( 1- \frac{\lambda }{\lambda _{k+1}}\right) \Vert v_{n}^{+}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )}&\le \int _{\Omega } (v_{n}+u_{t})_{+}^{2^{*}_{s}-1} v_{n}^{+}\, dx +{{\,\mathrm{\varepsilon }\,}}_{n} \Vert v_{n}^{+}\Vert _{\mathbb {H}_{0}^{s}(\Omega )}\\&\le \Vert v_{n}^{+}\Vert _{L^{2^{*}_{s}}(\Omega )} \Vert (v_{n}+u_{t})_{+}\Vert ^{2^{*}_{s}-1}_{L^{2^{*}_{s}}(\Omega )}+{{\,\mathrm{\varepsilon }\,}}_{n} \Vert v_{n}^{+}\Vert _{\mathbb {H}_{0}^{s}(\Omega )}\\&\le \eta \Vert v_{n}^{+}\Vert ^{2}_{L^{2^{*}_{s}}(\Omega )}+ C_{\eta } \Vert (v_{n}+u_{t})_{+}\Vert _{L^{2^{*}_{s}}(\Omega )}^{2(2^{*}_{s}-1)}+{{\,\mathrm{\varepsilon }\,}}_{n} \Vert v_{n}^{+}\Vert _{\mathbb {H}_{0}^{s}(\Omega )}\\&\le C \eta \Vert v_{n}^{+}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} + C_{\eta } \left( \frac{N}{s}\right) ^{{\frac{2(2^{*}_{s}-1)}{2^{*}_{s}}}} \left( c+ {{\,\mathrm{\varepsilon }\,}}_{n}\Vert v_{n}\Vert _{\mathbb {H}_{0}^{s}(\Omega )} +o(1)\right) ^{\frac{2(2^{*}_{s}-1)}{2^{*}_{s}}}+{{\,\mathrm{\varepsilon }\,}}_{n} \Vert v_{n}^{+}\Vert _{\mathbb {H}_{0}^{s}(\Omega )}\\&\le C \eta \Vert v_{n}^{+}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )}+ C'_{\eta } + C''_{\eta } {{\,\mathrm{\varepsilon }\,}}_{n}\left( \Vert v_{n}\Vert _{\mathbb {H}_{0}^{s}(\Omega )}^{\frac{2(2^{*}_{s}-1)}{2^{*}_{s}}}+ \Vert v_{n}^{+}\Vert _{\mathbb {H}_{0}^{s}(\Omega )}\right) , \end{aligned}$$

and choosing \(0<\eta <\frac{1}{C}\left( 1- \frac{\lambda }{\lambda _{k+1}}\right)\) (we recall that \(\lambda \in (\lambda _{k}, \lambda _{k+1})\)) we have

$$\begin{aligned} \Vert v_{n}^{+}\Vert _{\mathbb {H}_{0}^{s}(\Omega )}\le \bar{C} \left[ 1+ {{\,\mathrm{\varepsilon }\,}}_{n}\left( \Vert v_{n}\Vert _{\mathbb {H}_{0}^{s}(\Omega )}^{\frac{N+2s}{N}}+ \Vert v_{n}^{+}\Vert _{\mathbb {H}_{0}^{s}(\Omega )}\right) \right] . \end{aligned}$$
(2.33)

In a similar fashion, we can prove that

$$\begin{aligned} \Vert v_{n}^{-}\Vert _{\mathbb {H}_{0}^{s}(\Omega )}\le \tilde{C} \left[ 1+ {{\,\mathrm{\varepsilon }\,}}_{n}\left( \Vert v_{n}\Vert _{\mathbb {H}_{0}^{s}(\Omega )}^{\frac{N+2s}{N}}+ \Vert v_{n}^{-}\Vert _{\mathbb {H}_{0}^{s}(\Omega )}\right) \right] . \end{aligned}$$
(2.34)

Then, (2.33) and (2.34) yield \(\Vert v_{n}\Vert _{\mathbb {H}_{0}^{s}(\Omega )}\le C\) for all \(n\in \mathbb {N}\). Up to a subsequence, we may assume that

$$\begin{aligned} \begin{aligned}&v_{n}\rightharpoonup v \quad \text{ in } \mathbb {H}_{0}^{s}(\Omega ), \\&v_{n}\rightarrow v \quad \text{ in } L^{q}(\mathbb {R}^N), \, 1\le q<2^{*}_{s}, \\&v_{n}\rightarrow v \quad \text{ a.e. } \text{ in } \mathbb {R}^N. \end{aligned} \end{aligned}$$
(2.35)

