1 Introduction

In the standard framework of the p-Laplacian operator, there are a lot of interesting problems widely studied in the literature. A natural question is whether or not the results got in this classical context can be extended to the nonlocal case in the presence of fractional-type operators. Nonlocal fractional equations appear in many fields, and after the seminal works by Caffarelli and Silvestre [810], an increasing interest has been devoted to these topics; we refer, for instance, the recent papers [7, 12], the monograph [21] for several results on fractional variational problems and related topics, and [2527] as well as references therein for some existence and multiplicity results involving the fractional p-Laplacian operator.

In most of the papers concerning with fractional Laplacian equation, it is assumed that the right-hand side has a superlinear, but subcritical growth (cf., e.g., [5, 22, 28, 29] and references therein). In the recent papers [3, 14] it has been firstly studied in the nonlocal setting the so-called asymptotically linear case: as it is well known, the lack of the classical Ambrosetti–Rabinowitz assumption requires more efforts in order to obtain a compactness Palais–Smale-type condition.

Further difficulties arise in the so-called resonant case (cf., e.g., [1] and references therein): indeed, the resonance affects both the compactness property and the geometry of the Euler–Lagrange functional arising in a suitable variational approach. For the local case we recall the contributions [2, 18, 24].

Motivated by this interest in the current literature, we would like to focus here our attention on p-fractional “asymptotically linear” problems, also in the presence of resonance. Indeed, we look for solutions of the nonlocal elliptic problem

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle {(-\varDelta )^s_p u\ =\ \lambda |u|^{p-2}u+ f(x, u)} &{} \text{ in } \Omega ,\\ \displaystyle {u=0} &{} \text{ on } \mathbb {R} ^N{\setminus }\Omega ,\\ \end{array} \right. \end{aligned}$$
(1.1)

where \(1<p<+\infty , s\in ]0,1[, (-\varDelta )^s_p\) denotes the fractional p-Laplacian which (up to normalization factors) may be defined for any \(x\in {\mathbb {R}}^{N}\) as

$$\begin{aligned} (-\varDelta )^s_p\varphi (x)=2\lim _{\varepsilon \searrow 0^+}\int _{\mathbb {R} ^N{\setminus } B_{\varepsilon }(x)}\frac{\left| \varphi (x)-\varphi (y)\right| ^{p-2}(\varphi (x) -\varphi (y))}{|x-y|^{N+sp}}\mathrm {d}y \end{aligned}$$

along any \(\varphi \in C^\infty _0(\mathbb {R} ^N)\), where \(B_\varepsilon (x):=\left\{ y\in \mathbb {R} ^N\,:\,\left| x-y\right| <\varepsilon \right\} \), \(\Omega \) is an open bounded domain of \(\mathbb {R} ^N\) (\(N>sp\)) with Lipschitz boundary \(\partial \Omega \) and \(f:\Omega \times \mathbb {R} \rightarrow \mathbb {R} \) is a Carathéodory function such that:

\(({h}_1)\) :

\( \displaystyle \sup _{|t|\le a}|f(\cdot ,t)|\in L^\infty (\Omega ) \quad \forall \, a>0; \)

\(({h}_2)\) :

there exist

$$\begin{aligned} \lim _{|t|\rightarrow + \infty }\frac{f(x,t)}{|t|^{p-2}t}\ =\ 0 \end{aligned}$$
(1.2)

and

$$\begin{aligned} \lim _{t\rightarrow 0}\frac{f(x,t)}{|t|^{p-2}t} \ = \alpha \in \mathbb {R}\ \end{aligned}$$
(1.3)

uniformly with respect to a.e. \(x \in \Omega \).

Here we also deal with the resonant case, under the following further assumption (cf., e.g., [18]):

\((h_3)\) :

there exists

$$\begin{aligned} \lim _{|t|\rightarrow + \infty }\ \Big (f(x,t)t-pF(x,t)\Big )=\ +\infty \end{aligned}$$

uniformly with respect to a.e. \(x \in \Omega \), where, as usual, we set

$$\begin{aligned} F(x,t):=\displaystyle \int _0^t f(x,\tau ) \; \mathrm{d}\tau . \end{aligned}$$

In order to state our multiplicity result we introduce some notations. Recalling that the Gagliardo seminorm of a measurable function \(u:\mathbb {R} ^N\rightarrow \mathbb {R} \) is defined by

$$\begin{aligned}{}[u]_{s,p}:=\left( \int _{\mathbb {R} ^N\times \mathbb {R} ^N} \frac{|u(x)-u(y)|^{p}}{|x-y|^{N+sp}}\,\mathrm {d}x \mathrm {d}y\right) ^{\frac{1}{p}}, \end{aligned}$$

the fractional Sobolev space is defined by

$$\begin{aligned} W^{s,p}(\mathbb {R} ^N):=\{u\in L^p(\mathbb {R} ^N): [u]_{s,p}<+\infty \} \end{aligned}$$

and is equipped with the norm

$$\begin{aligned} \Vert u\Vert _{s,p}:=\left( |u|_p^p + [u]_{s,p}^p\right) ^{\frac{1}{p}}, \end{aligned}$$

where \(|\cdot |_p\) denotes the norm on \(L^p(\mathbb {R} ^N)\).

Our problem is set in the closed linear space

$$\begin{aligned} X(\Omega ):=\{u\in W^{s,p}(\mathbb {R} ^N): u(x)=0 \hbox { a.e. in } \mathbb {R} ^N{\setminus }\Omega \} \end{aligned}$$

endowed with the norm \(\Vert \cdot \Vert = [\cdot ]_{s,p}\). The space \((X(\Omega ), \Vert \cdot \Vert )\) is uniformly convex and setting

$$\begin{aligned} p^*_s:=\displaystyle \frac{pN}{N-sp} \end{aligned}$$

it results:

$$\begin{aligned} X(\Omega )\hookrightarrow L^{\mu }(\Omega ) \hbox { continuously for }\mu \in [1,p^*_s] \end{aligned}$$

and

$$\begin{aligned} X(\Omega )\hookrightarrow \hookrightarrow L^{\mu }(\Omega )\hbox { compactly for }\mu \in [1,p^*_s[. \end{aligned}$$
(1.4)

