1. Introduction
1.1. Features of the paper and historical comments
In this paper, we are concerned with the study of concentration and multiplicity properties of solutions for a class of fractional double phase problems with Choquard nonlinearity. The features of this paper are the following:
(i) the presence of several nonlocal operators with different growth, which generates a double phase associated energy;
(ii) the analysis developed in this paper is concerned with the combined effects of a nonstandard fractional operator with unbalanced growth and a nonlocal Choquard term;
(iii) the potential describing the absorption term satisfies a local condition and no information on the behaviour of the potential at infinity is imposed;
(iv) the main concentration property creates a bridge between the global maximum point of the solution and the global minimum of the potential;
(v) since the nonlinearity is only assumed to be continuous, one cannot apply the standard C 1-Nehari manifold arguments due to the lack of differentiability of the associated Nehari manifold;
(vi) our analysis combines the nonlocal nature of the fractional (p, q)-operator and of the Choquard nonlinearity with the local perturbation in the absorption term.
The problem studied in this paper combines both the above-mentioned features. More exactly, we are interested in concentration phenomena associated with a nonlinear Choquard problem driven by fractional Laplace operators with different power. This integro-differential operator appears in several relevant models of applied sciences. We only recall that the fractional power of the Laplace operator is the infinitesimal generator of the Lévy stable diffusion process and arises in anomalous diffusion in plasma, population dynamics, geophysical fluid dynamics, flame propagation, chemical reactions in liquids, and American options in finance. Moreover, fractional Sobolev spaces have been well known since the beginning of the last century, especially within the framework of harmonic analysis. The starting point in the study of fractional problems is attributed to the pioneering papers of Caffarelli et al. [Reference Caffarelli, Roquejoffre and Sire20–Reference Caffarelli and Silvestre22]. For a comprehensive introduction to the study of fractional equations and the use of variational methods in the treatment of these problems, we refer to the monographs by Di Nezza, Palatucci, and Valdinoci [Reference Di Nezza, Palatucci and Valdinoci27] and Molica Bisci, Rădulescu, and Servadei [Reference Molica Bisci, Rădulescu and Servadei45]. Therefore, the nonlocal operators are becoming increasingly popular in applied sciences, theoretical research, and real-world applications.
Since the content of this paper is at interplay between ‘double phase problems’ and ‘Choquard problems’, we now recall some pioneering achievements in these fields.
We start with a short description on the development of double phase problems. To the best of our knowledge, the first contributions to this field are due to the study by Ball [Reference Ball13, Reference Ball14], in relationship with problems in nonlinear elasticity and composite materials. Let $\Omega\subset\mathbb{R}^N$ be a bounded domain with a smooth boundary. If $u:\Omega\to\mathbb{R}^N$ is the displacement and if Du is the N × N matrix of the deformation gradient, then the total energy can be represented by an integral of the type
where the energy function $f=f(x,\xi):\Omega\times\mathbb{R}^{N\times N}\to\mathbb{R}$ is quasiconvex with respect to ξ, see the study by Morrey [Reference Morrey47]. A simple example considered by Ball is given by functions f of the type
where ${\mathrm{det}}\,\xi$ is the determinant of the N × N matrix ξ and g, h are non-negative convex functions, which satisfy the growth conditions
where c 1 is a positive constant and $1 \lt p \lt N$. The condition $p\leqslant N$ is necessary to study the existence of equilibrium solutions with cavities, that is, minima of the variational integral (1.1) that are discontinuous at one point where a cavity forms; in fact, every u with finite energy belongs to the Sobolev space $W^{1,p}(\Omega,\mathbb{R}^N)$, and thus it is a continuous function if p > N. In accordance with these problems arising in nonlinear elasticity, Marcellini [Reference Marcellini41, Reference Marcellini42] considered continuous functions $f=f(x,u)$ with unbalanced growth that satisfy
where c 1, c 2 are positive constants and $1\leqslant p\leqslant q$. Recently, a great deal of works have enriched the mathematical analysis of nonlinear models with unbalanced growth; we refer to the works of Bahrouni, Rădulescu, and Repovš [Reference Bahrouni, Rădulescu and Repovš12]; Crespo-Blanco, Gasínski, Harjulehto; and Winkert [Reference Crespo-Blanco, Gasínski, Harjulehto and Winkert23]; Liu and Papageorgiou [Reference Liu and Papageorgiou40]; and Papageorgiou, Pudełko, Rădulescu [Reference Papageorgiou, Pudełko and Rădulescu49]. In addition, an overview of recent developments on the regularity of nonlocal double problems can be found in Byun, Ok, and Kyeong [Reference Byun, Ok and Kyeong19] and De Filippis and Palatucci [Reference De Filippis and Palatucci25].
In the pioneering works of Fröhlich [Reference Fröhlich30] and Pekar [Reference Pekar50], the following Choquard equation was introduced,
which models quantum polaron and corresponds to the study of the free electrons in ionic lattices interacting with phonons associated with deformations of the lattices or with the polarization created on the medium. Choquard also used it to study steady states of the one component plasma approximation in the Hartree–Fock theory describing an electron trapped in its own hole [Reference Lieb37]. Equation (1.2) is also called the Schrödinger–Newton equation which combines the Schrödinger equation of quantum physics with nonrelativistic Newtonian gravity. Such a model can also be deduced from the Einstein–Klein–Gordon and Einstein–Dirac systems related to boson stars and the collapse of galaxy fluctuations of scalar field dark matter; we bring the reader’s attention to Elgart and Schlein [Reference Elgart and Schlein29], Jones [Reference Jones36], Lions [Reference Lions39], Penrose [Reference Penrose51, Reference Penrose56], and Schunck and Mielke [Reference Schunck and Mielke54].
In conclusion, based on the historical research background of the fractional double problem and the Choquard problem and the related theoretical foundation, focusing on the characteristics of such problems and the combined effects produced by their combination, we will apply analytical, topological, and variational methods and develop some new techniques to devote ourselves to the study of new phenomena of nonlocal double problems with nonlocal reaction terms, when the potential V satisfies a general local condition and the nonlinearity f possesses only the continuity property.
1.2. Statement of the problem and main result
In this paper, we focus on the existence, multiplicity, and concentration behaviour of positive solutions for the nonlinear fractional (p, q)-Choquard problem:
where ɛ is a positive parameter, $0 \lt s \lt 1$, $2 \leqslant p \lt q \lt \min\{2p, N / s\}$, $0 \lt \mu \lt sp$, $(- \Delta)_t^s$ $(t \in \left\{p,q\right\})$ is the fractional t-Laplace operator, $V : \mathbb{R}^N \mapsto \mathbb{R}$ and $f : \mathbb{R} \mapsto \mathbb{R}$ are continuous functions, F is the primitive function of f, and $*$ denotes the convolution product.
In the local setting corresponding to $s \rightarrow 1^-$ (up to normalization), the double phase problem (1.3) is motivated by numerous models arising in mathematical physics. For instance, we can refer to the following Born–Infeld equation [Reference Bonheure, d’Avenia and Pomponio16] that appears in electromagnetism:
Indeed, by the Taylor formula, we have
Taking $x=2\vert\text{D}u\vert^2$ and adopting the first-order approximation, we get a particular case of the problem (1.3) for p = 2 and q = 4. Furthermore, the n-th order approximation problem is driven by the multi-phase differential operator
Finally, we also refer to the following fourth-order relativistic operator:
which describes large classes of phenomena arising in relativistic quantum mechanics. Clearly, we can use the Taylor formula to conclude that the fourth-order relativistic operator can be approximated by the following double phase operator:
The purpose of this paper is to study a class of fractional unbalanced double phase problems in which the nonlocal term appears also in the nonlinear part. Problem (1.3) is closely related to the analysis of nonlinear problems and stationary waves for models arising in mathematical physics, such as phase transitions, anomalous diffusion, composite materials, image processing, fractional quantum mechanics in the study of particles on stochastic fields, fractional superdiffusion, and fractional white-noise limit, see the study by Pucci and Saldi [Reference Pucci and Saldi52]. For more details, we refer interested readers to the preliminary introduction of this topic in [Reference Di Nezza, Palatucci and Valdinoci27, Reference Molica Bisci, Rădulescu and Servadei45]. From the point of view of physics, in the semi-classical sense (namely, as $ \varepsilon \rightarrow 0$), it is of great significance to study the existence and shape of the standing wave solutions of problem (1.3), which can be used to describe the transition relationship between quantum mechanics and classical mechanics. From a mathematical point of view, in the framework of this semi-classical state, we can obtain more dynamic behaviour information, such as concentration, convergence, decay, bifurcation, and other properties.
Note that, if $p=q=2$, after rescaling, problem (1.3) reduces to the following nonlinear fractional Choquard equation:
When ɛ = 1, $F(u)=|u|^p$, and $f(u)=|u|^{p-2}u$, d’Avenia, Siciliano and Squassina [Reference d’Avenia, Siciliano and Squassina24] dealt with the regularity, existence and non-existence, symmetry, and decay properties of solutions to problem (1.4). Under assumption that the potential V has a local minimum, Ambrosio [Reference Ambrosio9] established the multiplicity and concentration of positive solutions to problem (1.4) via the penalization technique and Ljusternik-Schnirelmann theory.
When $p=q \neq 2$, after rescaling, problem (1.3) boils down to the following nonlinear fractional Choquard problem:
In [Reference Ambrosio7], Ambrosio investigated the existence, multiplicity, and concentration of positive solutions for the fractional Choquard problem (1.5) by assuming that the potential $V :\mathbb{R}^N \mapsto \mathbb{R}$ fulfills the following condition due to the study by Rabinowitz [Reference Rabinowitz53]:
(V) $V_\infty=\liminf_{|x| \rightarrow +\infty} V(x) \gt \inf_{x\in\mathbb{R}^N}V(x)$.
These solutions concentrate at global minimum points of V under the global hypothesis (V). In the local sense s = 1, problem (1.5) becomes the following quasi-linear Choquard problem with the p-Laplace operator
for which several existence, multiplicity, and concentration results of positive solutions have been studied by different authors, under suitable assumptions on the potential function V and the reaction f; see, for instance, Alves and Yang [Reference Alves and Yang3–Reference Alves and Yang5]. On the other hand, if s = 1 and $1 \lt p \lt q \lt N$, after rescaling, problem (1.3) reduces to the following (p, q)-Laplace problem:
Very recently, by assuming that the potential V and the nonlinear reaction f satisfy the following conditions:
$(\overline{V})$ $V \in C(\mathbb{R}^N,\mathbb{R})$ and $+ \infty \gt V_\infty=\liminf_{|x| \rightarrow +\infty} V(x) \gt \inf_{x\in\mathbb{R}^N}V(x) \gt 0$;
(f)$~~f \in C^1(\mathbb{R},\mathbb{R})$,
Zhang, Zhang, and Rădulescu studied the multiplicity and concentration behavior of positive solutions to problem (1.7) in [Reference Zhang, Zhang and Rădulescu60].