Hence, v is a weak solution to

$$\begin{aligned} \left\{ \begin{array}{ll} (-\Delta )^{s} v= \lambda v +(v+u_{t})_{+}^{2^{*}_{s}-1} &{} \text{ in } \Omega ,\\ v=0 &{} \text{ in } \mathbb {R}^{N}\setminus \Omega , \end{array} \right. \end{aligned}$$
(2.36)

that is, for any \(\phi \in \mathbb {H}_{0}^{s}(\Omega )\) it holds

$$\begin{aligned} \langle v, \phi \rangle _{\mathbb {H}_{0}^{s}(\Omega )} - \lambda \int _{\Omega } v \phi \, dx&= \int _{\Omega } (v+u_{t})_{+}^{2^{*}_{s}-1} \phi \, dx. \end{aligned}$$
(2.37)

In particular, taking \(\phi =v\) in (2.37), we get

$$\begin{aligned} \Vert v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} - \lambda \Vert v\Vert ^{2}_{L^{2}(\Omega )}&= \int _{\Omega } (v+u_{t})_{+}^{2^{*}_{s}-1} v \, dx\nonumber \\&=\int _{\Omega } (v+u_{t})_{+}^{2^{*}_{s}} \, dx - \int _{\Omega } (v+u_{t})_{+}^{2^{*}_{s}-1} u_{t} \, dx. \end{aligned}$$
(2.38)

By applying the Brezis–Lieb lemma [8, Theorem 1], we obtain

$$\begin{aligned}&\Vert v_{n}-v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )}= \Vert v_{n}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )}- \Vert v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )}+o(1), \end{aligned}$$
(2.39)

and

$$\begin{aligned}&\Vert (v_{n}+u_{t})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )}= \Vert (v_{n}-v)_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} + \Vert (v+u_{t})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )}+o(1). \end{aligned}$$
(2.40)

Then, by using (2.35), (2.39) and (2.40), we deduce

$$\begin{aligned} \mathcal {J}(v_{n})= \mathcal {J}(v) + \frac{1}{2} \Vert v_{n}-v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} - \frac{1}{2^{*}_{s}} \Vert (v_{n}-v)_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} +o(1). \end{aligned}$$
(2.41)

On the other hand, by using (2.35), (2.38), (2.39) and (2.40), we have

$$\begin{aligned} \langle \mathcal {J}'(v_{n}), v_{n}\rangle&= \Vert v_{n}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} -\lambda \Vert v_{n}\Vert ^{2}_{L^{2}(\Omega )}- \int _{\Omega } (v_{n}+ u_{t})_{+}^{2^{*}_{s}-1} v_{n} \, dx\\&=\left[ \Vert v_{n}-v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} + \Vert v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} +o(1) \right] -\lambda \left[ \Vert v\Vert ^{2}_{L^{2}(\Omega )}+ o(1)\right] \\&\quad - \Vert (v_{n}+ u_{t})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} + \int _{\Omega } (v_{n}+ u_{t})_{+}^{2^{*}_{s}-1} u_{t} \, dx \\&= \Vert v_{n}-v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} + \left[ \Vert v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} - \lambda \Vert v\Vert ^{2}_{L^{2}(\Omega )}\right] \\&\quad -\Vert (v_{n}-v)_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} - \Vert (v+ u_{t})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} + \int _{\Omega } (v_{n}+ u_{t})_{+}^{2^{*}_{s}-1} u_{t} \, dx +o(1) \\&= \Vert v_{n}-v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )}+ \Vert (v+ u_{t})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} - \int _{\Omega } (v+ u_{t})_{+}^{2^{*}_{s}-1} u_{t} \, dx \\&\quad -\Vert (v_{n}-v)_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} - \Vert (v+ u_{t})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} + \int _{\Omega } (v_{n}+ u_{t})_{+}^{2^{*}_{s}-1} u_{t} \, dx +o(1)\\&= \Vert v_{n}-v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} - \int _{\Omega } (v+ u_{t})_{+}^{2^{*}_{s}-1} u_{t} \, dx -\Vert (v_{n}-v)_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )}+ \int _{\Omega } (v_{n}+ u_{t})_{+}^{2^{*}_{s}-1} u_{t} \, dx +o(1)\\&= \Vert v_{n}-v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} -\Vert (v_{n}-v)_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} + \int _{\Omega } (v_{n}-v)_{+}^{2^{*}_{s}-1} u_{t} \, dx +o(1). \end{aligned}$$

Taking into account that \(\langle \mathcal {J}'(v_{n}), v_{n}\rangle \rightarrow 0\) and \(\int _{\Omega } (v_{n}-v)_{+}^{2^{*}_{s}-1} u_{t}\, dx \rightarrow 0\) as \(n\rightarrow \infty\), we deduce

$$\begin{aligned} \Vert v_{n}-v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )}= \Vert (v_{n}-v)_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} +o(1). \end{aligned}$$
(2.42)

Set \(w_{n}:= v_{n}-v\). Let

$$\begin{aligned} \ell :=\lim _{n\rightarrow \infty } \Vert w_{n}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} \ge 0. \end{aligned}$$