Let us recall that \(\lambda \in \mathbb {R} \) is an eigenvalue for \((-\varDelta )^s_p\) if there exists a nontrivial weak solution on \(X(\Omega )\) of the problem

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle {(-\varDelta )^s_p u\ =\ \lambda |u|^{p-2}u} &{} \text{ in } \Omega ,\\ \displaystyle {u=0} &{} \text{ on } \mathbb {R} ^N{\setminus }\Omega ,\\ \end{array} \right. \end{aligned}$$
(1.5)

that is \(u\in X(\Omega ){\setminus }\{0\}\) such that

$$\begin{aligned} \begin{aligned}&\iint _{\mathbb {R} ^N\times \mathbb {R} ^N}\frac{\left| u(x)-u(y)\right| ^{p-2}(u(x)-u(y))(\varphi (x)-\varphi (y))}{\left| x-y\right| ^{N+sp}}\,\mathrm {d}x\mathrm {d}y\\&\quad -\lambda \int _{\Omega }\left| u(x)\right| ^{p-2}u(x)\varphi (x) \mathrm {d}x=0, \end{aligned} \end{aligned}$$

for every \(\varphi \in X(\Omega )\).

As usual, the set of the eigenvalues is named spectrum and it is denoted by \(\sigma ((-\varDelta )^s_p))\).

It is well known that if \(p=2\) the spectrum of \((-\varDelta )^s\) in \(X(\Omega )\) consists of a diverging sequence \(\{\lambda _n\}_{n\in \mathbb {N}}\) of eigenvalues, repeated according to their multiplicity, satisfying \(0<\lambda _1<\lambda _2\le \cdots \le \lambda _k\le \cdots \) (see [29]). Moreover, for \(p=2\) the structure of \(\sigma ((-\varDelta )^s)\) is very closed to that of the local operator \(-\varDelta \) and also provides a decomposition by means of the eigenfunctions.

On the other hand, when \(p\not =2\) the full spectrum of \((-\varDelta )^s_p\) is still almost unknown, even if some important properties of the first eigenvalue and of the higher-order (variational) eigenvalues have been established in [15, 19]. We also point out that a sequence of eigenvalues has been introduced in [17] by means of the cohomological index.

As we shall see in the proof of our main result, the definition of the quasi-eigenvalues proposed here and inspired to that in [11] fits in with our purposes, as a suitable decomposition of \(X(\Omega )\) can be introduced and it turns out to be the known one for \(p=2\) (cf. Sect. 2 for the details).

Thus, the interaction of the nonlinearity f with the spectrum \(\sigma ((-\varDelta )^s_p)\) of the fractional p-Laplacian operator must be taken into account. To our knowledge the only contribution in this direction is by [17], where the authors obtain existence results via Morse theory.

Now we state our main result.

Theorem 1.1

Let \(s\in ]0,1[, N>sp\) and \(\Omega \) be an open bounded subset of \(\mathbb {R} ^N\) with continuous boundary. Assume that \((h_1)\)\((h_2)\) hold, \(f(x,\cdot )\) is odd for a.e. \(x\in \Omega \) and that

\((h_4)\) :

there exist \(h, k \in \mathbb {N} \) with \(k\ge h\) such that

$$\begin{aligned} \alpha + \lambda<\ \beta _h\le \ \gamma _k\ <\lambda , \end{aligned}$$

where \(\{\beta _n\}_{n\in \mathbb {N}}, \{\gamma _n\}_{n\in \mathbb {N}}\) are, respectively, as in (2.11) and (2.13) below.

Then, problem (1.1) has at least \(k-h+1\) distinct pairs of nontrivial weak solutions, provided either

(a):

\(\lambda \not \in \sigma ((-\varDelta )^s_p)\) or

(b):

\(\lambda \in \sigma ((-\varDelta )^s_p)\) and \((h_3)\) holds true.

For the sake of completeness we recall here that a function \(u\in X(\Omega )\) is a weak solution of problem (1.1) if

for every \(\varphi \in X(\Omega )\).

This paper is organized as follows. In Sect. 2 we introduce the sequences of quasi-eigenvalues and depict some properties; then, in Sect. 3 we prove Theorem 1.1 by making use of a pseudo-index result recalled in Appendix as well as other classical tools.

2 Splitting of the fractional space \(X(\Omega )\)

In this section we adapt the arguments in [11, Section 5] to the nonlocal setting, thus constructing a first sequence of quasi-eigenvalues for \((-\varDelta )_p^s\) on \(X(\Omega )\). Let us define the subset

$$\begin{aligned} {\mathcal S}:= \left\{ v \in X(\Omega ): \int _{\mathbb {R} ^N} |v(x)|^p \hbox {d}x = 1\right\} , \end{aligned}$$
(2.1)

the functional \(\Phi :X(\Omega ) \rightarrow \mathbb {R} \) by

$$\begin{aligned} \Phi (u):= \Vert u\Vert ^p=[u]_{s,p}^p \end{aligned}$$
(2.2)

and

$$\begin{aligned} \beta _1 := \inf _{v \in {\mathcal S}}\Phi (v) > 0 \end{aligned}$$

(\(\beta _1\) is indeed the first eigenvalue of the p-fractional operator). Let us remark that

  • \(X(\Omega )\) is a reflexive Banach space;

  • \({\mathcal S}\) is weakly closed in \(X(\Omega )\) (by (1.4));

  • \(\Phi \) is coercive on \(X(\Omega )\);

  • \(\Phi \) is weakly lower semicontinuous on \(X(\Omega )\).