For related concentration and multiplicity properties of solutions, we refer to the recent paper by Alves and de Morais Filho [Reference Alves and de Morais Filho2]; Alves, Ambrosio, and Isernia [Reference Alves, Ambrosio and Isernia1]; Ambrosio [Reference Ambrosio8]; Ambrosio, Isernia, and Rădulescu [Reference Ambrosio, Isernia and Rădulescu10]; Ambrosio and Rădulescu [Reference Ambrosio and Rădulescu11]; Del Pino and Felmer [Reference Del Pino and Felmer26]; Gao, Tang, and Chen [Reference Gao, Tang and Chen31]; Gu and Tang [Reference Gu and Tang32]; Ji and Rădulescu [Reference Ji and Rădulescu34, Reference Ji and Rădulescu35]; Moroz and Van Schaftingen [Reference Moroz and Van Schaftingen46]; Zhang and Zhang [Reference Zhang and Zhang61]; Zhang, Zhang, and Rădulescu [Reference Zhang, Zhang and Rădulescu59]; and Zhang, Tang, and Rădulescu [Reference Zhang, Tang and Rădulescu62]. We also cite the study by Mingione and Rădulescu [Reference Mingione and Rădulescu44] for an overview of recent results concerning elliptic variational problems with nonstandard growth and nonuniform ellipticity.
Inspired by the above works, in this paper, we are interested in studying the multiplicity and concentration behaviour as $\varepsilon \rightarrow 0$ of positive solutions for problem (1.3), when we suppose a local condition on the potential V, and the nonlinear term f is only assumed to be continuous (without C 1 differentiability). Precisely, we first impose that V is a continuous map satisfying the following assumptions:
there exists $V_0 \gt 0$ such that $V_0 := \inf_{x \in \mathbb{R}^N} V(x)$;
there exists an open bounded domain $\Lambda \subset \mathbb{R}^N$ such that
\begin{align*}V_0 \lt \min_{\partial \Lambda}V\quad \ \text{and }\quad M :=\{x \in \Lambda : V(x) = V_0\} \neq \emptyset.\end{align*}
Next, we assume that f is a merely continuous function that verifies the following conditions:
(f1) $\lim_{t \rightarrow 0} \frac{|f(t)|}{|t|^{p - 1}} = 0$;
(f2) there exists $\nu \in (q, q_s^*(2 - \mu / N) / 2)$ such that
\begin{align*}\lim_{|t| \rightarrow + \infty} \frac{|f(t)|}{|t|^{\nu - 1}} = 0,\end{align*}where $q_s^* = Nq / (N - sq)$;
(f3) $0 \lt p F(t) : = p \int^t_0 f(\tau) d\tau \leqslant f(t)t$ for all t > 0;
(f4) the map $t \mapsto \frac{f(t)}{t^{q - 1}}$ is increasing for all t > 0.
Remark 1.1. We would like to point out that, since the hypotheses on V and f are different from [Reference Ambrosio7, Reference Rabinowitz53, Reference Zhang, Zhang and Rădulescu59–Reference Zhang and Zhang61], our arguments are totally distinct with them and improve the previous results for the singularly perturbed fractional problem with nonlocal Choquard reaction, in the sense that we establish multiplicity results and concentration behaviour for the fractional (p, q)-problems involving continuous nonlinearity and by imposing a local condition on the potential V. Compared with the local case s =1, we point out that our result improves theorem 1.2 in [Reference Zhang, Zhang and Rădulescu60]. In addition, we believe that the ideas contained here can be used in other cases to study problems driven by more general operators, under local conditions on the potential V and without the differentiability of the nonlinearity f.
In order to look for positive solutions of problem (1.3), we may assume that $f(t) = 0$ for all $t \leqslant 0$.
Recall that if A is a closed subset of a topological space Y, then we use ${\mathrm{cat}}_Y(A)$ to denote the Ljusternik–Schnirelmann category of A in Y, that is, the smallest number of closed and contractible sets in Y which cover A. For more details, we refer to the study by Willem [Reference Willem57].
Our main result in this paper establishes the following concentration and multiplicity properties.
Theorem 1.2 Let $0 \lt \mu \lt sp.$ Assume that the nonlinearity f fulfils $(f_1)$–$(f_4)$ with $\nu \lt (N - \mu) q / (N - sq)$ and the potential V verifies hypotheses $(V_1)$–$(V_2).$ Then for all δ > 0 such that $M_\delta :=\left\{x \in \mathbb{R}^N : {\mathrm{dist}\,}(x,M) \leqslant \delta\right\} \subset \Lambda,$ there exists $\varepsilon_\delta \gt 0$ such that, for any $\varepsilon \in (0,\varepsilon_\delta)$ problem (1.3) has at least $cat_{M_\delta}(M)$ positive solutions. Moreover, if uɛ denotes one of these solutions and $x_\varepsilon \in \mathbb{R}^N$ is the global maximum point of $u_\varepsilon,$ then $\lim_{\varepsilon \rightarrow 0} V(\varepsilon x_\varepsilon) = V_0.$
The proof of theorem 1.2 is based on topological and variational methods and refined analytic techniques. Let us now take the next steps to outline our strategies and methods for proving theorem 1.2:
(1) concerning our variational approach: because there is no information on the behaviour of V at infinity, as in [Reference Del Pino and Felmer26], we first modify in a convenient way the nonlinear reaction outside the set Λ, and then we will study an auxiliary problem. The main characteristic of the corresponding modified energy functional is that it verifies all the conditions of the mountain pass theorem [Reference Ambrosetti and Rabinowitz6]. Finally, we prove that, for ɛ > 0 sufficiently small, the solutions of the auxiliary problem are indeed solutions of the original one;
(2) topological techniques: in order to get multiple solutions of the auxiliary problem, we will use the generalized Neahri manifold method and some well-known topological techniques proposed by Benci and Cerami in [Reference Benci and Cerami15] based on accurate comparisons between the category of some sub-level sets of the modified functional and the category of the set M. Note that the nonlinearity is only continuous, and we stress that standard C 1-Nehari manifold arguments as in [Reference Zhang, Zhang and Rădulescu60] cannot be employed in our setting due to the lack of differentiability of the associated Nehari manifold. To this end, we take inspiration by some ideas developed in [Reference Szulkin, Weth, Gao and Motreanu55] and make use of some abstract critical point results obtained in [Reference Szulkin, Weth, Gao and Motreanu55] to overcome this obstacle;
(3) the combination effects of nonlocality and nonhomogeneous: the lack of homogeneity caused by fractional (p, q)-Laplacian operator, together with the appearance of nonlocal Choquard reaction term, makes our analysis more refined and interesting than the above works and also brings some new difficulties to our analysis. In particular, we need to develop a suitable Moser iteration scheme to obtain $L^\infty$-estimates and absorb some ideas from [Reference Ambrosio and Rădulescu11, Reference Zhang, Tang and Rădulescu62] to build on the Hölder regularity results under this work.
Throughout this paper, C, C 0, C 1, C 2, ... denote positive constants whose exact values are inessential and can change from line to line, and the same C, C 0, C 1, C 2, ... may represent different constants; $B_\rho(y)$ denotes the open ball centred at $y \in \mathbb{R}^N$ with radius ρ > 0, and $B_\rho^c(y)$ denotes the complement of $B_\rho(y)$ in $\mathbb{R}^N$. In particular, Bρ and $B_\rho^c$ denote $B_\rho(0)$ and $B_\rho^c(0)$, respectively.
2. Auxiliary results
2.1. Notations
Let $u : \mathbb{R}^N \mapsto \mathbb{R}$. For $0 \lt s \lt 1$ and p > 1, we define the fractional Sobolev space
equipped with the natural norm
For all u, $v \in W^{s,p}(\mathbb{R}^N)$, we define
We now recall some fundamental embeddings. Let $s \in (0,1)$ and $p \in (1, + \infty)$ such that N > sp. Then for any $r \in [p,p_s^*]$ there exists a constant such that $C_r \gt 0$ such that
for all $u \in W^{s,p}(\mathbb{R}^N)$. Moreover, $W^{s,p}(\mathbb{R}^N)$ is compactly embedded in $L^r_{\mathrm{loc}}(\mathbb{R}^N)$ for any $r \in [1,p_s^*)$.
In order to treat the nonlinear fractional (p, q)-Choquard problem, we use the following space:
endowed with the norm
Additionally, since $W^{s,r}(\mathbb{R}^N)$ ($1 \lt r \lt + \infty$) is a separable reflexive Banach space, then X is also a separable reflexive Banach space.
2.2. The penalized problem
In order to overcome the lack of compactness of problem (1.3), we shall adapt the penalization method introduced by del Pino and Felmer [Reference Del Pino and Felmer26] to deal with the nonlinear fractional (p, q)-Choquard problem. Furthermore, without loss of generality, we can assume that
So, we need to find a constant K > 0 (which is determined later, see lemma 2.6) and take a unique number a > 0 such that $f(a) = \frac{V_0}{K} \left(a^{p - 1} + a^{q - 1}\right)$.
Now, we define
and
where $\chi_\Omega$ is the characteristic function on $\Omega \subset \mathbb{R}^N$.
We can easily observe that g is a Carathéodory function and fulfills the following properties:
(g1) $\lim_{t \rightarrow 0^+} \frac{g(x,t)}{t^{p - 1}} = 0$ uniformly for all $x \in \mathbb{R}^N$;
(g2) $g(x,t) \leqslant f(t)$ for all $x \in \mathbb{R}^N$ and t > 0;
(g3)i $0 \lt p G(x,t) := p \int^t_0 g(x,\tau) d\tau \leqslant g(x,t) t$ for all $x \in \Lambda$ and t > 0;
(g3)ii $0 \lt p G(x,t) \leqslant g(x,t) t \leqslant \frac{V_0}{K} \left(t^p + t^q\right)$ for all $x \in \Lambda^c$ and t > 0;
(g4) the maps $t \mapsto \frac{g(x,t)}{t^{\frac{q}{2} - 1}}$ and $t \mapsto \frac{G(x,t)}{t^\frac{q}{2}}$ are both increasing for all $x \in \mathbb{R}^N$ and t > 0.
Remark 2.1. We shall consider the following penalized problem:
in $\mathbb{R}^N.$ If uɛ is a solution of problem (2.2) such that $u_\varepsilon(x) \leqslant a$ for all $x \in \Lambda_\varepsilon^c,$ where $\Lambda_\varepsilon:=\left\{x \in \mathbb{R}^N : \varepsilon x \in \Lambda \right\},$ then $g(\varepsilon x, u_\varepsilon) = f(u_\varepsilon),$ $G(\varepsilon x, u_\varepsilon) = F(\varepsilon x,u_\varepsilon).$ So, uɛ is also a solution of problem (1.3).
For any $\varepsilon \geqslant 0$, we define the space
equipped with the norm
where $\|u\|_{V_\varepsilon,t}^t := [u]_{s,t}^t + \int_{\mathbb{R}^N} V(\varepsilon x) |u|^t dx$ for $t \in \left\{p,q\right\}$.