If \(\ell =0\), then \(v_{n}\rightarrow v\) in \(\mathbb {H}_{0}^{s}(\Omega )\) as \(n\rightarrow \infty\) and so \(0<\alpha \le c= \mathcal {J}(v)\), that is v is a nontrivial solution to (2.1). Let \(\ell >0\). Our goal is to prove that \(v\not \equiv 0\). Assume by contradiction that \(v=0\). Then, by the Sobolev inequality [13, Theorem 1.1] and (2.42), we can infer

$$\begin{aligned} \Vert w_{n}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )}&\ge S_{*} \Vert w_{n}\Vert ^{2}_{L^{2^{*}_{s}}(\Omega )} \nonumber \\&\ge S_{*} \Vert (w_{n})_{+}\Vert ^{2}_{L^{2^{*}_{s}}(\Omega )} \nonumber \\&\ge S_{*} \left[ \Vert w_{n}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} + o(1)\right] ^{\frac{2}{2^{*}_{s}}} \end{aligned}$$

which gives

$$\begin{aligned} \ell \ge S_{*} \ell ^{\frac{N-2s}{N}}, \, \text{ i.e. } \ell \ge S_{*}^{\frac{N}{2s}}. \end{aligned}$$
(2.43)

Now, from (2.41), (2.42), (2.43), \(v=0\) and recalling that \(u_{t}\le 0\), we get

$$\begin{aligned} c+o(1)= \mathcal {J}(v_{n})= \mathcal {J}(0)+ \left( \frac{1}{2}- \frac{1}{2^{*}_{s}}\right) \Vert (v_{n})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} + o(1) = \frac{s}{N} \ell \ge \frac{s}{N} S_{*}^{\frac{N}{2s}}, \end{aligned}$$

which contradicts Lemma 8. Therefore, \(v\not \equiv 0\). From (2.36) we also derive that \(v\ge 0\) in \(\mathbb {R}^{N}\).

Remark 1

It is also possible to prove that \(\mathcal {J}\) satisfies the Palais-Smale condition at any level \(d\in \mathbb {R}\) such that \(0<d<\frac{s}{N}S_{*}^{\frac{N}{2s}}\).

Remark 2

When \(0<\lambda <\lambda _{1}\), we can argue as at the beginning of Sect. 2 to prove that there exists \(t_{0}=t_{0}(h)<0\) such that, if \(t<t_{0}\), (1.1) has a nonpositive solution \(u_{t}\in \mathbb {H}_{0}^{s}(\Omega )\). If in addition we require that \(N>6s\), we can obtain the existence of a second solution to (1.1) by applying the mountain pass theorem (see [25, Theorem 2.2]).

3 Proof of Theorem 2

Let us consider the problem (1.1) with \(\lambda =\lambda _{1}\), namely

$$\begin{aligned} \left\{ \begin{array}{ll} (-\Delta )^{s} u = \lambda _{1} u + u_{+}^{2^{*}_{s}-1} + f(x) &{} \text{ in } \Omega ,\\ u=0 &{} \text{ in } \mathbb {R}^{N}\setminus \Omega , \end{array} \right. \end{aligned}$$
(3.1)

by assuming that (1.6) and (1.7) are satisfied. Let \(\mathbb {H}_{0}^{s}(\Omega )= \mathbb {H}^{-}\oplus \mathbb {H}^{+}\), where \(\mathbb {H}^{-}:= \mathrm{span}\{e_{1}\}\) and \(\mathbb {H}^{+}:=(\mathbb {H}^{-})^{\perp }\). The Euler-Lagrange functional \(\mathcal {J}:\mathbb {H}_{0}^{s}(\Omega )\rightarrow \mathbb {R}\) associated with (3.1) is given by

$$\begin{aligned} \mathcal {J}(u):= \frac{1}{2} \Vert u\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} - \frac{\lambda _{1}}{2} \Vert u\Vert ^{2}_{L^{2}(\Omega )} - \frac{1}{2^{*}_{s}} \Vert u_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} - \int _{\Omega } f u \, dx, \end{aligned}$$

and its differential is

$$\begin{aligned} \langle \mathcal {J}'(u), \phi \rangle = \langle u, \phi \rangle _{\mathbb {H}_{0}^{s}(\Omega )} - \lambda _{1} \int _{\Omega } u\phi \, dx - \int _{\Omega } u_{+}^{2^{*}_{s}-1} \phi \, dx - \int _{\Omega } f\phi \, dx, \end{aligned}$$

for \(u, \phi \in \mathbb {H}_{0}^{s}(\Omega )\). For any \(u\in \mathbb {H}_{0}^{s}(\Omega )\) there are \(t\in \mathbb {R}\) and \(v\in \mathbb {H}^{+}\) such that \(u= te_{1}+v\). Then, the functional \(\mathcal {J}\) becomes

$$\begin{aligned} \mathcal {J}(u)= \frac{1}{2} \Vert v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} -\frac{\lambda _{1}}{2} \Vert v\Vert ^{2}_{L^{2}(\Omega )}- \frac{1}{2^{*}_{s}} \Vert (v+te_{1})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} - \int _{\Omega } f\, (v+te_{1})\, dx, \end{aligned}$$

where \(u=v+te_{1}\) and \(t= \int _{\Omega } ue_{1}\, dx\).