Then, by the generalized Weierstrass theorem, there exists a function \(\psi _1 \in X(\Omega )\) such that

$$\begin{aligned} \int _{\mathbb {R} ^N} |\psi _1(x)|^p \mathrm {d}x = 1,\quad \Phi (\psi _1) = [\psi _1]_{s,p}^p =\beta _1; \end{aligned}$$
(2.3)

therefore,

$$\begin{aligned} \beta _1 |u|_p^p \le \Phi (u) \quad \hbox {for all}\, u\in X(\Omega ). \end{aligned}$$

Let us consider the linear operator \({\mathcal L}_1: L^p(\mathbb {R} ^N) \rightarrow \mathbb {R} \) related to \(\psi _1\) (\(\psi _1\in W^{s,p}(\mathbb {R} ^N)\Rightarrow \psi _1\in L^p(\mathbb {R} ^N)\) thus \(|\psi _1|^{p-1}\in L^{p'}(\mathbb {R} ^N)\), where \(p'\) is the conjugate of p) defined by

$$\begin{aligned} {\mathcal L}_1u := \int _{\mathbb {R} ^{N}} |\psi _1(x)|^{p-2}\psi _1(x)u(x)\,\mathrm {d}x. \end{aligned}$$

By definition, \({\mathcal L}_1 \in L^{p'}(\mathbb {R} ^N)\), while (2.3) implies \({\mathcal L}_1\psi _1 = 1\), then \(\Vert {\mathcal L}_1\Vert _{L^{p'}} = 1\). Denoting again by \({\mathcal L}_1\) the restriction to the subspace \(X(\Omega )\), it is also \({\mathcal L}_1 \in (X(\Omega ))'\).

Now, we define the new constraint

$$\begin{aligned} {\mathcal S}_1 := \{v \in {\mathcal S}: {\mathcal L}_1 v = 0\} = \ker ({\mathcal L}_1|_{\mathcal S}) \end{aligned}$$

and the corresponding constrained infimum

$$\begin{aligned} \beta _2 := \inf _{v \in {\mathcal S}_1}\Phi (v). \end{aligned}$$

We have that \(\beta _2 > \beta _1\) (the first eigenvalue is isolated). We claim that also \({\mathcal S}_1\) is weakly closed in \(X(\Omega )\). In fact, taking a sequence \(\{v_m\}_{m\in \mathbb {N}} \subset {\mathcal S}_1\) and \(v\in X(\Omega )\) such that

$$\begin{aligned} v_m \rightharpoonup v\; \hbox { weakly in}\,X(\Omega ), \end{aligned}$$

by (1.4) it follows that

$$\begin{aligned} v_m \rightarrow v\; \hbox { strongly in}\,L^p(\mathbb {R} ^N) \end{aligned}$$

and, since \({\mathcal L}_1 \in L^{p'}\) and \(v_m \in {\mathcal S}_1\) for each \(m \in \mathbb {N} \), we get that \({\displaystyle \int _{\mathbb {R} ^N} |v(x)|^p \mathrm {d}x = 1}\) and \({\mathcal L}_1 v_m \rightarrow {\mathcal L}_1 v\); therefore, \(v \in {\mathcal S}_1\).

Thus, the generalized Weierstrass theorem applies again and there exists \(\psi _2 \in {\mathcal S}_1\) such that \(\Phi (\psi _2)=\beta _2\), i.e.,

$$\begin{aligned} \int _{\mathbb {R} ^N} |\psi _2(x)|^p \hbox {d}x = 1,\quad {\mathcal L}_1\psi _2=0,\quad \Phi (\psi _2)= \beta _2. \end{aligned}$$

The procedure can be repeated, so fixing any \(n\in \mathbb {N} \) we can define some positive numbers

$$\begin{aligned} \beta _1 < \beta _2 \le \cdots \le \beta _n \end{aligned}$$

and some functions

$$\begin{aligned} \psi _1,\psi _2,\ldots ,\psi _n \in {\mathcal S}\end{aligned}$$

such that, for each \(i \in \{1,\ldots ,n\}\), related to \(\psi _i\) we can consider the linear operator \({\mathcal L}_{i}\in L^{p'}\) defined by

$$\begin{aligned} {\mathcal L}_{i}u := \int _{\mathbb {R} ^N} |\psi _{i}(x)|^{p-2}\psi _{i}(x)u(x)\mathrm {d}x \end{aligned}$$
(2.4)

such that

$$\begin{aligned}{}[\psi _i]_{s,p}^p = \beta _{i} \end{aligned}$$
(2.5)

and

$$\begin{aligned} {\mathcal L}_i\psi _i =\int _{\mathbb {R} ^N} |\psi _{i}(x)|^{p} \mathrm {d}x = 1, \end{aligned}$$
(2.6)

hence

$$\begin{aligned} \Vert {\mathcal L}_i\Vert _{L^{p'}} = 1, \end{aligned}$$

while \({\mathcal L}_j\psi _i=0\) for all \(j\in \{1,\ldots , i-1\}\), thus

$$\begin{aligned} \psi _i \in \bigcap _{j=1}^{i-1}\ker ({\mathcal L}_j|_{\mathcal S}), \quad \hbox {if } i \ge 2. \end{aligned}$$

Therefore, we can define \({\mathcal S}_0:={\mathcal S}\),

$$\begin{aligned} {\mathcal S}_n := \{v \in {\mathcal S}: {\mathcal L}_1 v =\dots = {\mathcal L}_n v = 0\} = \bigcap _{i=1}^{n}\ker ({\mathcal L}_i|{\mathcal S}) \quad \hbox {if } n\in \mathbb {N}, \end{aligned}$$

and the corresponding constrained infimum

$$\begin{aligned} \beta _{n+1} := \inf _{v \in {\mathcal S}_n}\Phi (v) \quad \hbox {if } n \ge 0. \end{aligned}$$
(2.7)

We claim that there exists \(\psi _{n+1} \in {\mathcal S}_n\) such that

$$\begin{aligned}{}[\psi _{n+1}]_{s,p}^p = \beta _{n+1}. \end{aligned}$$
(2.8)

To this aim, arguing as above, it is enough proving that each \({\mathcal S}_n\) is weakly closed. Indeed, if \(\{v_m\}_{m\in \mathbb {N}}\subset {\mathcal S}_n\) weakly converges to \(v\in X(\Omega )\), by (1.4) and \({\mathcal L}_i \in L^{p'}\), it follows that \(|v|_p=1\) and \({\mathcal L}_i v_m \rightarrow {\mathcal L}_i v\) for all \(i \in \{1,\ldots ,n\},\) hence \(v \in {\mathcal S}_n\). Thus, \(\Phi \) attains its infimum on \({\mathcal S}_n\) and (2.8) holds.