From now on, we focus on the critical points of the Euler–Lagrange functional $J_\varepsilon : {X_\varepsilon} \mapsto \mathbb{R}$ defined by
for all $u \in {X_\varepsilon}$. By a standard argument, we can infer that $J_\varepsilon \in C^1({X_\varepsilon},\mathbb{R})$ and its derivative is given by
Let us define the Nehari manifold associated with problem (2.2), that is,
We use ${X_\varepsilon^+}$ to denote the open set defined by
and we introduce the set $S_\varepsilon^+ := S_\varepsilon \cap {X_\varepsilon^+}$, where $S_\varepsilon : = \left\{u \in {X_\varepsilon} : \|u\|_{X_\varepsilon} = 1\right\}$. We first observe that $S_\varepsilon^+$ is an incomplete $C^{1,1}$-manifold of codimension one. So, for all $u \in S_\varepsilon^+$ we have $X_\varepsilon = T_u (S_\varepsilon^+) \bigoplus \mathbb{R} u$, where
Due to the fact that f is only continuous, the following result is crucial to bypass the non-differentiability of $\mathcal{N}_\varepsilon$.
Lemma 2.2. Assume that $(f_1)$–$(f_4)$ and $(V_1)$–$(V_2)$ are fulfilled. Then the following properties hold true:
(a) for all $u \in {X_\varepsilon^+},$ there exists a unique $t_u \gt 0$ such that $t_u u \in \mathcal{N}_\varepsilon.$ Furthermore, $\hat{m}_\varepsilon(u) = t_u u$ is the unique maximum of $\ell_u(t) :=J_\varepsilon(t u);$
(b) there exists τ > 0 independent of u such that $t_u \geqslant \tau$ for all $u \in S_\varepsilon^+.$ Moreover, for each compact set $\mathcal{W} \subset S_\varepsilon^+,$ there exists a constant $C_\mathcal{W} \gt 0$ such that $t_u \leqslant C_\mathcal{W}$ for all $u \in \mathcal{W};$
(c) the mapping $\hat{m}_\varepsilon : {X_\varepsilon^+} \mapsto \mathcal{N}_\varepsilon$ is continuous and $m_\varepsilon :=\hat{m}_\varepsilon |_{S_\varepsilon^+}$ is a homeomorphism between $S_\varepsilon^+$ and $\mathcal{N}_\varepsilon,$ and the inverse of mɛ is given by $m_\varepsilon^{-1}(u) := u / \|u\|_{X_\varepsilon};$
(d) $c_\varepsilon : = \inf_{u \in \mathcal{N}_\varepsilon} J_\varepsilon(u) \geqslant \varrho_0 \gt 0$ and Jɛ is bounded below $\mathcal{N}_\varepsilon,$ where ϱ 0 is independent of $\varepsilon,$ K, and $a;$
(e) let $0 \lt \mu \lt sp$ and $\nu \lt (N - \mu) q / (N - sq).$ If there exists a sequence $\left\{u_n\right\}_{n \in \mathbb{N}} \subset S_\varepsilon^+$ such that ${\mathrm{dist}\,}(u_n,\partial S_\varepsilon^+) \rightarrow 0$ as $n \rightarrow \infty,$ then $\|m_\varepsilon(u_n)\|_{X_\varepsilon} \rightarrow + \infty$ and $J_\varepsilon(m_\varepsilon(u_n)) \rightarrow + \infty$ as $n \rightarrow \infty.$
Proof. (a) For each $u \in {X_\varepsilon^+}$ and t > 0, $\ell_u(0) = 0$. From theorem 4.3 of Lieb and Loss [Reference Lieb and Loss38], $(g_2)$, $(f_1)$–$(f_2)$, (2.1), and hypothesis $(V_1)$ we can deduce that there exists some constant C > 0 such that
Due to $2p \gt q$, we see that $\ell_u(t) \gt 0$ for t > 0 sufficiently small. Using $(g_3)_i$ and $(g_3)_{ii}$, we can find a constant
such that
for all t > 1. Applying $2p \gt q$ again, we know that $\ell_u(t) \lt 0$ for t > 1 large enough. Hence, $\max_{t \geqslant 0} \ell_u(t)$ is attained at some $t_u \gt 0$ verifying $\ell_u'(t_u)= 0$ and $t_u u \in \mathcal{N}_\varepsilon$. We point out that
According to q > p and $(g_4)$, we conclude that the right-hand side in (2.3) is an increasing function with respect to t > 0. Therefore, the uniqueness of tu is now obvious.
(b) For any $u \in S_\varepsilon^+$, in view of lemma 2.2-(a), there exists $t_u \gt 0$ such that
By theorem 4.3 of Lieb and Loss [Reference Lieb and Loss38], together with $(g_2)$, $(f_1)$, $(f_2)$, (2.1), and hypothesis $(V_1)$, we can infer that for any σ > 0, there exists some constant $C_\sigma \gt 0$ such that for all $u \in S_\varepsilon^+$,
Assume that $t_u \leqslant 1$. Choosing $\sigma = \frac{1}{2}$, we have
Assume that $t_u \gt 1$. Taking σ = 1 and applying $1 = \|u\|_{X_\varepsilon} \geqslant \|u\|_{V_\varepsilon,p}$, we get
So, there exists τ > 0 independent of u such that $t_u \geqslant \tau$ for all $u \in S_\varepsilon^+.$
Let $\mathcal{W} \subset S_\varepsilon^+$ be a compact set. Arguing by contradiction, we may assume that there exists a sequence $\{u_n\}_{n \in \mathbb{N}} \subset \mathcal{W}$ such that $1 \leqslant t_n :=t_{u_n} \rightarrow + \infty$ as $n \rightarrow \infty$. Since $\mathcal{W}$ is a compact set, there is $u \in \mathcal{W}$ such that $u_n \rightarrow u$ in Xɛ as $n \rightarrow \infty$. As in the proof of lemma 2.2-(a), together with Fatou’s Lemma, we can see that
In addition, for any fixed $\varphi \in \mathcal{N}_\varepsilon$, we have $\langle J_\varepsilon'(\varphi), \varphi \rangle = 0$. Combining $(g_3)_i$ with $(g_3)_{ii}$, we have
Taking $\varphi_n = t_n u_n \in \mathcal{N}_\varepsilon$ in the above inequality, we get
This is a contradiction.
(c) As in lemma 2.2-(a), we can define the maps $\hat{m}_\varepsilon : {X_\varepsilon^+} \mapsto \mathcal{N}_\varepsilon$ and $m_\varepsilon : S_\varepsilon^+ \mapsto \mathcal{N}_\varepsilon $ by
Firstly, we note that $\hat{m}_\varepsilon$, mɛ, and $m_\varepsilon^{- 1}$ are well-defined. Indeed, using lemma 2.2 (a), for any fixed $u \in {X_\varepsilon^+}$ it follows that there exists a unique $\hat{m}_\varepsilon(u) \in \mathcal{N}_\varepsilon$. In addition, if $u \in \mathcal{N}_\varepsilon$, and so $u \in {X_\varepsilon^+}$. Otherwise, we obtain $|{\mathrm{supp}\,}(u^+)| = 0$. The above equality, hypothesis $(V_1)$, the definition of g yield that u = 0. This is impossible since u ≠ 0. Thus, the inverse of mɛ is given by $m_\varepsilon^{- 1}(u) = \frac{u}{\|u\|_{X_\varepsilon}} \in S_\varepsilon^+$ for all $u \in \mathcal{N}_\varepsilon$. Consequently, $m_\varepsilon^{- 1}$ is well-defined and continuous. On the other hand, for all $u \in S_\varepsilon^+$, we can deduce that
This yields that mɛ is bijection.
Next, we show that $\hat{m}_\varepsilon$ is a continuous map. To this end, let $\{u_n,u\}_{n \in \mathbb{N}} \subset {X_\varepsilon^+}$ such that $u_n \rightarrow u$ in Xɛ as $n \rightarrow \infty$. On account of the fact that $\hat{m}_\varepsilon (t u) = \hat{m}_\varepsilon(u)$ for all t > 0, we can assume that $\|u_n\|_{X_\varepsilon} = \|u\|_{X_\varepsilon} = 1$ for all $n \in \mathbb{N}$. According to lemma 2.2 (b), we know that there exists $t_n :=t_{u_n} \rightarrow t_0 \gt 0$ as $n \rightarrow \infty$ such that $t_n u_n \in \mathcal{N}_\varepsilon$, then we have
In the above relation, we pass to the limit as $n \rightarrow \infty$. Then,
This implies that $t_0 u \in \mathcal{N}_\varepsilon$. From lemma 2.2 (a), we know that $t_u = t_0$. Hence, it follows that $\hat{m}_\varepsilon(u_n) \rightarrow \hat{m}_\varepsilon(u)$ in ${X_\varepsilon^+}$ as $n \rightarrow \infty$. Thus, $\hat{m}_\varepsilon$ and mɛ are continuous mappings.
(d) For ɛ > 0, $0 \lt t \lt 1$, and $u \in S_\varepsilon^+$, using theorem 4.3 of Lieb and Loss [Reference Lieb and Loss38], $(g_2)$, $(f_1)$–$(f_2)$, (2.1), and hypothesis $(V_1)$ again, we can conclude that there exists C > 0 such that
Thus, we can find $\varrho_0 \gt 0$ such that $J_\varepsilon(t u) \geqslant \varrho_0$ for $0 \lt t \lt 1$ sufficiently small (since $2 p \gt q \gt 0$), where ϱ 0 is independent of $\varepsilon,$ K, and a. In addition, by lemma 2.2 (a), (b), and (c), we observe (see the study by Szulkin and Weth [Reference Szulkin, Weth, Gao and Motreanu55]) that
So, $J_\varepsilon(u) |_{\mathcal{N}_\varepsilon} \geqslant\varrho_0$.
(e) Let $\left\{u_n\right\}_{n \in \mathbb{N}} \subset S_\varepsilon^+$ be a sequence such that ${\mathrm{dist}\,}(u_n,\partial S_\varepsilon^+) \rightarrow 0$ as $n \rightarrow \infty$. For any $\varphi \in \partial S_\varepsilon^+$ and $n \in \mathbb{N}$, then we obtain $|u_n^+| \leqslant |u_n - \varphi|$ a.e. in $\Lambda_\varepsilon$. Hence, from $(V_1)$ and Sobolev embedding, for any $r \in [p,q_s^*]$ and $n \in \mathbb{N}$, it follows that
Taking into account $\|u_n\|_{X_\varepsilon} = 1$ and using hypothesis $(V_1)$, we have
Note that $0 \lt \mu \lt sp$ and $\nu \lt (N - \mu) q / (N - sq).$ Then, for all t > 0, we can deduce from lemma 2.6, $(V_1)$, $(g_3)_{ii}$, $(g_2)$, and $(f_1)$–$(f_2)$ that
where C 1, C 2, $\hat{C}_p$, and $\hat{C}_\nu$ are some positive constants. So, we have
as $n \rightarrow \infty$. Moreover, for any t > 1, we infer that
Recalling that the definition of mɛ and invoking relations (2.6) and (2.7), for all t > 1, we can deduce that
The above inequality combined with the definition of Jɛ and the arbitrariness of t > 1 means that
and so $\|m_\varepsilon(u_n)\|_{X_\varepsilon} \rightarrow + \infty$ as $n \rightarrow \infty$. This proof is now complete.