Lemma 9

For any given \(v\in \mathbb {H}^{+}\), \(\mathcal {J}\) is bounded from above in \(\mathbb {H}^{-}\).

Proof

Fix \(v\in \mathbb {H}^{+}\). Let us define

$$\begin{aligned} \psi (t):=\mathcal {J}(v+te_{1}), \, t\in \mathbb {R}. \end{aligned}$$

Note that, for \(t<0\), the Hölder inequality yields

$$\begin{aligned} \psi (t)\le \frac{1}{2} \Vert v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )}-\frac{\lambda _{1}}{2} \Vert v\Vert ^{2}_{L^{2}(\Omega )} + \Vert f\Vert _{L^{2}(\Omega )} \Vert v\Vert _{L^{2}(\Omega )}. \end{aligned}$$

Now, let \(t>0\). We first show that

$$\begin{aligned} \lim _{t\rightarrow \infty }\left\{ \frac{1}{2^{*}_{s}} \Vert (v+te_{1})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} + \int _{\Omega } f\, (v+te_{1})\, dx\right\} =\infty . \end{aligned}$$
(3.2)

Put \(M:=\sup _{x\in \Omega } e_{1}(x)\) and choose an open set \(\Omega _{0} \Subset \Omega\) such that \(e_{1}(x)> \frac{M}{4}\) for \(x\in \Omega _{0}\). Let \(\delta := \frac{|\Omega _{0}|}{2}\). Combining Lusin’s theorem with the Tietze extension theorem, we find \(h\in C^{0}(\mathbb {R}^{N})\) such that

$$\begin{aligned} |\{x\in \Omega _{0} : h(x) \ne v(x)\}|< \delta . \end{aligned}$$

Set \(A:=\{x\in \Omega _{0} : h(x)=v(x)\}\) and note that

$$\begin{aligned} |A|> |\Omega _{0}|-\delta =\frac{|\Omega _{0}|}{2}. \end{aligned}$$

Put \(L:=\sup _{x\in A} |v(x)|\). For \(x\in A\), we get

$$\begin{aligned} e_{1}(x)+ \frac{v(x)}{t}\ge \frac{M}{4}- \frac{L}{t}\ge \frac{M}{8} \quad \text{ for } t\ge T:=\frac{8L}{M}, \end{aligned}$$

and thus there exists \(\kappa >0\) such that

$$\begin{aligned} \frac{1}{2^{*}_{s}}\int _{\Omega } (v+te_{1})_{+}^{2^{*}_{s}} dx =\frac{t^{2^{*}_{s}}}{2^{*}_{s}} \int _{\Omega } \left( \frac{v}{t}+ e_{1} \right) _{+}^{2^{*}_{s}}\, dx \ge \kappa t^{2^{*}_{s}} \quad \text{ for } t\ge T. \end{aligned}$$

The above inequality implies that (3.2) holds. Using the continuity of \(\psi\) in [0, T], we can infer that \(\mathcal {J}\) is bounded from above in \(\mathbb {H}^{-}\).

It is possible to prove that for each \(v\in \mathbb {H}^{+}\) there exists a unique \(t(v)\in \mathbb {R}\) such that

$$\begin{aligned} \psi (t(v)):=\max _{t\in \mathbb {R}} \psi (t), \end{aligned}$$

that is \(\mathcal {J}(v+te_{1})\le \mathcal {J}(v+t(v)e_{1})\) for all \(t\in \mathbb {R}\). By invoking the implicit function theorem, we can define a continuously differentiable map

$$\begin{aligned} v\in \mathbb {H}^{+} \rightarrow t(v)\in \mathbb {R}\end{aligned}$$

such that \(\mathcal {J}(v+te_{1})< \mathcal {J}(v+t(v)e_{1})\) for all \(t\in \mathbb {R}\) such that \(t\ne t(v)\). Since, for \(v\in \mathbb {H}^{+}\), \(\psi '(t(v))=0\), we have

$$\begin{aligned} \int _{\Omega } (v+t(v)e_{1})_{+}^{2^{*}_{s}-1} e_{1} \,dx + \int _{\Omega } f e_{1} \,dx=0. \end{aligned}$$
(3.3)

Taking \(v=0\) in (3.3), we find

$$\begin{aligned} \int _{\Omega } (t(0)e_{1})_{+}^{2^{*}_{s}-1} e_{1} \,dx + \int _{\Omega } f e_{1} \,dx=0, \end{aligned}$$

which combined with (1.6) shows that \(t(0)>0\) and

$$\begin{aligned} t(0)^{2^{*}_{s}-1} \int _{\Omega } e_{1}^{2^{*}_{s}} dx =-\int _{\Omega } f e_{1} \,dx. \end{aligned}$$
(3.4)