Summing up, by induction, we construct a sequence of positive numbers \(\{\beta _n\}_{n\in \mathbb {N}}\), of functions \(\{\psi _n\}_{n\in \mathbb {N}} \subset X(\Omega )\) and of linear operators \(\{{\mathcal L}_n\}_{n\in \mathbb {N}} \subset L^{p'}\) such that (2.4)–(2.6) hold for all \(n\in \mathbb {N} \); furthermore, it is

$$\begin{aligned} 0 < \beta _1 \le \beta _2 \le \cdots \le \beta _n\le \cdots \end{aligned}$$

and \(\psi _n \ne \psi _m\) if \(n \ne m\).

Now we recall that if \(V\subseteq X\) is a closed subspace of a Banach space X, a subspace \(W\subseteq X\) is a topological complement of V, briefly \(X=V\oplus W\), if W is closed and every \(x\in X\) can be uniquely written as \(v+w\), with \(v\in V\) and \(w\in W\); furthermore, the projection operators onto V and W are linear and continuous; hence, there exists \(L:=L(V,W)>0\) such that

$$\begin{aligned} \Vert v\Vert +\Vert w\Vert \le L\Vert v+w\Vert . \end{aligned}$$
(2.9)

When \(X=V\oplus W\) and V has finite dimension, we say that W has finite codimension, with \(\mathrm{codim}\, W=\dim V\).

By using again (1.4), the proofs of [11, Lemmas 5.2 and 5.3 and Proposition 5.4] can be adapted with minor changes to our setting; thus, the following properties can be stated:

  • the increasing sequence \(\{\beta _n\}_{n\in \mathbb {N}}\) diverges positively;

  • fixing any \(n \ge 1\) and setting

    $$\begin{aligned} X_n := & {} \mathrm{span}\{\psi _1,\ldots ,\psi _n\} = \left\{ v \in X(\Omega ): \; \exists \ b_1,\ldots ,b_n \in \mathbb {R}\ \hbox {s.t.}\ v = \sum _{i=1}^n b_i\psi _i\right\} , \\ Y_n := & {} \bigcap _{i=1}^{n}\ker ({\mathcal L}_i) = \{w \in X(\Omega ): {\mathcal L}_1 w =\dots = {\mathcal L}_{n}w = 0\}, \end{aligned}$$

    we have

    $$\begin{aligned} X(\Omega ) = X_n \oplus Y_n; \end{aligned}$$
    (2.10)
  • taking \(n \ge 1\), for all \(w\in Y_n\) we get

    $$\begin{aligned} \beta _{n+1}\ \int _{\mathbb {R} ^N} |w(x)|^p \mathrm {d}x \le \ [w]_{s,p}^p; \end{aligned}$$
    (2.11)
  • the sequence \(\{\psi _n\}_{n\in \mathbb {N}}\) generates the whole space \(X(\Omega )\).

Now, following [18] we introduce another sequence of positive numbers. For all \(n\in {\mathbb {N}}\), taking \(\psi _1\) as in (2.3), we set

$$\begin{aligned} {\mathbb {W} _n}=\{Z: Z \hbox { is a subspace of } X(\Omega ), \psi _1\in Z \hbox { and } \dim Z\ge n\} \end{aligned}$$
(2.12)

and

$$\begin{aligned} \gamma _n:=\inf _{Z\in {\mathbb {W} _n}}\sup _{u\in {\mathcal S}\cap Z}\Phi (u), \end{aligned}$$
(2.13)

with \({\mathcal S}\) as in (2.1). By the previous definitions, it follows that \(\beta _1=\gamma _1\) and \({\mathbb {W} _{n+1}}\subseteq {\mathbb {W} _n}\); hence, \(\{\gamma _n\}_{n\in \mathbb {N}}\) is an increasing sequence of quasi-eigenvalues.

Remark 2.1

For \(p=2\) the sequences \(\{\beta _n\}_{n\in \mathbb {N}}\) and \(\{\gamma _n\}_{n\in \mathbb {N}}\) reduce to the known sequence of eigenvalues \(\{\lambda _n\}_{n\in \mathbb {N}}\) of \((-\varDelta )^s\) (see, for instance, the paper [29]).

By using the genus we can construct a sequence of eigenvalues \(\{\mu _n\}_{n\in \mathbb {N}}\) for the nonlinear operator \((-\varDelta )_s^p\) on \(X(\Omega )\), alike in the case of local p-Laplacian as in [16, 20].

Let us consider the nonlinear eigenvalue problem (1.5) and set

$$\begin{aligned} \mu _n \,:=\, \inf _{A\in \Sigma _n}\sup _{u\in A{\setminus }\{0\}}\frac{\Phi (u)}{\displaystyle \int _{\mathbb {R} ^N} |u(x)|^p\; \mathrm{d}x},\quad n\in \mathbb {N} \end{aligned}$$
(2.14)

where (cf. Sect. 1) \(\displaystyle \Sigma _n:=\{A\in \Sigma : \gamma (A)\ge n\}\) with \(\Phi \) as in (2.2),

$$\begin{aligned} \Sigma :=\{A \subseteq X(\Omega ), \hbox {closed and symmetric w.r.t. the origin}\} \end{aligned}$$
(2.15)

and consider the even functional

$$\begin{aligned} \Psi (u)\ :=\ \frac{\Phi (u)}{\displaystyle \int _{\mathbb {R} ^N} |u(x)|^p\; \mathrm{d}x} \quad \hbox { on } X(\Omega ){\setminus }\{0\}. \end{aligned}$$

The critical values and the critical points of \(\Psi \) restricted to the manifold \({\mathcal {S}}\) defined in (2.1) are eigenvalues and eigenfunctions of \((-\varDelta )^s_p\) on \(X(\Omega )\), respectively. We can state the following proposition.

Proposition 2.2

For every \(n\in \mathbb {N} \) the numbers \(\mu _n\) in (2.14) are eigenvalues for the nonlinear operator \((-\varDelta )^s_p\) on \(X(\Omega )\).