Now, we introduce the functionals
defined by $\hat{\psi}_\varepsilon(u) := J_\varepsilon(\hat{m}_\varepsilon(u))$ for any $u \in {X_\varepsilon^+}$ and $\psi_\varepsilon := \hat{\psi}_\varepsilon |_{S_\varepsilon^+}$.
Using lemma 2.2 and corollary 2.3 in the study by Szulkin and Weth [Reference Szulkin, Weth, Gao and Motreanu55], we deduce that the following result holds true.
Lemma 2.3. Assume that $(f_1)$–$(f_4)$ and $(V_1)$–$(V_2)$ are satisfied. Then,
(a) $\hat{\psi}_\varepsilon \in C^1({X_\varepsilon^+},\mathbb{R})$ and $\langle \hat{\psi}'_\varepsilon(u), v\rangle = \frac{\|\hat{m}_\varepsilon(u)\|_{X_\varepsilon}}{\|u\|_{X_\varepsilon}} \langle J_\varepsilon'(\hat{m}_\varepsilon(u)),v\rangle$ for all $u \in X_\varepsilon^+,$ all $v \in {X_\varepsilon};$
(b) $\psi_\varepsilon \in C^1(S_\varepsilon^+,\mathbb{R})$ and $\langle \psi'_\varepsilon(u),v\rangle = \|m_\varepsilon(u)\|_{X_\varepsilon} \langle J_\varepsilon'(m_\varepsilon(u)),v\rangle$ for all $u \in S_\varepsilon^+,$ all $v \in T_u(S_\varepsilon^+);$
(c) if $\{u_n\}_{n \in \mathbb{N}}$ is a Palais–Smale sequence for the functional $\psi_\varepsilon,$ then $\{m_\varepsilon(u_n)\}_{n \in \mathbb{N}}$ is a Palais–Smale sequence for the functional $J_\varepsilon.$ If $\{u_n\}_{n \in \mathbb{N}} \subset \mathcal{N}_\varepsilon$ is bounded Palais–Smale sequence for the functional $J_\varepsilon,$ then, $\{m_\varepsilon^{- 1}(u_n)\}_{n \in \mathbb{N}} \subset S_\varepsilon^+$ is a Palais–Smale sequence for $\psi_\varepsilon;$
(d) $u \in S_\varepsilon^+$ is a critical point of the functional ψɛ if and only if $m_\varepsilon(u) \in \mathcal{N}_\varepsilon$ is a critical point of the functional $J_\varepsilon.$ Moreover, the corresponding critical values coincide and
\begin{align*}\inf_{u \in S_\varepsilon^+} \psi_\varepsilon(u) = \inf_{u \in \mathcal{N}_\varepsilon}J_\varepsilon(u) = c_\varepsilon.\end{align*}
Lemma 2.4. The modified functional Jɛ admits a Palais–Smale sequence $\{u_n\}_{n \in \mathbb{N}} \subset {X_\varepsilon}$ at the level $c_\varepsilon,$ that is, $J_\varepsilon(u_n) \rightarrow c_\varepsilon$ in $\mathbb{R}$ and $J_\varepsilon'(u_n) \rightarrow 0$ in ${X_\varepsilon}^*$ as $n \rightarrow \infty,$ where cɛ is given in lemma 2.2. Furthermore, there exists some constant ϑ > 0 $($independent of $\varepsilon,$ K, and $a)$ such that $c_\varepsilon \lt \vartheta$ for all ɛ sufficiently small.
Proof. In view of lemma 2.2, we only need to verify that Jɛ possesses a mountain pass geometry, that is, the functional Jɛ satisfies the following properties:
(i) there exist some constants $\rho_1 \gt 0$ and $\delta_1 \gt 0$ such that $J_\varepsilon(u) \geqslant \delta_1$ for $\|u\|_{X_\varepsilon} = \rho_1$;
(ii) there exists an element $ e \in {X_\varepsilon}$ with $\|e\|_{X_\varepsilon} \gt \rho_1$ such that $J_\varepsilon(e) \lt 0$.
(i) Arguing as in the proof of lemma 2.2 (a), we find some constants C > 0 and $C' \gt 0$ such that
for $\|u\|_{X_\varepsilon} = \rho_1 \in (0,1)$. So, using $2 p \gt q$, we obtain (i) if we take ρ 1 small enough.
(ii) We choose a suitable function $\varphi \in C_0^\infty(\mathbb{R}^N)$ such that $\varphi \geqslant 0$, $\varphi \not\equiv 0$, and ${\mathrm{supp}}\, (\varphi) \subset \Lambda$. Then, for all ɛ sufficiently small, it is obvious that $G(\varepsilon x, \varphi) = F(\varphi)$ for all $x \in \mathbb{R}^N$. Hence, with arguments as in the proof of lemma 2.2 (a), there exist two constants
and C > 0 such that
for all t > 1. Hence, (ii) holds true for $e = t \varphi$ and for some t > 1 sufficiently large.
According to the mountain pass theorem without the Palais–Smale condition (see the study by Brezis and Nirenberg [Reference Brezis and Nirenberg18]), we establish the existence of a Palais–Smale sequence $\{u_n\}_{n \in \mathbb{N}} \subset {X_\varepsilon}$ at the level cɛ. We recall that ${\mathrm{supp}}\, (\varphi) \subset \Lambda$, and then we can infer that there exists a constant ϑ > 0 such that $c_\varepsilon \lt \vartheta$ for all ɛ small enough, where ϑ is independent of ɛ, K, and a. This completes the proof of the lemma.
Lemma 2.5. Assume that $\{u_n\}_{n \in \mathbb{N}} \subset {X_\varepsilon}$ is the Palais–Smale sequence of Jɛ at the level $c \leqslant \vartheta,$ where ɛ > 0 small enough. Then, for ɛ > 0 small enough, the sequence $\{u_n\}_{n \in \mathbb{N}} \subset {X_\varepsilon}$ is bounded and
Proof. According to $(g_3)_{i,ii}$ and using hypothesis $(V_1)$, for any ɛ > 0 small enough, we deduce that
So, for ɛ > 0 small enough, we deduce that the sequence $\{u_n\}_{n \in \mathbb{N}} \subset {X_\varepsilon}$ is bounded and
for all $n \in \mathbb{N}$ large enough.
According to (2.8), we now define the following set:
where ϑ is given in lemma 2.4. Using the above notation, we can show the following estimate.
Lemma 2.6. Assume that $(f_1)$–$(f_4)$ hold true, $0 \lt \mu \lt sp$, and $\nu \lt (N - \mu) q / (N - sq),$ then there is a constant K > 0 such that
Proof. Firstly, $(g_2)$ and $(f_1)$–$(f_2)$ imply that there exists a constant C > 0 (which is independent of ɛ) such that
Therefore,
for some constant C > 0 (which is independent of $x \in \mathbb{R}^N$ and ɛ > 0).
Let us choose
since $0 \lt \mu \lt sp$ and $\nu \lt (N - \mu) q / (N - sq).$ Then, combining the Hölder inequality, (2.1), and the definition of $\mathcal{B}$, we can easily see that
for some constant C > 0 (which is independent of $x \in \mathbb{R}^N$ and ɛ > 0). So, from the above information, we complete the proof of this lemma.
Now, we show that the modified functional Jɛ satisfies the Palais–Smale condition.
Lemma 2.7. Let $0 \lt \mu \lt sp$ and $\nu \lt (N - \mu) q / (N - sq).$ Then, for all ɛ > 0 sufficiently small, the modified functional Jɛ satisfies the Palais–Smale condition with $c \leqslant \vartheta$.
Proof. For all ɛ > 0 sufficiently small, let $\{u_n\}_{n \in \mathbb{N}} \subset {X_\varepsilon}$ be a Palais–Smale sequence of the functional Jɛ at the level c. From lemma 2.5, we know that $\{u_n\}_{n \in \mathbb{N}} \subset {X_\varepsilon}$ is bounded for all ɛ > 0 small enough. So, passing to a subsequence, we may assume that there exists some $u \in {X_\varepsilon}$ such that $u_n \xrightarrow w u$ in Xɛ, $u_n(x) \rightarrow u(x)$ a.e. in $\mathbb{R}^N$ and $u_n \rightarrow u$ in $L^r_{\mathrm{loc}}(\mathbb{R}^N)$ for all $r \in [1,q_s^*)$ as $n \rightarrow \infty$. We first assert that the following property is fulfilled.
Claim 1: The following properties hold up to a subsequence:
(a) $\lim_{n \rightarrow \infty}\int_{\mathbb{R}^N} \left(\frac{1}{|x|^\mu} * \left(G(\varepsilon x,u_n) -G(\varepsilon x,u)\right)\right) g(\varepsilon x,u) \varphi dx = 0$ for all $\varphi \in C^\infty_0(\mathbb{R}^N)$;
(b) $\lim_{n \rightarrow \infty}\int_{\mathbb{R}^N} \left(\frac{1}{|x|^\mu} * G(\varepsilon x,u_n)\right) \left(g(\varepsilon x,u_n) - g(\varepsilon x,u) \right) \varphi dx = 0$ for all $\varphi \in C^\infty_0(\mathbb{R}^N)$;
(c) we have
\begin{align*} &\lim_{n \rightarrow \infty} \int_{\mathbb{R}^N}\left(\frac{1}{|x|^\mu} * G(\varepsilon x, u_n)\right) g(\varepsilon x,u_n) \varphi dx\\ =\ & \int_{\mathbb{R}^N}\left(\frac{1}{|x|^\mu} * G(\varepsilon x, u)\right) g(\varepsilon x,u) \varphi dx \ \text{for all}\ \varphi \in C^\infty_0(\mathbb{R}^N). \end{align*}
(a) Note that
On the other hand, since $u_n(x) \rightarrow u(x)$ a.e. in $\mathbb{R}^N$ as $n \rightarrow \infty$, we obtain
From proposition 5.4.7 in the study by Willem [Reference Willem58], it follows that
Let us define
Then, it follows from theorem 4.3 of [Reference Lieb and Loss38], $(g_2)$, $(f_1)$–$(f_2)$, Hölder’s inequality, (2.1), and $u\in {X_\varepsilon}$ that H is linear bounded functional. Combining this with (2.10), we derive that (a) holds true.
(b) Using the boundedness of the sequence $\{u_n\}_{n \in \mathbb{N}} \subset {X_\varepsilon}$, together with theorem 4.3 in the study by Lieb and Loss [Reference Lieb and Loss38], $(g_2)$, $(f_1)$–$(f_2)$, and (2.1), we see that $G(\varepsilon \cdot,u_n)$ is also bounded in $L^{\frac{2 N}{2 N -\mu}}(\mathbb{R}^N)$. Then, applying the boundedness of the sequence $\{u_n\}_{n \in \mathbb{N}} \subset {X_\varepsilon}$ again and combining with the compact embeddings and Dominated Convergence Theorem, we deduce that (b) is also true.