Let us now introduce \(\mathcal {F}:\mathbb {H}^{+}\rightarrow \mathbb {R}\) given by

$$\begin{aligned} \mathcal {F}(v):= \mathcal {J}(v+ t(v)e_{1}). \end{aligned}$$

We claim that \(\mathcal {F}\) attains its minimum on some ball \(B_{\rho }\). It is easy to see that (1.6) and (3.4) give

$$\begin{aligned} \mathcal {F}(0)&=-\frac{1}{2^{*}_{s}} t(0)^{2^{*}_{s}} \int _{\Omega } e_{1}^{2^{*}_{s}} dx - t(0)\int _{\Omega } f e_{1} \,dx \nonumber \\&=\frac{N+2s}{2N} \left( -\int _{\Omega } f e_{1} \,dx\right) ^{\frac{2N}{N+2s}}\left( \int _{\Omega } e_{1}^{2^{*}_{s}} dx \right) ^{-\frac{N-2s}{N+2s}}>0. \end{aligned}$$
(3.5)

Lemma 10

There exists \(\beta >0\) such that

$$\begin{aligned} \mathcal {F}(v)\ge \beta \ge \mathcal {F}(0) \, \text{ whenever } \Vert v\Vert _{\mathbb {H}_{0}^{s}(\Omega )}=\varrho _{0}, \end{aligned}$$
(3.6)

where

$$\begin{aligned} \varrho _{0}:=\left( 1- \frac{\lambda _{1}}{\lambda _{2}}\right) ^{\frac{N-2s}{4s}} \left( \frac{N}{N+2s}\right) ^{\frac{N-2s}{4s}} S_{*}^{\frac{N}{4s}}. \end{aligned}$$

Proof

Note that, for all \(v\in \mathbb {H}^{+}\), it holds

$$\begin{aligned} \Vert v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} \ge \lambda _{2}\Vert v\Vert ^{2}_{L^{2}(\Omega )}. \end{aligned}$$

Using this, the Hölder inequality and the Sobolev inequality [13, Theorem 1.1], we have

$$\begin{aligned} \mathcal {F}(v)&= \mathcal {J}(v+ t(v)e_{1}) \ge \mathcal {J}(v) \nonumber \\&\ge \frac{1}{2}\left( 1-\frac{\lambda _{1}}{\lambda _{2}}\right) \Vert v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} - \frac{1}{2^{*}_{s}} \Vert v\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )}- \Vert f\Vert _{L^{2}(\Omega )}\Vert v\Vert _{L^{2}(\Omega )} \nonumber \\&\ge \frac{1}{2}\left( 1-\frac{\lambda _{1}}{\lambda _{2}}\right) \Vert v\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} - \frac{1}{2^{*}_{s}} S_{*}^{-\frac{N}{N-2s}} \Vert v\Vert ^{2^{*}_{s}}_{\mathbb {H}_{0}^{s}(\Omega )}- \Vert f\Vert _{L^{2}(\Omega )} \lambda _{2}^{-\frac{1}{2}} \Vert v\Vert _{\mathbb {H}_{0}^{s}(\Omega )}. \end{aligned}$$
(3.7)

Define

$$\begin{aligned} \kappa (\zeta ):=\frac{1}{2} \left( 1-\frac{\lambda _{1}}{\lambda _{2}}\right) \, \zeta ^{2} - \frac{1}{2^{*}_{s}} S_{*}^{-\frac{N}{N-2s}}\,\zeta ^{2^{*}_{s}}- \Vert f\Vert _{L^{2}(\Omega )} \lambda _{2}^{-\frac{1}{2}} \zeta =:\zeta j(\zeta ), \, \zeta \ge 0. \end{aligned}$$

Let us compute \(j'(\zeta )\):

$$\begin{aligned} j'(\zeta )=\frac{1}{2}\left( 1-\frac{\lambda _{1}}{\lambda _{2}}\right) - \frac{N+2s}{2N} S_{*}^{-\frac{N}{N-2s}} \zeta ^{\frac{4s}{N-2s}}. \end{aligned}$$

Then \(j(\zeta )\) achieve its maximum at \(\varrho _{0}\) given by

$$\begin{aligned} \varrho _{0}:= \left( \left( 1-\frac{\lambda _{1}}{\lambda _{2}}\right) S_{*}^{\frac{N}{N-2s}}\frac{N}{N+2s}\right) ^{\frac{N-2s}{4s}}= \left( 1- \frac{\lambda _{1}}{\lambda _{2}}\right) ^{\frac{N-2s}{4s}} \left( \frac{N}{N+2s}\right) ^{\frac{N-2s}{4s}} S_{*}^{\frac{N}{4s}}. \end{aligned}$$