Proof

By [13, Lemma 4] and (1.4) the functional \(\Psi \big |_{{\mathcal {S}}}\) satisfies the Palais–Smale condition. Then, by using a suitable version of the deformation lemma (cf., e.g., [6]), standard mini–max arguments give the result. \(\square \)

Furthermore, we point out that \(\mu _1=\beta _1=\gamma _1\). Slight changes in the proof of [2, Proposition 2.9] and in [18, Remark 1.1(4)] provide the following proposition stating a comparison among the sequence \(\{\mu _n\}_{n\in \mathbb {N}}\) and the sequences of quasi-eigenvalues \(\{\beta _n\}_{n\in \mathbb {N}}\) and \(\{\gamma _n\}_{n\in \mathbb {N}}\) of \((-\varDelta )^s_p\).

Proposition 2.3

For all \(n\in \mathbb {N} \) we have that \(\beta _n\le \mu _n\le \gamma _n\).

Remark 2.4

The properties of \(\{\beta _n\}_{n\in \mathbb {N}}\) and Proposition 2.3 imply that \(\{\mu _n\}_{n\in \mathbb {N}}\) and \(\{\gamma _n\}_{n\in \mathbb {N}}\) are diverging sequences. Moreover, as \(\{\gamma _n\}_{n\in \mathbb {N}}\) is increasing, we have also \(\beta _h\le \gamma _k\) for \(k\ge h\ge 1\); therefore, this inequality is not an assumption in \((h_4)\).

3 Proof of Theorem 1.1

From \((h_1)\) and (1.2) for all \(\varepsilon >0\) there exists \(K_\varepsilon >0\) such that

$$\begin{aligned} |f(x,t)|\le \varepsilon |t|^{p-1} + K_\varepsilon \quad \hbox { for a.e. } x\in \Omega \hbox { and for all } t\in \mathbb {R}. \end{aligned}$$
(3.1)

The weak solutions of problem (1.1) are the critical points of the \(C^1\)-functional

$$\begin{aligned} J_\lambda (u):=\frac{1}{p}\Vert u\Vert ^p - \frac{{\lambda }}{p} \int _\Omega |u(x)|^p \; \mathrm{d}x - \int _\Omega F(x,u(x))\mathrm{d}x \end{aligned}$$
(3.2)

on \(X(\Omega )\) whose derivative is given by

for any \(\varphi \in X(\Omega )\).

For the sake of simplicity we introduce the operator \(A:X(\Omega )\rightarrow (X(\Omega ))^*\), defined for all \(u,\varphi \in X(\Omega )\) by

$$\begin{aligned} \langle A(u),\varphi \rangle := \iint _{\mathbb {R} ^N\times \mathbb {R} ^N}\frac{\left| u(x)-u(y)\right| ^{p-2}(u(x)-u(y))(\varphi (x)-\varphi (y))}{\left| x-y\right| ^{N+sp}}\mathrm {d}x\mathrm {d}y. \end{aligned}$$

In next proposition we prove that the functional J satisfies the \((\mathrm C)\) condition (cf. Appendix) both in the nonresonant case and in the resonant one, up to assume also assumption \((h_3)\).

Proposition 3.1

Assume that \((h_1)\)\((h_2)\) hold. Then

  1. (i)

    if \(\lambda \not \in \sigma ((-\varDelta )^s_p)\), the functional \(J_\lambda \) in (3.2) satisfies \((\mathrm C)\) in \(\mathbb {R};\)

  2. (ii)

    if \(\lambda \in \sigma ((-\varDelta )^s_p)\) and \((h_3)\) holds, the functional \(J_\lambda \) in (3.2) satisfies \((\mathrm C)\) in \(\mathbb {R} \).

Proof

(i) Let \(c\in \mathbb {R} \) and \(\{u_m\}_{m\in \mathbb {N}}\) be a sequence in \(X(\Omega )\) such that (3.30) holds; then in particular

$$\begin{aligned} \begin{aligned} \langle A(u_m),\varphi \rangle&- \lambda \int _\Omega |u_m(x)|^{p-2} u_m(x)\varphi (x)\mathrm{d}x \\&-\int _\Omega f(x,u_m(x))\varphi (x)\mathrm{d}x = o(1), \end{aligned} \end{aligned}$$
(3.3)

for every \(\varphi \in X(\Omega )\), where o(1) denotes an infinitesimal sequence.

In order to prove the statement, it is enough to show that \(\{\Vert u_m\Vert \}_{m\in \mathbb {N}}\) is bounded (cf. [23, Proposition 1.3]). Then, arguing by contradiction, let us assume that

$$\begin{aligned} \Vert u_m\Vert \rightarrow +\infty \quad \hbox { as } m\rightarrow + \infty . \end{aligned}$$
(3.4)

Setting \(w_m:=\displaystyle \frac{u_m}{\Vert u_m\Vert }\), \(\{w_m\}_{m\in \mathbb {N}}\) is bounded in \(X(\Omega )\) and there exists \(w\in X(\Omega )\) such that, up to subsequences, we have

$$\begin{aligned} w_m\rightharpoonup w \quad \hbox { weakly in } X(\Omega ) \end{aligned}$$
(3.5)

and

$$\begin{aligned} w_m\rightarrow w \quad \hbox { strongly in } L^{p}(\Omega ). \end{aligned}$$
(3.6)

Evaluating (3.3) in \(w_m-w\) and dividing by \(\Vert u_m\Vert ^{p-1}\), we get

$$\begin{aligned} \langle A(w_m),w_m-w\rangle= & {} \lambda \int _\Omega |w_m(x)|^{p-2} w_m(x)\,(w_m-w)(x) \,\mathrm{d}x \, \nonumber \\&+ \int _\Omega \frac{f(x,u_m(x))}{\Vert u_m\Vert ^{p-1}}(w_m-w)(x)\;\mathrm{d}x + o(1). \end{aligned}$$
(3.7)

Let us analyze this last equation. Firstly, by (3.6) it follows that

$$\begin{aligned} \displaystyle {\left| \int _\Omega |w_m(x)|^{p-2} w_m(x) (w_m-w)(x)\mathrm{d}x\right| \le |w_m|_p^{p-1}|w_m-w|_p=o(1).} \end{aligned}$$

Furthermore, (3.1), (3.4) and (3.6) imply that

$$\begin{aligned} \left| \int _\Omega \frac{f(x,u_m(x))}{\Vert u_m\Vert ^{p-1}}(w_m-w)(x)\mathrm{d}x \right|\le & {} \varepsilon |w_m|_p^{p-1}|w_m-w|_p \nonumber \\&+\, \frac{K_\varepsilon }{\Vert u_m\Vert ^{p-1}}|w_m-w|_1 =o(1). \end{aligned}$$
(3.8)

Hence, by (3.7)

$$\begin{aligned} \langle A(w_m),w_m-w\rangle =o(1) \end{aligned}$$

and by [23, Proposition 1.3]

$$\begin{aligned} w_m\rightarrow w \quad \hbox { strongly in } X(\Omega ). \end{aligned}$$
(3.9)

Thus, by the definition of \(w_m\) it follows \(w\not =0\).