(c) This result is a consequence of (a) and (b).
Next, we consider the following sequence:
and we also introduce the following function
where $p' = \frac{p}{p - 1}$. Then, $\{b_n\}_{n \in \mathbb{N}} \subset L^{p'}(\mathbb{R}^{2 N})$ is bounded and $b_n(x,y) \rightarrow b(x,y)$ a.e. in $\mathbb{R}^{2 N}$ as $n \rightarrow \infty$. Then, we deduce from proposition 5.4.7 in the study by Willem [Reference Willem58] that
On the other hand, for all $\varphi \in C^\infty_0(\mathbb{R}^N)$, we see that
From this and (2.11), we have
In a similar fashion, we also have
Finally, it is easy to check that
From the above information, together with the fact that $\langle J_\varepsilon'(u_n),\varphi \rangle \rightarrow 0$ as $n \rightarrow \infty$, it follows that $\langle J_\varepsilon'(u),\varphi \rangle = 0$ for all $\varphi \in C^\infty_0(\mathbb{R}^N)$. Using the density of $C^\infty_0(\mathbb{R}^N)$ in Xɛ, we know that u is a critical point of the functional Jɛ. Consequently, $\langle J_\varepsilon'(u),u\rangle = 0$.
In order to show that the Palais–Smale sequence satisfies the Palais–Smale condition, we need to establish the following property.
Claim 2: We have
We first observe that, for all ɛ > 0 small enough, there exists $ n_0 \in \mathbb{N}$ such that
For any R > 0, let $\eta_R \in C^\infty(\mathbb{R}^N)$ be such that $0 \leqslant \eta_R \leqslant 1$, $\eta_R = 0$ in $B_{\frac{R}{2}}$, $\eta_R = 1$ in $B_R^c$, and $|\nabla \eta_R| \leqslant C / R$ for some constant C > 0 (which is independent of R). Taking into account the boundedness of the sequence $\{\eta_R u_n\}_{n \in \mathbb{N}} \subset {X_\varepsilon}$, we see that $\langle J_\varepsilon'(u_n),\eta_R u_n \rangle \rightarrow 0$ as $n \rightarrow \infty$. It follows that
Fix ɛ > 0 small enough. Let R > 0 sufficiently large such that $\Lambda_\varepsilon \subset B_{\frac{R}{2}}$. Using the definitions of ηR and K, together with $(g_3)_{ii}$, for $n \geqslant n_0$ large enough, we have
From the Hölder inequality and the boundedness of the sequence $\{u_n\}_{n \in \mathbb{N}} \subset {X_\varepsilon}$, it follows that
In addition, by the definition of ηR, polar coordinates, and the boundedness of the sequence $\{u_n\}_{n \in \mathbb{N}} \subset {X_\varepsilon}$, we have
Using (2.13) and (2.14), we see that
Also, we have
So, we deduce from (2.12), (2.15), and (2.16) that Claim 2 holds true.
Using Claim 2 and applying the locally compact embedding $X_\varepsilon \hookrightarrow L^p_{\mathrm{loc}}(\mathbb{R}^N)$, we can derive that $u_n \rightarrow u$ in $L^p(\mathbb{R}^N)$ as $n \rightarrow \infty$. In addition, we deduce from the interpolation inequality that $u_n \rightarrow u$ in $L^r(\mathbb{R}^N)$ as $n \rightarrow \infty$ for all $r \in [p,q_s^*)$. Then, from theorem 4.3 in the study by Lieb–Loss [Reference Lieb and Loss38], $(g_2)$, $(f_1)$–$(f_2)$, the Dominated Convergence Theorem, and the boundedness of the sequence $\{u_n\}_{n \in \mathbb{N}} \subset {X_\varepsilon}$, it follows that
Also, we have
and
Hence, the above fact and (2.17) imply that
Furthermore, according to lemma 2.10 (a), (b), we see that
and
as $n \rightarrow \infty$. Thus, we conclude that
This fact means that $\|u_n - u\|_{X_\varepsilon} \rightarrow 0$ as $n \rightarrow \infty$. Now, we complete the proof of the lemma.
Lemma 2.8. Let $0 \lt \mu \lt sp$ and $\nu \lt (N - \mu) q / (N - sq).$ Assume that $(f_1)$–$(f_4)$ and $(V_1)$–$(V_2)$ are fulfilled, then for all ɛ > 0 small enough, problem (2.2) possesses a non-negative solution $u_\varepsilon \in {X_\varepsilon}.$
Proof. Using lemmas 2.4 and 2.7, we can employ the Mountain Pass Theorem to infer that for all ɛ > 0 sufficiently small, there exists $u_\varepsilon \in {X_\varepsilon}$ such that $J_\varepsilon'(u_\varepsilon) = 0$ and $J_\varepsilon(u_\varepsilon) = c_\varepsilon$. Furthermore, choosing $u_\varepsilon^- := \min\{u_\varepsilon,0\}$ and recalling that $g(\varepsilon \cdot,t) = 0$ for all $t \leqslant 0$ and $\langle J_\varepsilon'(u_\varepsilon),u_\varepsilon^- \rangle = 0$, we can infer that
This ends the proof of the lemma.
Corollary 2.9. Let $0 \lt \mu \lt sp$ and $\nu \lt (N - \mu) q / (N - sq).$ Then, for all ɛ > 0 sufficiently small, the functional ψɛ satisfies the Palais–Smale condition at the level $c \leqslant \vartheta$ on $S_\varepsilon^+.$
Proof. Let $\{u_n\}_{n \in \mathbb{N}} \subset S_\varepsilon^+$ be a Palais–Samle sequence for the functional ψɛ at the level c, that is,
Using lemma 2.3 (c), we see that $\left\{m_\varepsilon(u_n)\right\}_{n \in \mathbb{N}} \subset {X_\varepsilon}$ is also a Palais–Samle sequence for the functional Jɛ at the level c. Therefore, we can deduce from lemma 2.7 that the functional Jɛ verifies the Palais–Smale condition. Hence, passing to a subsequence, we can find some $u \in S_\varepsilon^+$ such that $m_\varepsilon(u_n) \rightarrow m_\varepsilon(u)$ in Xɛ as $n \rightarrow \infty$. Combining this fact with lemma 2.2 (c), we conclude that $u_n \rightarrow u$ in $S_\varepsilon^+$ as $n \rightarrow \infty$.
We end this section by showing the following result:
Lemma 2.10. Let $\left\{u_n\right\}_{n \in \mathbb{N}} \subset {X_\varepsilon}$ be a sequence such that $u_n \xrightarrow w u$ in Xɛ, and let $\mathcal{A}_t : \mathbb{R} \mapsto \mathbb{R}$ defined by $\mathcal{A}_t(\tau) = |\tau|^{t - 2} \tau$ $(t \in \left\{p,q\right\}).$ Setting $w_n = u_n - u,$ then for all $\varepsilon \geqslant 0$ and $n \in \mathbb{N}$ large enough, we have
(a) $\left([u_n]_{s,p}^p + [u_n]_{s,q}^q \right) - \left([w_n]_{s,p}^p + [w_n]_{s,q}^q \right) - \left([u]_{s,p}^p + [u]_{s,q}^q \right) = o_n(1);$
(b) $\left(\left|V(\varepsilon \cdot)^{1/p} u_n\right|_p^p + \left|V(\varepsilon \cdot)^{1/q} u_n\right|_q^q\right) - \left(\left|V(\varepsilon \cdot)^{1/p} w_n\right|_p^p + \left|V(\varepsilon \cdot)^{1/q} w_n\right|_q^q\right) -\left(\left|V(\varepsilon \cdot)^{1/p} u\right|_p^p + \left|V(\varepsilon \cdot)^{1/q} u\right|_q^q\right) = o_n(1);$
(c) $\left| \mathcal{A}_p(u_n)- \mathcal{A}_p(w_n) - \mathcal{A}_p(u) \right|_{{\frac{p}{p - 1}}}^{\frac{p}{p - 1}}+\left| \mathcal{A}_q(u_n)- \mathcal{A}_q(w_n) - \mathcal{A}_q(u) \right|_{{\frac{q}{q - 1}}}^{\frac{q}{q - 1}} = o_n(1);$
(d) $\int_{\mathbb{R}^N}\int_{\mathbb{R}^N} \Bigg|\frac{\mathcal{A}_p(u_n(x) - u_n(y))}{|x - y|^{\frac{N + sp}{p / (p - 1)}}} - \frac{\mathcal{A}_p(w_n(x) - w_n(y))}{|x - y|^{\frac{N + sp}{p / (p - 1)}}} - \frac{\mathcal{A}_p(u(x) - u(y))}{|x - y|^{\frac{N + sp}{p / (p - 1)}}} \Bigg|^{\frac{p}{p - 1}}dxdy$
$ + \int_{\mathbb{R}^N}\int_{\mathbb{R}^N} \Bigg|\frac{\mathcal{A}_q(u_n(x) - u_n(y))}{|x - y|^{\frac{N + sq}{q / (q - 1)}}} - \frac{\mathcal{A}_q(w_n(x) - w_n(y))}{|x - y|^{\frac{N + sq}{q / (q - 1)}}} - \frac{\mathcal{A}_q(u(x) - u(y))}{|x - y|^{\frac{N + sq}{q / (q - 1)}}} \Bigg|^{\frac{q}{q - 1}}dxdy$
$ = o_n(1);$
(e) $\int_{\mathbb{R}^N} \left(\frac{1}{|x|^\mu} * F(u_n)\right) F(u_n) dx - \int_{\mathbb{R}^N} \left(\frac{1}{|x|^\mu} * F(w_n)\right) F(w_n) dx - \int_{\mathbb{R}^N} \left(\frac{1}{|x|^\mu} * F(u)\right) F(u) dx = o_n(1)$
and for any $\varphi \in {X_\varepsilon}$ with $\|\varphi\|_{X_\varepsilon} \leqslant 1$, it holds
(f) $\int_{\mathbb{R}^N} \left(\frac{1}{|x|^\mu} * F(u_n)\right) f(u_n) \varphi dx - \int_{\mathbb{R}^N} \left(\frac{1}{|x|^\mu} * F(w_n)\right) f(w_n) \varphi dx - \int_{\mathbb{R}^N} \left(\frac{1}{|x|^\mu} * F(u)\right) f(u) \varphi dx = o_n(1).$
Proof. We first point out that the proofs of (a) and (b) follow from the Brezis–Lieb Lemma (see [Reference Brezis and Lieb17]). Moreover, arguing as in the proof lemma 3.2 of Mercuri and Willem [Reference Mercuri and Willem43], we see that (c) holds true.
Next, we give the details of the proofs of (d), (e), and (f) for the convenience of readers.
(d) Case 1: $2 \leqslant t \in \left\{p,q\right\}$.