Therefore,

$$\begin{aligned} \kappa (\varrho _{0})&=\varrho _{0} \left[ \frac{1}{2} \left( 1- \frac{\lambda _{1}}{\lambda _{2}}\right) ^{\frac{N+2s}{4s}} \!\!\left( \frac{N}{N+2s}\right) ^{\frac{N-2s}{4s}}\!\!S_{*}^{\frac{N}{4s}} - \frac{N-2s}{2N} \left( 1- \frac{\lambda _{1}}{\lambda _{2}}\right) ^{\frac{N+2s}{4s}} \!\!\left( \frac{N}{N+2s}\right) ^{\frac{N+2s}{4s}} \!\!S_{*}^{\frac{N}{4s}} - \Vert f\Vert _{L^{2}(\Omega )} \lambda _{2}^{-\frac{1}{2}} \right] \\&=\varrho _{0} \left[ \left( 1- \frac{\lambda _{1}}{\lambda _{2}}\right) ^{\frac{N+2s}{4s}} S_{*}^{\frac{N}{4s}} \left( \frac{N}{N+2s}\right) ^{\frac{N-2s}{4s}} \frac{2s}{N+2s} -\Vert f\Vert _{L^{2}(\Omega )} \lambda _{2}^{-\frac{1}{2}} \right] , \end{aligned}$$

and using (1.8) and the first condition in (1.7), we see that

$$\begin{aligned} \Vert f\Vert _{L^{2}(\Omega )} \lambda _{2}^{-\frac{1}{2}} \le \frac{s}{N+2s} S_{*}^{\frac{N}{4s}} \left( \frac{N}{N+2s}\right) ^{\frac{N-2s}{4s}} \left( 1- \frac{\lambda _{1}}{\lambda _{2}}\right) ^{\frac{N+2s}{4s}}. \end{aligned}$$

Thus (3.7) yields

$$\begin{aligned} \mathcal {F}(v)&\ge \frac{s \varrho _{0}}{N+2s} \left( 1- \frac{\lambda _{1}}{\lambda _{2}}\right) ^{\frac{N+2s}{4s}}S_{*}^{\frac{N}{4s}}\left( \frac{N}{N+2s}\right) ^{\frac{N-2s}{4s}}\nonumber \\&= \frac{s}{N+2s} \left( 1- \frac{\lambda _{1}}{\lambda _{2}}\right) ^{\frac{N}{2s}}S_{*}^{\frac{N}{2s}}\left( \frac{N}{N+2s}\right) ^{\frac{N-2s}{2s}}\nonumber \\&= \frac{s}{N} S_{*}^{\frac{N}{2s}}\left( 1- \frac{\lambda _{1}}{\lambda _{2}}\right) ^{\frac{N}{2s}} \left( \frac{N}{N+2s}\right) ^{\frac{N}{2s}}=:\beta \end{aligned}$$
(3.8)

provided \(\Vert v\Vert _{\mathbb {H}_{0}^{s}(\Omega )}= \varrho _{0}\). Let us observe that (3.5), the second condition in (1.7) and (1.9) yield

$$\begin{aligned} \mathcal {F}(0)&<\frac{N+2s}{2N}\frac{M_{2}^{\frac{2N}{N+2s}}}{\Vert e_{1} \Vert _{L^{2^{*}_{s}}(\Omega )}^{\frac{2N}{N+2s}}}\nonumber \\&\le \frac{s}{N} S_{*}^{\frac{N}{2s}}\left( 1- \frac{\lambda _{1}}{\lambda _{2}}\right) ^{\frac{N}{2s}} \left( \frac{N}{N+2s}\right) ^{\frac{N}{2s}}, \end{aligned}$$
(3.9)

which combined with (3.8) ensures that (3.6) is satisfied.

Proof of Theorem 2

Let us define

$$\begin{aligned} m:= \inf \{\mathcal {F}(v) : v\in \bar{B}_{\varrho _{0}}\}. \end{aligned}$$
(3.10)

We claim that (3.1) has a nontrivial solution \(v_{0}\in B_{\varrho _{0}}\). From (3.9) easily follows that

$$\begin{aligned} \mathcal {F}(0)<\frac{s}{N} S_{*}^{\frac{N}{2s}}, \end{aligned}$$

and in particular

$$\begin{aligned} m<\frac{s}{N} S_{*}^{\frac{N}{2s}}. \end{aligned}$$
(3.11)

Let \(\{v_{n}\}_{n\in \mathbb {N}}\subset \mathbb {H}_{0}^{s}(\Omega )\) be a minimizing sequence of (3.10). Since \(\Vert v_{n}\Vert _{\mathbb {H}_{0}^{s}(\Omega )}\le \varrho _{0}\), we may assume that, up to a subsequence,

$$\begin{aligned} \begin{aligned}&v_{n}\rightharpoonup v_{0} \, \text{ in } \, \mathbb {H}_{0}^{s}(\Omega ),\\&v_{n}\rightarrow v_{0} \, \text{ in } \, L^{q}(\mathbb {R}^{N}), 1\le q<2^{*}_{s}, \\&v_{n}\rightarrow v_{0} \, \text{ a.e. } \text{ in } \, \mathbb {R}^{N}. \end{aligned} \end{aligned}$$
(3.12)