Now, dividing (3.3) by \(\Vert u_m\Vert ^{p-1}\), for all \(\varphi \in X(\Omega )\) we have that

$$\begin{aligned} \langle A(w_m),\varphi \rangle= & {} \lambda \int _\Omega |w_m(x)|^{p-2} w_m(x)\varphi (x)\mathrm{d}x \nonumber \\&+\,\int _\Omega \frac{f(x,u_m(x))}{\Vert u_m\Vert ^{p-1}}\varphi (x)\mathrm{d}x + o(1). \end{aligned}$$
(3.10)

Again (3.1), (3.4) and (3.6) give

$$\begin{aligned} \lim _{m\rightarrow + \infty }\int _\Omega \frac{f(x,u_m(x))}{\Vert u_m\Vert ^{p-1}}\varphi (x)\mathrm{d}x =0\,\,\,\hbox { for all } \varphi \in X(\Omega ). \end{aligned}$$
(3.11)

Therefore, by (3.9) and (3.11), passing to the limit in (3.10), we get

$$\begin{aligned} \langle A(w),\varphi \rangle = \lambda \int _\Omega |w(x)|^{p-2}w(x)\varphi (x)\mathrm{d}x \quad \hbox { for all } \varphi \in X(\Omega ). \end{aligned}$$

But this means that \(\lambda \in \sigma ((-\varDelta )^s_p)\), against our assumption; thus, the proof is complete.

(ii) Let \(c\in \mathbb {R} \) and \(\{u_m\}_{m\in \mathbb {N}}\) be a sequence in \(X(\Omega )\) such that (3.30) holds. Set

$$\begin{aligned} g(x,t):=\lambda t+f(x,t) \quad \hbox { for a.e. } x\in \Omega ,\,\, \forall \, t\in \mathbb {R}. \end{aligned}$$
(3.12)

By using (3.12) we have that

$$\begin{aligned} \frac{1}{p}\Vert u_m\Vert ^p - \int _{\Omega } G(x,u_m(x))\mathrm{d}x = c+ o(1) \end{aligned}$$
(3.13)

and

$$\begin{aligned} \Vert u_m\Vert ^p- \int _{\Omega } g(x,u_m(x))u_m(x)\mathrm{d}x = o(1), \end{aligned}$$
(3.14)

with \(\displaystyle G(x,t):=\int _0^t g(x,\tau )\mathrm{d}\tau .\)

By assumption \(({h}_3)\) there exists \(\eta _1>0\) such that

$$\begin{aligned} g(x,t)t-p\,G(x,t) \ge 0 \quad \quad \hbox {if}\,|t|\ge \eta _1, \hbox {for a.e.} x\in \Omega . \end{aligned}$$
(3.15)

On the other hand, by using condition \(({h}_1)\) there exists \(C_1=C_1(\eta _1)>0\) such that

$$\begin{aligned} \int _{\{|u_m|\le \eta _1\}} \left( g(x,u_m(x))u_m(x)-p\,G(x,u_m(x))\right) \mathrm{d}x \ge -C_1, \end{aligned}$$
(3.16)

for every \(m\in \mathbb {N} \). Fixing \(\varepsilon >0\), by (3.12) in addition to (1.2) of \(({h}_2)\), there exists \(\eta _\varepsilon >0\) such that

$$\begin{aligned} |g(x,t)|\le (|\lambda | + \varepsilon ) |t| \quad \quad \hbox {if}\, |t|>\eta _\varepsilon , \hbox {for a.e.} x\in \Omega . \end{aligned}$$
(3.17)

Now, taking \(q\in ]p,p_s^*[\), there exists \(C>0\) such that

$$\begin{aligned} |u|_{q}\le C^{\frac{1}{p}}\Vert u\Vert \quad \quad \hbox {for all } u\in X(\Omega ) \end{aligned}$$
(3.18)

(cf. (1.4)). Hence, let us set

$$\begin{aligned} \kappa := (2c + C_1)(2 (|\lambda | + \varepsilon )C)^{\frac{q}{q-p}}, \end{aligned}$$
(3.19)

with c as in (3.13) and \(C_1\) as in (3.16).

Again by \(({h}_3)\) we get the existence of \(\eta _2 :=\eta _2(\kappa )>\max \{\eta _1, \eta _\varepsilon \}\) such that

$$\begin{aligned} g(x,t)t-p\,G(x,t) \ge \kappa \quad \quad \hbox {if}\,|t|\ge \eta _2, \hbox {for a.e.} x\in \Omega . \end{aligned}$$
(3.20)

Then, for \(\eta _2\) as above, we define

$$\begin{aligned} A_m:=\{x\in \Omega : |u_m(x)|\ge \eta _2\} \end{aligned}$$

and

$$\begin{aligned} B_m :=\{x\in \Omega : |u_m(x)|\le \eta _2\}, \end{aligned}$$

for every \(m\in \mathbb {N} \).