We deduce from the Mean Value Theorem, the Young inequality, and $t \geqslant 2$ that for any fixed σ > 0, there exists $C_\sigma \gt 0$ such that
In (2.18), we take
and we have
Now, we introduce the mapping $H_{\sigma,n}^1 : \mathbb{R}^{2 N} \mapsto \mathbb{R}_+$ given by
Then, $H_{\sigma,n}^1(x,y) \rightarrow 0$ a.e. in $\mathbb{R}^{2 N}$ as $n \rightarrow \infty$ and
for some constant C > 0. Hence, from the Dominated Convergence Theorem, we obtain
In addition, according to the definition of $H_{\sigma,n}^1$, we have
for some constant C > 0. This implies that
holds true for $n \in \mathbb{N}$ large enough.
Case 2: $1 \lt t \lt 2$.
Invoking lemma 3.1 in the study by Mercuri and Willem [Reference Mercuri and Willem43] and applying the Dominated Convergence Theorem, for $n \in \mathbb{N}$ large enough, we obtain that
holds true.
Now, combining Case 1 and Case 2, we complete the proof of (d).
(e) Using the Mean Value Theorem, hypotheses $(f_1)$–$(f_2)$, and Young’s inequality, we can infer that for any σ > 0, there exists $C_\sigma \gt 0$ such that
Define the following mapping:
Then, arguing as in the proof of lemma 2.10 (d) and using the boundedness of the sequence $\left\{u_n\right\}_{n \in \mathbb{N}} \subset {X_\varepsilon}$, for $n \in \mathbb{N}$ sufficiently large, we can derive that
On the other hand, we can deduce from theorem 4.3 in the study by Lieb and Loss [Reference Lieb and Loss38] that
for $n \in \mathbb{N}$ large enough.
We note that
Using the above equality and taking into account that $F(w_n) \xrightarrow w 0$ in $L^{\frac{2 N }{2 N - \mu}}(\mathbb{R}^N)$ as $n \rightarrow \infty$, together with (2.19)–(2.20), we can show that (e) holds true.
(f) We first prove that
for $n \in \mathbb{N}$ large enough.
For any fixed $0 \lt \sigma \lt 1$, we deduce from hypothesis $(f_1)$ that there exists $0 \lt \lambda_0 := \lambda_0(\sigma) \lt 1$ such that
In addition, by hypothesis $(f_2)$, we can find $\lambda_1 := \lambda_1(\sigma) \gt 2$ such that
Next, by the continuity of f, there is a positive constant $\gamma :=\gamma(\sigma) \lt \lambda_0$ such that
Additionally, combining $(f_1)$ and $(f_2)$, we can conclude that there exists $C(\sigma) \gt 0$ such that
We now estimate the following term:
Since $u \in {X_\varepsilon}$, together with (2.25), the Hölder inequality (2.1), and $(V_1)$, we know that there exists $R = R(\sigma) \gt 0$ such that
for some constant C > 0.
Let us define
According to (2.22), the Hölder inequality, (2.1), $(V_1)$, and the boundedness of the sequence $\left\{u_n\right\}_{n \in \mathbb{N}} \subset {X_\varepsilon}$, we have
for some constant C > 0.
Set
Arguing as in (2.27), together with (2.23), we can obtain
for some positive constant C > 0.
Now, we introduce the following set:
It is easy to check that $|D_{u_n}^3| \lt + \infty$, since $u_n \in {X_\varepsilon}$. So, from (2.24), it follows that
for some constant C > 0.
Thus, we deduce from (2.27)–(2.29) that
Thanks to $u \in {X_\varepsilon}$, we have $|(\mathbb{R}^N \setminus B_R) \cap \{|u| \geqslant \gamma\}| \rightarrow 0$ as $R \rightarrow \infty$. Using (2.25) again, we deduce that there exists $R := R(\sigma) \gt 0$ such that
where C is a positive constant. Combining this inequality with (2.26) and (2.30), we deduce that there is a constant C > 0 such that for all $n \in \mathbb{N}$,
Recalling that $u_n \xrightarrow w u$ in Xɛ as $n \rightarrow \infty$ and passing to a subsequence (still denoted by $\left\{u_n\right\}_{n \in \mathbb{N}}$), we can assume that $u_n \rightarrow u$ strongly in $L^{\frac{2 N p}{2 N - \mu}}(B_R)$ (since $0 \lt \mu \lt 2 sp$) and there is a function $d \in L^{\frac{2 N p}{2 N - \mu}}(B_R)$ such that $|u_n(x)|$, $|u(x)| \leqslant d(x)$ a.e. $x \in B_R$.
For $n \in \mathbb{N}$ large enough, we have
where C > 0 is some constant.
Set $D_{u_n}^4 :=\left\{x \in B_R : |u_n(x) - u(x)| \geqslant 1\right\}$. So, we get
where C > 0 is some constant. We observe that $|D_{u_n}^4| \rightarrow 0$ in $\mathbb{R}$ as $n \rightarrow \infty$, and then we can deduce that there exists some constant C > 0 such that
provided that $n \in \mathbb{N}$ is sufficiently large.
In addition, from $u \in {X_\varepsilon}$, it follows that
Using the above fact and invoking (2.25) again, we can infer that there exists $L :=L(\sigma) \gt 0$ such that
where C > 0 is a constant. Additionally, we can deduce from the Dominated Convergence Theorem that
for $n \in \mathbb{N}$ large enough.
It follows that
for some constant C > 0.
Hence, we deduce from (2.32), (2.33), (2.34), and (2.35) that there exists some constant C > 0 such that
for large enough $n \in \mathbb{N}$.
Putting together (2.31) and (2.36), we conclude that (2.21) holds true.
On account of the fact that $\frac{1}{|x|^\mu}*F(u) \in L^\frac{2 N}{\mu}(\mathbb{R}^N)$, for any $0 \lt \sigma \lt 1$, there exists $R_1 :=R_1(\sigma) \gt 0$ such that
Using the above all information and by a straightforward computation, for $n \in \mathbb{N}$ large enough, we can conclude that
where Cu and C are positive constants.
Next, we prove the following relation:
for $n \in \mathbb{N}$ sufficiently large.
Indeed, it is easy to check that, for any $\sigma \in (0,1)$ and some $R_2 :=R(\sigma) \gt 0$, there exists some constant C > 0 such that
for $n \in \mathbb{N}$ large enough.
To prove (2.38), now we need to estimate the following part:
To this end, we divide this into two parts for discussion.
($*$) If $f(u(x)) \varphi(x) = 0$ a.e. on $B_{R_2}$, so, we have
($**$) If $|\left\{x \in B_{R_2} : f(u(x)) \varphi(x) \neq 0\right\}| \gt 0$, that is,
In addition, we can easily derive that
Let us define
In the case ($**$), we can apply the above relations, theorem 4.3 in the study by Lieb and Loss [Reference Lieb and Loss38], the Sobolev continuous embedding and the local compactness Sobolev embedding to infer that, for $n \in \mathbb{N}$ sufficiently large,
Thus, relations (2.39), (2.41), and (2.42) and ($*$) imply that (2.38) holds true.
Finally, we point out that
Then, the above equality combined with (2.19), (2.20), (2.21), (2.37), and (2.38); theorem 4.3 in the study by Lieb and Loss [Reference Lieb and Loss38]; the Hölder inequality; the boundedness of the sequence $\left\{u_n\right\}_{n \in\mathbb{N}} \subset {X_\varepsilon}$; and the fact that $\|\varphi\|_{X_\varepsilon} \leqslant 1$ yields that (f) is true. This proof is now complete.
3. The limit problem
To the best of our knowledge, there is no result about the nonlinear fractional (p, q)-Choquard problem. That is why we need to consider the following limit problem associated with problem (1.3):
It is worth pointing out that we denote by $X_{V_0}$ the space Xɛ when ɛ = 0. For all $u \in {X_{V_0}}$, we introduce the corresponding energy functional associated with problem (3.1) defined by
It is obvious that the functional $I_{V_0}$ is well-defined and belongs to C 1, with its differential given by
for all u, $v \in {X_{V_0}}$. So, it is easy to check that all the solutions of problem (3.1) correspond to critical points of the functional $I_{V_0}$.
Next, we denote by $\mathcal{M}_{V_0}$ the Nehari manifold associated with the functional $I_{V_0}$, that is,
Moreover, we define $c_{V_0} : =\inf_{u \in \mathcal{M}_{V_0}} I_{V_0}(u)$.
Now, we define the following sets:
where $S_{V_0}$ is the unit sphere of $X_{V_0}$. We point out that $S_{V_0}^+$ is also an incomplete $C^{1,1}$-manifold of codimension one and contained in $X_{V_0}^+$. Hence, $X_{V_0} = T_u (S_{V_0}^+)\bigoplus \mathbb{R} u$ for each $u \in S_{V_0}^+$, where
With arguments as in the proof of lemma 2.2, we can deduce that the following property holds true.
Lemma 3.1. Assume that $(f_1)$–$(f_4)$ and $(V_1)$ are fulfilled, then we have the following properties:
(a) for all $u \in {X_{V_0}^+}$, there exists a unique $t_u \gt 0$ such that $t_u u \in \mathcal{M}_{V_0}.$ Furthermore, $\hat{m}_{V_0}(u) = t_u u$ is the unique maximum of $I_{V_0}(t u);$
(b) there exists τ > 0 independent of u such that $t_u \geqslant \tau$ for all $u \in S_{V_0}^+.$ Moreover, for each compact set $\mathcal{W} \subset S_{V_0}^+,$ there exists a constant $C_\mathcal{W} \gt 0$ such that $t_u \leqslant C_\mathcal{W}$ for all $u \in \mathcal{W};$
(c) the mapping $\hat{m}_{V_0} : {X_{V_0}^+} \mapsto \mathcal{M}_{V_0}$ is continuous, $m_{V_0} :=\hat{m}_{V_0} |_{S_{V_0}^+}$ is a homeomorphism between $S_{V_0}^+$ and $\mathcal{M}_{V_0},$ and the inverse of $m_{V_0}$ is given by $m_{V_0}^{-1}(u) := u / \|u\|_{X_{V_0}};$
(d) $c_{V_0} : = \inf_{u \in \mathcal{M}_{V_0}} I_{V_0}(u) \gt 0$, and $I_{V_0}$ is bounded below on $\mathcal{M}_{V_0}$ by some positive constant$;$
(e) let $0 \lt \mu \lt sp$ and $\nu \lt (N - \mu) q / (N - sq).$ If there exists a sequence $\left\{u_n\right\}_{n \in \mathbb{N}} \subset S_{V_0}^+$ such that ${\mathrm{dist}\,}(u_n,\partial S_\varepsilon^+) \rightarrow 0$ as $n \rightarrow \infty,$ then $\|m_{V_0}(u_n)\|_{X_{V_0}} \rightarrow + \infty$ and $J_\varepsilon(m_\varepsilon(u_n)) \rightarrow + \infty$ as $n \rightarrow \infty;$
(f) $I_{V_0}$ is coercive on $\mathcal{M}_{V_0}.$
Now, let us define the mappings
by $\hat{\psi}_{V_0}(u) := I_{V_0}(\hat{m}_{V_0}(u))$ for all $u \in {X_{V_0}^+}$ and $\psi_{V_0} := \hat{\psi}_{V_0} |_{S_{V_0}^+}$.