From the weak lower semicontinuity of the norm \(\Vert \cdot \Vert _{\mathbb {H}_{0}^{s}(\Omega )}\), we have

$$\begin{aligned} \Vert v_{0}\Vert _{\mathbb {H}_{0}^{s}(\Omega )} \le \lim _{n\rightarrow \infty } \Vert v_{n}\Vert _{\mathbb {H}_{0}^{s}(\Omega )} \le \varrho _{0}. \end{aligned}$$

By Ekeland’s variational principle [23, Theorem 4.1], we may assume that

$$\begin{aligned} \mathcal {F}(v_{n})\rightarrow m \quad \text{ and } \quad \mathcal {F}'(v_{n})\rightarrow 0, \end{aligned}$$

as \(n\rightarrow \infty\). Consequently,

$$\begin{aligned}&\frac{1}{2}\Vert v_{n}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} - \frac{\lambda _{1}}{2} \Vert v_{n}\Vert ^{2}_{L^{2}(\Omega )} - \frac{1}{2^{*}_{s}} \Vert (v_{n} + t(v_{n}) e_{1})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} - \int _{\Omega } f(v_{n}+ t(v_{n})e_{1})\, dx \nonumber \\&= m+o(1), \end{aligned}$$
(3.13)

and

$$\begin{aligned} \Vert v_{n}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )}- \lambda _{1} \Vert v_{n}\Vert ^{2}_{L^{2}(\Omega )} - \int _{\Omega } (v_{n} + t(v_{n}) e_{1})_{+}^{2^{*}_{s}-1} v_{n} \,dx - \int _{\Omega } f v_{n} \,dx=o(1). \end{aligned}$$
(3.14)

From (3.12) we derive that \(v_{0}\) is a weak solution to

$$\begin{aligned} \left\{ \begin{array}{ll} (-\Delta )^{s} v_{0} = \lambda _{1} v_{0} + (v_{0}+t(v_{0})e_{1})^{2^{*}_{s}-1}_{+}+ f(x) &{} \text{ in } \Omega ,\\ v_{0}=0 &{} \text{ in } \mathbb {R}^{N}\setminus \Omega . \end{array} \right. \end{aligned}$$
(3.15)

Using \(v_{0}\) as test function in the weak formulation of (3.15), we acquire

$$\begin{aligned} \Vert v_{0}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} - \lambda _{1} \Vert v_{0}\Vert ^{2}_{L^{2}(\Omega )} - \int _{\Omega } (v_{0}+ t(v_{0})e_{1})_{+}^{2^{*}_{s}-1} v_{0} \,dx- \int _{\Omega } f v_{0} \,dx=0. \end{aligned}$$
(3.16)

On the other hand, choosing \(e_{1}\) as test function in the weak formulation of (3.15) and recalling that

$$\begin{aligned} \left\{ \begin{array}{ll} (-\Delta )^{s} e_{1} = \lambda _{1} e_{1} &{} \text{ in } \Omega ,\\ e_{1}=0 &{} \text{ in } \mathbb {R}^{N}\setminus \Omega , \end{array} \right. \end{aligned}$$

we obtain

$$\begin{aligned} \int _{\Omega } (v_{0}+ t(v_{0})e_{1})_{+}^{2^{*}_{s}-1} e_{1} \,dx+ \int _{\Omega } f e_{1}\, dx=0. \end{aligned}$$
(3.17)

Next we show that \(v_{0}\not \equiv 0\). We start by observing that

$$\begin{aligned} \lim _{n\rightarrow \infty } t(v_{n})= t(v_{0}). \end{aligned}$$

Indeed, if we assume by contradiction that \(t(v_{n})\rightarrow t_{1}\ne t(v_{0})\), using (3.3), we see that

$$\begin{aligned} \int _{\Omega } (v_{n}+ t(v_{n})e_{1})_{+}^{2^{*}_{s}-1} e_{1} \,dx+ \int _{\Omega } f e_{1}\, dx=0, \end{aligned}$$

and

$$\begin{aligned} \int _{\Omega } (v_{0}+ t(v_{0})e_{1})_{+}^{2^{*}_{s}-1} e_{1} \,dx+ \int _{\Omega } f e_{1} \,dx=0, \end{aligned}$$

imply

$$\begin{aligned} \int _{\Omega } (v_{0}+ t_{1}e_{1})_{+}^{2^{*}_{s}-1} e_{1} dx= \int _{\Omega } (v_{0}+ t(v_{0})e_{1})_{+}^{2^{*}_{s}-1} e_{1} dx, \end{aligned}$$

that is a contradiction. Set \(w_{n}:= v_{n}- v_{0}\). Using (3.13), (3.14), and the Brezis–Lieb lemma [8, Theorem 1], we get

$$\begin{aligned}&\frac{1}{2}\Vert w_{n}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} - \frac{1}{2^{*}_{s}}\Vert (w_{n})_{+} \Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} + \frac{1}{2} \Vert v_{0}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} -\frac{\lambda _{1}}{2} \Vert v_{0}\Vert ^{2}_{L^{2}(\Omega )} \\&\quad - \frac{1}{2^{*}_{s}} \Vert (v_{0}+ t(v_{0})e_{1})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}} (\Omega )}- \int _{\Omega } f(v_{0}+ t(v_{0})e_{1}) \, dx=m+o(1), \end{aligned}$$

namely

$$\begin{aligned} \mathcal {F}(v_{0})+\frac{1}{2} \Vert w_{n}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )}- \frac{1}{2^{*}_{s}} \Vert (w_{n})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )}= m+o(1). \end{aligned}$$
(3.18)