By (3.13)–(3.16) and (3.20) it follows that

$$\begin{aligned} \begin{aligned} pc + o(1)&= \int _{\Omega } \left( g(x,u_m(x))u_m(x)-p\,G(x,u_m(x))\right) \mathrm{d}x\\&= \int _{A_m} \left( g(x,u_m(x))u_m(x)-p\,G(x,u_m(x))\right) \mathrm{d}x\\&\quad + \int _{\{|u_m|\le \eta _1\}} \left( g(x,u_m(x))u_m(x)-p\,G(x,u_m(x))\right) \mathrm{d}x\\&\quad +\int _{\{\eta _1\le |u_m|\le \eta _2\}} \left( g(x,u_m(x))u_m(x)-p\,G(x,u_m(x))\right) \mathrm{d}x\\&\ge \ \kappa \, \mathrm{meas}(A_m) - C_1. \end{aligned} \end{aligned}$$

Hence, from the above inequality, one has

$$\begin{aligned} \mathrm{meas}(A_m)\le \frac{2c + C_1}{\kappa } + o(1) \quad \hbox {for all } m\in \mathbb {N}. \end{aligned}$$
(3.21)

Taking \(r>p\), by (3.13) and (3.14) we have that

$$\begin{aligned} \begin{aligned} \left( \frac{1}{p} - \frac{1}{r}\right) \Vert u_m\Vert ^p&- \int _{\Omega } \left( G(x,u_m(x)) - \frac{1}{r} \, g(x,u_m(x))u_m(x)\right) \mathrm{d}x\\&= c + o(1). \end{aligned} \end{aligned}$$
(3.22)

Moreover, by \(({h}_1)\) there exists \(C_2:=C_2(\Omega ,g,\eta _2, r)>0\) such that

$$\begin{aligned} \left| \int _{B_m} \left( G(x,u_m(x)) - \frac{1}{r}\, g(x,u_m(x))u_m(x)\right) \;\mathrm{d}x \right| \le C_2,\,\,\,\forall m\in \mathbb {N}. \end{aligned}$$
(3.23)

Hence, by (3.22) and (3.23) we infer that

$$\begin{aligned} \begin{aligned} c + o(1)&\ge \left( \frac{1}{p} - \frac{1}{r}\right) \Vert u_m\Vert ^p\\&- \int _{A_m} \left( G(x,u_m(x)) - \frac{1}{r}\, g(x,u_m(x))u_m(x)\right) \mathrm{d}x - C_2. \end{aligned} \end{aligned}$$

Further, by (3.15) and (3.17) it follows that

$$\begin{aligned} \begin{aligned} c + o(1)&\ge \ \left( \frac{1}{p} - \frac{1}{r}\right) \Vert u_m\Vert ^p\\&- \int _{A_m} \left( \frac{1}{p} g(x,u_m(x))u_m(x) - \frac{1}{r}\, g(x,u_m(x))u_m(x)\right) \;\mathrm{d}x - C_2\\&\ge \ \left( \frac{1}{p} - \frac{1}{r}\right) \left( \Vert u_m\Vert ^p - \int _{A_m} (|\lambda | +\varepsilon )|u_m(x)|^p \;\mathrm{d}x\right) - C_2. \end{aligned} \end{aligned}$$

Now, by the Hölder inequality, (3.18), (3.19) and (3.21) we have that

$$\begin{aligned} \begin{aligned} c + o(1)&\ge \left( \frac{1}{p} - \frac{1}{r}\right) \left( \Vert u_m\Vert ^p - (\lambda +\varepsilon )\ |u_m|^p_{q}\ \mathrm{meas}(A_m)^{\frac{q-p}{q}}\right) - C_2\\&\ge \left( \frac{1}{p} - \frac{1}{r}\right) \Vert u_m\Vert ^p \left( 1 - (|\lambda | + \varepsilon )C\left( \frac{1}{(2 (\lambda + \varepsilon )C)^{\frac{q}{q-p}}} + o(1)\right) ^{\frac{q-p}{q}}\right) \\&-C_2. \end{aligned} \end{aligned}$$

Thus, the sequence \(\{\Vert u_m\Vert \}_{m\in \mathbb {N}}\) is bounded in \(X(\Omega )\). \(\square \)

Lemma 3.2

Assume that \((h_1)\)\((h_2)\) hold. Let \(\beta _h\) be as in \((h_4)\) and \(Y_{h-1}\) as in (2.10). Then, there exist \(\rho >0\) and \(c_0>0\) such that, setting \(S_\rho :=\{u\in X(\Omega ): \Vert u\Vert = \rho \},\) the functional \(J_\lambda \) in (3.2) verifies

$$\begin{aligned} J_\lambda (u)\ge c_0 \quad \hbox { for all } u\in S_\rho \cap Y_{h-1}. \end{aligned}$$
(3.24)

Proof

By \((h_2)\) it follows that, uniformly with respect to almost every \(x \in \Omega \), there exist

$$\begin{aligned} \lim _{|t|\rightarrow +\infty }\frac{F(x,t)}{|t|^{p}} \ = 0 \end{aligned}$$

and

$$\begin{aligned} \lim _{t\rightarrow 0}\frac{F(x,t)}{|t|^{p}} \ = \frac{\alpha }{p}. \end{aligned}$$

Therefore, for every \(\varepsilon >0\) there exist \(R_\varepsilon , \delta _\varepsilon >0\) such that, for almost every \(x\in \Omega \),

$$\begin{aligned} |F(x,t)|\le \frac{\varepsilon }{p}|t|^p \quad \quad \hbox {if}\,|t|> R_\varepsilon \end{aligned}$$
(3.25)

and

$$\begin{aligned} \left| F(x,t)-\frac{\alpha }{p}|t|^p\right| \le \frac{\varepsilon }{p}|t|^p \quad \quad \hbox {if}\,|t|< \delta _\varepsilon , \end{aligned}$$
(3.26)

without loss of generality with \(R_\varepsilon \ge 1\). On the other hand, by \((h_1)\), taking any \(l\in [0, p_s^*-p[\), there exists \(k_{R_\varepsilon }>0\) such that, for almost every \(x\in \Omega \),

$$\begin{aligned} |F(x,t)| \le k_{R_\varepsilon }{|t|}^{l+p} \quad \quad \hbox {if}\,\delta _\varepsilon \le |t|\le R_\varepsilon . \end{aligned}$$
(3.27)

The inequalities (3.25)–(3.27) imply that for any \(\varepsilon >0\) there exists \(k_\varepsilon >0\) such that

$$\begin{aligned} F(x,t)\le \frac{\alpha + \varepsilon }{p}|t|^p + {k_\varepsilon }|t|^{l+p} \quad \quad \hbox {for a.e.} x\in \Omega , \hbox {for all}\,t\in \mathbb {R}. \end{aligned}$$