Lemma 3.2. Assume that $(f_1)$–$(f_4)$ and $(V_1)$ hold true. Then,
(a) $\hat{\psi}_{V_0} \in C^1({X_{V_0}^+},\mathbb{R})$ and $\langle \hat{\psi}'_{V_0}(u), v\rangle = \frac{\|\hat{m}_{V_0}(u)\|_{X_{V_0}}}{\|u\|_{X_{V_0}}} \langle I_{V_0}'(\hat{m}_{V_0}(u)),v\rangle$ for all $u \in {X_{V_0}^+},$ all $v \in {X_{V_0}};$
(b) $\psi_{V_0} \in C^1(S_{V_0}^+,\mathbb{R})$ and $\langle \psi'_{V_0}(u),v\rangle = \|m_{V_0}(u)\|_{X_{V_0}} \langle I_{V_0}'(m_{V_0}(u)),v\rangle$ for all $u \in S_{V_0}^+,$ all $v \in T_u(S_{V_0}^+);$
(c) if $\{u_n\}_{n \in \mathbb{N}}$ is a Palais–Smale sequence for $\psi_{V_0},$ then $\{m_{V_0}(u_n)\}_{n \in \mathbb{N}}$ is a Palais–Smale sequence for $I_{V_0}.$ If $\{u_n\}_{n \in \mathbb{N}} \subset \mathcal{M}_{V_0}$ is a bounded Palais–Smale sequence for $I_{V_0},$ then $\{m_{V_0}^{- 1}(u_n)\}_{n \in \mathbb{N}} \subset S_{V_0}^+$ is a Palais–Smale sequence for $\psi_{V_0};$
(d) $u \in S_{V_0}^+$ is a critical point of $\psi_{V_0}$ if and only if $m_{V_0}(u) \in \mathcal{M}_{V_0}$ is a critical point of $I_{V_0}.$ Moreover, the corresponding critical values coincide and
\begin{align*}\inf_{u \in S_{V_0}^+} \psi_{V_0}(u) = \inf_{u \in \mathcal{M}_{V_0}}I_{V_0}(u) = c_{V_0}.\end{align*}
Remark 3.3. The following variational characterization for $c_{V_0}$ is fulfilled:
Lemma 3.4. Assume that $\left\{u_n\right\}_{n \in \mathbb{N}} \subset {X_{V_0}}$ is a Palais–Smale sequence of the functional $I_{V_0}$ at the level $c_{V_0}.$ Then, the sequence $\left\{u_n\right\}_{n \in \mathbb{N}} \subset {X_{V_0}}$ is bounded, and there exist a sequence $\left\{y_n\right\}_{n \in \mathbb{N}} \subset \mathbb{R}^N$ and some constants $R,$ α > 0 such that
Proof. Arguing as in lemma 2.5, it is obvious to see that $\left\{u_n\right\}_{n \in \mathbb{N}} \subset {X_{V_0}}$ is bounded. Now, arguing by contradiction, suppose that for any R > 0, the following relation
holds true. Then, using lemma 2.2 in the study by Alves, Ambrosio, and Isernia [Reference Alves, Ambrosio and Isernia1], we know that for all $r \in (p,q_s^*)$,
So, we deduce from theorem 4.3 in the study by Lieb and Loss [Reference Lieb and Loss38], hypotheses $(f_1)$–$(f_2)$, the boundedness of the sequence $\left\{u_n\right\}_{n \in \mathbb{N}} \subset {X_{V_0}}$, and (3.2) that
Moreover, by the boundedness of the sequence $\left\{u_n\right\}_{n \in \mathbb{N}} \subset {X_{V_0}}$, we see that $\langle I_{V_0}'(u_n),u_n \rangle = o_n(1)$ as $n \rightarrow \infty$, that is,
This implies that $\|u_n\|_{X_{V_0}} = o_n(1)$ as $n \rightarrow \infty$. We get a contradiction since $I_{V_0}(u_n) \rightarrow c_{V_0} \gt 0$ in $\mathbb{R}$ as $n \rightarrow \infty$. The proof is now complete.
Theorem 3.5 Problem (3.1) has a positive ground state solution.
Proof. We deduce from a variant of the Mountain Pass Theorem without the Palais–Smale condition that there is a Palais–Smale sequence $\left\{u_n\right\}_{n\in \mathbb{N}} \subset {X_{V_0}}$ of the functional $I_{V_0}$ at the level $c_{V_0}$. As in the proof of lemma 2.5, we can conclude that the sequence $\left\{u_n\right\}_{n \in \mathbb{N}} \subset {X_{V_0}}$ is bounded. Then, we are able to deduce from lemma 3.4 that there exist a sequence $\left\{y_n\right\}_{n \in \mathbb{N}} \subset \mathbb{R}^N$ and some constants R, α > 0
Set $v_n(x) : = u_n(x + y_n)$. Then, we have
It is easy to check that $I_{V_0}(v_n) \rightarrow c_{V_0}$ in $\mathbb{R}$ and $I_{V_0}'(v_n) \rightarrow 0$ in ${X_{V_0}}^*$ as $n \rightarrow \infty$. Clearly, the sequence $\left\{v_n\right\}_{n \in \mathbb{N}} \subset {X_{V_0}}$ is also bounded. Thus, we may suppose that there exists $0 \neq v \in {X_{V_0}}$ such that $v_n \xrightarrow w v$ (up to a subsequence) in ${X_{V_0}}$ as $n \rightarrow \infty$. Moreover, arguing as in lemma 2.7, we have $I_{V_0}'(v) = 0$. Due to v ≠ 0, we can derive that $v \in \mathcal{M}_{V_0}$, and so $I_{V_0}(v) \geqslant c_{V_0}$. On the other hand, from hypothesis $(f_3)$ and Fatou’s Lemma, we deduce that
So, we obtain $I_{V_0}(v) = c_{V_0}$.
We can also show that this ground state solution v is positive. Set $v^- :=\min\left\{v,0\right\}$, and then $v^- \in {X_{V_0}}$. Recalling that $f(t) = 0$ for $t \leqslant 0$ and using $\langle I_{V_0}'(v),v^- \rangle = 0$, we obtain
This leads to $v^- = 0$, and so $v \geqslant 0$ on $\mathbb{R}^N$. Thus, $v \geqslant 0$ and $v \not \equiv 0$. Similar to the proof of lemma 6.1, we see that $v \in L^\infty(\mathbb{R}^N)$. In addition, from corollary 2.1 in the study by Ambrosio and Rădulescu [Reference Ambrosio and Rădulescu11], it follows that $v \in C^\sigma(\mathbb{R}^N)$ for some $\sigma \in (0,1)$. Now, it follows from the proof of theorem 1.1 (ii) in the study by Jarohs [Reference Jarohs33] that v > 0 on $\mathbb{R}^N$. This ends the proof.
Next, we introduce a compactness result for the autonomous problem, which will be very useful in the sequel.
Lemma 3.6. Let $0 \lt \mu \lt sp$ and $\nu \lt (N - \mu) q / (N - sq).$ Assume that $\left\{u_n\right\}_{n \in \mathbb{N}} \subset \mathcal{M}_{V_0}$ is a sequence satisfying $I_{V_0}(u_n) \rightarrow c_{V_0}$ in $\mathbb{R}$ as $n \rightarrow \infty,$ then the sequence $\left\{u_n(\cdot + y_n)\right\}_{n \in \mathbb{N}} \subset {X_{V_0}}$ possesses a convergent subsequence for some sequence $\left\{y_n\right\}_{n \in \mathbb{N}} \subset \mathbb{R}^N.$
Proof. On account of the fact that $\left\{u_n\right\}_{n \in \mathbb{N}} \subset \mathcal{M}_{V_0}$ and $I_{V_0}(u_n) \rightarrow c_{V_0}$ in $\mathbb{R}$ as $n \rightarrow \infty$, lemma 3.1 (c), lemma 3.2 (d), and the definition of $c_{V_0}$ imply that
and
Now, we introduce the following map:
It is worth pointing out the following essential facts:
(i) $(\overline{S_{V_0}^+}, \delta_{V_0})$, where $\delta_{V_0}(u,v) = \|u - v\|_{X_{V_0}}$, is a complete metric space;
(ii) $\mathcal{I} \in C(\overline{S_{V_0}^+},\mathbb{R} \cup \left\{+ \infty\right\})$, by lemma 3.1 (e);
(iii) $\mathcal{I}$ is bounded below, by lemma 3.2 (d).
We deduce from the Ekeland variational principle (see the study by Ekeland [Reference Ekeland28]) that there exists a Palais–Smale sequence $\left\{\hat{w}_n\right\}_{n \in \mathbb{N}} \subset S_{V_0}^+$ of the functional $\psi_{V_0}$ at the level $c_{V_0}$ such that $\|\hat{w}_n - w_n\|_{X_{V_0}} = o_n(1)$ as $n \rightarrow \infty$, that is, $\psi_{V_0}(\hat{w}_n) \rightarrow c_{V_0}$ in $\mathbb{R}$ and $\psi_{V_0}'(\hat{w}_n) \rightarrow 0$ in ${T_{\hat{w}_n}(S_{V_0}^+)}^*$ as $n \rightarrow \infty$. Therefore, from lemma 3.2 (c), it follows that the sequence $\left\{m_{V_0}(\hat{w}_n)\right\}_{n \in \mathbb{N}} \subset \mathcal{M}_{V_0}$ is the Palais–Smale sequence of the functional $I_{V_0}$ at the level $c_{V_0}$. Then, we can conclude that the sequence $\left\{m_{V_0}(\hat{w}_n)\right\}_{n \in \mathbb{N}} \subset {X_{V_0}}$ is the bounded Palais–Smale sequence at the level $c_{V_0}$. Now, let $v_n :=m_{V_0}(\hat{w}_n)$. According to theorem 3.5, we see that there exist a sequence $\left\{y_n\right\}_{n \in \mathbb{N}} \subset \mathbb{R}^N$ and $0 \neq \hat{v} \in {X_{V_0}}$ such that
Set $\widetilde{v}_n := \hat{v}_n - \hat{v}$. Applying lemma 2.10, the Hölder inequality, and (3.3), we can deduce that
Consequently, $\widetilde{v}_n \rightarrow 0$ in ${X_{V_0}}$ as $n \rightarrow \infty$, that is, $\hat{v}_n \rightarrow \hat{v}$ in ${X_{V_0}}$ as $n \rightarrow \infty$. From lemma 3.1 (c) and (3.3), we infer that $u_n(\cdot + y_n) \rightarrow \hat{v} \in \mathcal{M}_{V_0}$ as $n \rightarrow \infty$. So, we now complete the proof.