Analogously, from (3.14), (3.16), (3.17) and the Brezis–Lieb lemma [8, Theorem 1], we have

$$\begin{aligned} \Vert w_{n}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )}&- \Vert (w_{n})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )} + \Vert v_{0}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )}-\lambda _{1} \Vert v_{0}\Vert ^{2}_{L^{2}(\Omega )} \\&\quad -\Vert (v_{0}+ t(v_{0})e_{1})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )}-\int _{\Omega } f(v_{0}+ t(v_{0})e_{1}) \, dx=o(1), \end{aligned}$$

that is

$$\begin{aligned} \Vert w_{n}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )}= \Vert (w_{n})_{+}\Vert ^{2^{*}_{s}}_{L^{2^{*}_{s}}(\Omega )}+o(1). \end{aligned}$$
(3.19)

Let

$$\begin{aligned} \ell :=\lim _{n\rightarrow \infty } \Vert w_{n}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} \ge 0. \end{aligned}$$

If \(\ell =0\), then we have done. Suppose \(\ell >0\). By the Sobolev inequality [13, Theorem 1.1], we know that

$$\begin{aligned} \Vert w_{n}\Vert ^{2}_{\mathbb {H}_{0}^{s}(\Omega )} \ge S_{*} \Vert (w_{n})_{+}\Vert ^{2}_{L^{2^{*}_{s}}(\Omega )}. \end{aligned}$$

Passing to the limit as \(n\rightarrow \infty\) in the above inequality, and using (3.19) and \(\ell >0\), we find

$$\begin{aligned} \ell \ge S_{*} \ell ^{\frac{N-2s}{N}}, \, \text{ i.e. } \ell \ge S_{*}^{\frac{N}{2s}}. \end{aligned}$$
(3.20)

Combining (3.11), (3.18), (3.19) and (3.20), we get

$$\begin{aligned} \frac{s}{N} S_{*}^{\frac{N}{2s}}>m= \mathcal {F}(v_{0}) + \left( \frac{1}{2}- \frac{1}{2^{*}_{s}}\right) \ell = \mathcal {F}(v_{0}) + \frac{s}{N}\ell \ge \mathcal {F}(v_{0}) + \frac{s}{N}S_{*}^{\frac{N}{2s}}, \end{aligned}$$

which implies that \(\mathcal {F}(v_{0})<0\). This last fact and (3.5) shows that \(v_{0}\not \equiv 0\). Since \(\mathcal {F}(v)\ge \beta >0\) for all \(\Vert v\Vert _{\mathbb {H}_{0}^{s}(\Omega )}= \varrho _{0}\), we can conclude that \(v_{0}\in B_{\varrho _{0}}\). The proof of Theorem 2 is complete.

Remark 3

Following [16], it is also possible to discuss the local bifurcation at \(\lambda =\lambda _{n}\) with \(n>1\).

Remark 4

If we replace \((-\Delta )^{s}\) by the more general nonlocal integro-differential operator \(-\mathcal {L}_{K}\) given by

$$\begin{aligned} \mathcal {L}_{K}u(x):=\int _{\mathbb {R}^{N}} (u(x+y)+u(x-y)-2u(x))K(y)\, dy, \quad x\in \mathbb {R}^{N}, \end{aligned}$$

where \(K:\mathbb {R}^{N}\setminus \{0\}\rightarrow (0, \infty )\) is a function such that:

  • \(mK\in L^{1}(\mathbb {R}^{N})\), where \(m(x):=\min \{|x|^{2}, 1\}\),

  • there exists \(\theta >0\) such that \(K(x)\ge \theta |x|^{-(N+2s)}\) for any \(x\in \mathbb {R}^{N}\setminus \{0\}\),

  • \(K(x)=K(-x)\) for any \(x\in \mathbb {R}^{N}\setminus \{0\}\),

then we can easily adapt our arguments to extend Theorem 1 and Theorem 2 to the following problem:

$$\begin{aligned} \left\{ \begin{array}{ll} -\mathcal {L}_{K} u = \lambda u + u_{+}^{2^{*}_{s}-1} + f(x) &{} \text{ in } \Omega ,\\ u=0 &{} \text{ in } \mathbb {R}^{N}\setminus \Omega . \end{array} \right. \end{aligned}$$