We infer that

$$\begin{aligned} \int _\Omega F(x,u(x))\mathrm{d}x\le \frac{\alpha + \varepsilon }{p}|u|^p_p + {k_\varepsilon }|u|^{l+p}_{l+p} \quad \hbox {for all } u\in X(\Omega ). \end{aligned}$$

For a suitable \(k'_\varepsilon >0\) we have

$$\begin{aligned} J_\lambda (u)\ge \frac{1}{p}\Vert u\Vert ^p - \frac{\lambda + \alpha + \varepsilon }{p}|u|_p^p - k_\varepsilon ' \Vert u\Vert ^{l+p}\quad \hbox {for all } u\in X(\Omega ). \end{aligned}$$
(3.28)

Let us recall that by the decomposition (2.10) it is \(X(\Omega )=X_{h-1}\oplus Y_{h-1}\), where \(X_{h-1}:=\mathrm{span}\{\psi _1,\ldots ,\psi _{h-1}\}\) and \(Y_{h-1}\) is its complement. Thus by (2.11) and (3.28) it follows that

$$\begin{aligned} J_\lambda (u)\ge \frac{1}{p} \left( 1 - \frac{\lambda + \alpha + \varepsilon }{\beta _{h}}\right) \Vert u\Vert ^p - k_\varepsilon ' \Vert u\Vert ^{l+p}\quad \hbox {for all}\,u\in Y_{h-1} \end{aligned}$$

and by \((h_4)\), for a suitable \(\varepsilon \), there exists \(k''_\varepsilon >0\) such that

$$\begin{aligned} J_\lambda (u)\ge k''_\varepsilon \Vert u\Vert ^p- k_\varepsilon '\Vert u\Vert ^{l+p} \quad \hbox {for all } u\in Y_{h-1}. \end{aligned}$$

Thus we conclude that if \(\rho \) is small enough there exists \(c_0>0\) such that (3.24) holds. \(\square \)

Lemma 3.3

Assume that \((h_1)\) and (1.2) hold. Let \(\gamma _k\) as in \((h_4)\), \(\mathbb {W} _{k}\) as in (2.12) and \(c_0\) as in Lemma 3.2. Then, there exist a k-dimensional space \(V\in \mathbb {W} _k\) and \(c_\infty >c_0\) such that the functional \(J_\lambda \) in (3.2) verifies

$$\begin{aligned} J_\lambda (u)\le c_\infty \quad \hbox { for all } u\in V. \end{aligned}$$
(3.29)

Proof

By (3.1), fixing any \(\varepsilon >0\) there exists \(C_\varepsilon >0\) such that

$$\begin{aligned} J_\lambda (u)\le \frac{1}{p}\Vert u\Vert ^p -\frac{\lambda }{p}|u|^p + \frac{\varepsilon }{2p}|u|^p_p + C_\varepsilon |u|_p \quad \hbox {for all } u\in X(\Omega ). \end{aligned}$$

Let \(\gamma _k\) be as in \((h_4)\) and take \(\varepsilon >0\) such that \(\gamma _k+\varepsilon <\lambda \). From definition (2.13), for such a fixed \(\varepsilon >0\) there exists a subspace \(V^\varepsilon _k\) in \(\mathbb {W} _k\), with \(\dim V^\varepsilon _k\ge k\), such that

$$\begin{aligned} \gamma _k\le \sup _{u\in V^\varepsilon _k{\setminus }\{0\}}\frac{\Vert u\Vert ^p}{|u|^p}<\gamma _k+\frac{\varepsilon }{2}. \end{aligned}$$

Thus it results that

$$\begin{aligned} J_\lambda (u)\le \frac{1}{p}\left( \gamma _k+\varepsilon - \lambda \right) |u|^p + C_\varepsilon |u|_p \quad \hbox {for all } u\in V^\varepsilon _k \end{aligned}$$

and, as without loss of generality we can assume that \(V^\varepsilon _k\) is a k-dimensional subspace, the functional \(J_\lambda \) tends to \(-\infty \) as \(\Vert u\Vert \) diverges in \(V^\varepsilon _k\), so there exists \(c_\infty =c_\infty (\varepsilon )\) (with \(c_\infty >c_0\)), such that (3.29) holds. \(\square \)

Proof of Theorem 1.1

(a) Firstly, by Proposition 3.1—part (i) the functional \(J_\lambda \) in (3.2) satisfies \((\mathrm C)\) in \(\mathbb {R} \), and by assumption, it is even.

Let us consider \(\beta _h\), \(Y_{h-1}\), \(\rho , c_0\) as in Lemma 3.2 and \(\gamma _k, \mathbb {W} _{k}\), \(V_k^\varepsilon \), \(c_\infty \) as in Lemma 3.3.

Then, we consider the pseudo-index theory \((S_\rho \cap Y_{h-1}, {\mathcal {H}}^*, \gamma ^*)\) related to the genus and \(S_\rho \cap Y_{h-1}\). By Remark 3.7 applied to \(V:=V^\varepsilon _k\), \(\partial B:= S_\rho \) and \(W:=Y_{h-1}\), we get

$$\begin{aligned} \gamma \left( V^\varepsilon _k\cap h\left( S_\rho \cap Y_{h-1}\right) \right) \ge \dim V^\varepsilon _k - \text{ codim } Y_{h-1} \quad \hbox {for all } h\in {\mathcal {H}}^*, \end{aligned}$$

which implies

$$\begin{aligned} \gamma ^*(V^\varepsilon _k)\ge k-h+1. \end{aligned}$$

The proof is then complete: in fact Theorem 3.6 applies with \({\tilde{A}}:= V^\varepsilon _k\) and \(S:=S_\rho \cap Y_{h-1}\) and J has at least \(k-h+1\) distinct pairs of critical points corresponding to at most \(k-h+1\) distinct critical values \(c_i\), where \(c_i\) is as in (3.32).

(b) In the resonant case, by Proposition 3.1—part (ii) the functional \(J_\lambda \) satisfies \((\mathrm C)\), and we can proceed as above. \(\square \)

Remark 3.4

We point out that Theorem 1.1 holds also with slight changes in the proof when \((h_4)\) is replaced by

$$\begin{aligned} \lambda<\ \beta _h\le \ \gamma _k\ < \alpha + \lambda . \end{aligned}$$