4. The barycenter map
Next, we need to establish a relationship between the topology of M and the number of positive solutions for problem (2.2). To this end, we first choose δ > 0 such that
and take $\eta \in C^\infty([0, +\infty),[0,1])$ non-increasing satisfying $\eta(t) = 1$ for $t \in (0,\delta / 2)$, $\eta(t) = 0$ for $t \in [\delta, + \infty)$, and $|\eta'(t)| \leqslant C$ for some constant C > 0.
Let ω be a positive ground state solution to the autonomous problem (3.1). For any $y \in M$, we define the following function:
with the unique number $t_\varepsilon \gt 0$ satisfying
and let us consider the mapping $\Phi_\varepsilon : M \mapsto \mathcal{N_\varepsilon}$ defined by
Lemma 4.1. The mapping $\Phi_\varepsilon$ verifies the following property:
Proof. Arguing by contradiction, we may assume that there exist $\delta_0 \gt 0$, $\left\{y_n\right\}_{n \in \mathbb{N}} \subset M$, and $\varepsilon_n \rightarrow 0$ in $\mathbb{R}$ as $n \rightarrow \infty$ such that
We first note that for each $ n \in \mathbb{N}$ and for all $ z \in B_{\frac{\delta}{\varepsilon_n}}$, $ \varepsilon_n z\in B_\delta$, and hence $\varepsilon_n z + y_n \in B_\delta(y_n) \subset M_\delta \subset \Lambda$.
Applying the change of variable $z = (\varepsilon_n x - y_n) / {\varepsilon_n}$ and recalling that G = F in Λ and $\eta(t) = 0$ for $t \geqslant \delta$, we have
Next, we prove that the sequence $\left\{t_{\varepsilon_n}\right\}_{n \in \mathbb{N}} \subset \mathbb{R}$ verifies $t_{\varepsilon_n} \rightarrow 1$ in $\mathbb{R}$ as $n \rightarrow \infty.$ Using the definition of $t_{\varepsilon_n}$, we see that $t_{\varepsilon_n} \Psi_{\varepsilon_n,y_n} \in \mathcal{N}_{\varepsilon_n}$, that is,
Taking into account $\eta (|x|) = 1$ for $x \in B_{\frac{\delta}{2}}$ and $B_{\frac{\delta}{2}} \subset B_{\frac{\delta}{\varepsilon_n}}$ for $n \in \mathbb{N}$ large enough, we deduce from (4.4) that
On the other hand, ω is a continuous and positive function in $\mathbb{R}^N$, so there exists $\bar{z} \in \mathbb{R}^N$ such that
Thus, we deduce from hypothesis $(f_4)$ that
According to the Dominated Convergence Theorem and lemma 2.2 in the study by Ambrosio [Reference Ambrosio7], we can deduce that
and
Therefore, if $t_{\varepsilon_n} \rightarrow + \infty$, we can conclude that the left-hand side of (4.5) satisfies the following property:
since q > p. On the other hand, we deduce from hypothesis $(f_3)$ that
From (4.5), (4.8), and (4.9) we get a contradiction. Then, we pass to a subsequence and assume that there exists t 0 such that $t_{\varepsilon_n} \rightarrow t_0 \geqslant 0$. In addition, applying (4.4) and (4.6) and combining with $(f_1)$–$(f_2)$, we deduce that $t_0 \gt 0$.
In (4.4), we pass to the limit as $n \rightarrow \infty$, and then we can use (4.6) and the Dominated Convergence Theorem to infer that
Due to $\omega \in \mathcal{M}_{V_0}$, we derive that
From the fact that the functions $\frac{F(t)}{t^{\frac{q}{2}}}$ and $\frac{f(t)}{t^{\frac{q}{2} - 1}}$ are increasing for t > 0, together with (4.10) and (4.11), it follows that $t_0 = 1$.
In (4.3), we pass to the limit as $n \rightarrow \infty$, and together with (4.7), we have
This is a contradiction, since (4.2). So, we complete the proof of the lemma.
Let us consider the function $h : \mathbb{R}^+ \mapsto \mathbb{R}^+$ such that $h(\varepsilon) := \sup_{y \in M} \left|J_\varepsilon(\Phi_\varepsilon(y)) - c_{V_0}\right|$ for all ɛ > 0. Then, we define the following subset of $\mathcal{N}_\varepsilon$:
From lemma 4.1, it follows that $h(\varepsilon) \rightarrow 0$ as $\varepsilon \rightarrow 0$. Additionally, we deduce from the definition of h that $\Phi_\varepsilon(y) \in \hat{\mathcal{N}}_\varepsilon$ for any $y \in M$ and ɛ > 0, and so $\hat{\mathcal{N}}_\varepsilon \neq \emptyset$.
For any δ > 0 given by (4.1), let us choose $\rho := \rho(\delta) \gt 0$ such that $M_\delta \subset B_\rho$. Then, we introduce the map $\zeta : \mathbb{R}^N \mapsto \mathbb{R}^N$ defined by
Now, we introduce the following barycenter map $\beta_\varepsilon : \mathcal{N}_\varepsilon \mapsto \mathbb{R}^N$ defined by
for all $u \in \mathcal{N}_\varepsilon$.
Then, from the above information, we can give the following lemma.
Lemma 4.2. The function βɛ has the following property:
Proof. Arguing by contradiction, assume that there exist $\delta_0 \gt 0$, $\left\{y_n\right\}_{n \in \mathbb{N}} \subset M$, and $\varepsilon_n \rightarrow 0$ in $\mathbb{R}$ as $n \rightarrow \infty$ such that
Employing the definitions of $\Phi_{\varepsilon_n}$, $\beta_{\varepsilon_n}$, and ζ and using the change of variable $z = (\varepsilon_n x - y_n) / \varepsilon_n$, we can conclude that
On account of $\left\{y_n\right\}_{n \in \mathbb{N}} \subset M \subset M_\delta$, combining the Dominated Convergence Theorem, we can derive that
which contradicts relation (4.12). This proof is now complete.
Lemma 4.3. Let $0 \lt \mu \lt sp$ and $\nu \lt (N - \mu) q / (N - sq).$ Assume that the sequences $\left\{\varepsilon_n\right\}_{n \in \mathbb{N}} \subset \mathbb{R}$ and $\left\{u_n\right\}_{n \in \mathbb{N}} \subset \mathcal{N}_{\varepsilon_n}$ satisfy $\varepsilon_n \rightarrow 0$ in $\mathbb{R}$ and $J_{\varepsilon_n}(u_n) \rightarrow c_{V_0}$ in $\mathbb{R}$ as $n \rightarrow \infty$; then, there is a sequence $\left\{\hat{y}_n\right\}_{n \in \mathbb{N}} \subset \mathbb{R}^N$ such that the sequence $\left\{\hat{u}_n(x) := u_n(x + \hat{y}_n)\right\}_{n \in \mathbb{N}}$ admits a subsequence which converges in ${X_{V_0}}.$ Furthermore, the sequence $ \left\{y_n := \varepsilon_n \hat{y}_n\right\}_{n \in \mathbb{N}} \subset \mathbb{R}^N$ has a subsequence $\left\{y_n\right\}_{n \in \mathbb{N}}$ $($still denoted by itself $)$ such that $y_n \rightarrow y_0 \in M$ as $n \rightarrow \infty.$
Proof. It is easy to verify that $\left\{u_n\right\}_{n \in \mathbb{N}} \subset X_{V_0}$ is bounded. Then, using lemma 3.4, we know that there exist a sequence $\left\{\hat{y}_n\right\}_{n \in \mathbb{N}} \subset \mathbb{R}^N$ and some constants $R,$ α > 0 such that
Let $\hat{u}_n(x) := u_n(x + \hat{y}_n)$. Consequently, $\left\{\hat{u}_n\right\}_{n \in \mathbb{N}} \subset {X_{V_0}}$ is bounded, and so passing to a subsequence, we can suppose that there exists some $0 \neq \hat{u} \in {X_{V_0}}$ such that $\hat{u}_n \xrightarrow w \hat{u}$ in ${X_{V_0}}$ as $n \rightarrow \infty$. Let $t_n \gt 0$ be such that $\hat{v}_n := t_n \hat{u}_n \in \mathcal{M}_{V_0}$, and let $y_n :=\varepsilon_n \hat{y}_n$. Thus, we deduce that
This implies that $I_{V_0}(\hat{v}_n) \rightarrow c_{V_0}$ in $\mathbb{R}$ as $n \rightarrow \infty$ and $\hat{v}_n \in \mathcal{M}_{V_0}.$ Clearly, the sequence $\left\{\hat{v}_n\right\}_{n \in \mathbb{N}} \subset {X_{V_0}}$ is bounded. Therefore, up to a subsequence if necessary, still denoted by itself, we may assume that there exists $\hat{v} \in {X_{V_0}}$ such that $\hat{v}_n \xrightarrow w \hat{v}$ in ${X_{V_0}}$ as $n \rightarrow \infty$. It is easy to see that the sequence $\left\{t_n\right\}_{n \in \mathbb{N}} \subset \mathbb{R}$ is bounded, and it holds that $t_n \rightarrow t_0 \geqslant 0$ as $n \rightarrow \infty$. Indeed, $t_0 \gt 0$. Otherwise, $t_0 = 0$, so, we infer from the boundedness of the sequence $\left\{\hat{v}_n\right\}_{n \in \mathbb{N}} \subset {X_{V_0}}$ that $\|\hat{v}_n\|_{X_{V_0}} = t_n \|\hat{u}_n\|_{X_{V_0}} \rightarrow 0$ in $\mathbb{R}$ as $n \rightarrow \infty$, and so $I_{V_0}(\hat{v}_n) \rightarrow 0$ in $\mathbb{R}$ as $n \rightarrow \infty$, but this is impossible, since $c_{V_0} \gt 0$. Thus, $t_0 \gt 0$. We deduce from the uniqueness of the weak limit that $\hat{v} = t_0 \hat{u}$ and $\hat{u} \neq 0$. Then, from lemma 3.6, it follows that
and so $\hat{u}_n \rightarrow \hat{u}$ in ${X_{V_0}}$ as $n \rightarrow \infty$. Moreover, $I_{V_0}(\hat{v}) = c_{V_0}$ and $\langle I_{V_0}'(\hat{v}), \hat{v} \rangle = 0$.
Next, we shall prove that the sequence $\left\{y_n\right\}_{n \in \mathbb{N}} \subset \mathbb{R}^N$ has a subsequence, still denoted by itself, such that $y_n \rightarrow y_0 \in M$ as $n \rightarrow \infty$. We first show the boundedness of the sequence $\left\{y_n\right\}_{n \in \mathbb{N}} \subset \mathbb{R}^N$. Otherwise, the sequence $\left\{y_n\right\}_{n \in \mathbb{N}} \subset \mathbb{R}^N$ is not bounded. So, we may assume that there exists a subsequence, still denoted by itself, such that $|y_n| \rightarrow + \infty$ in $\mathbb{R}$ as $n \rightarrow \infty$. Then, we choose R > 0 large enough such that $\Lambda \subset B_R$, we may suppose that $|y_n| \gt 2 R$ for $n \in \mathbb{N}$ sufficiently large, and so for all $x \in B_{R / {\varepsilon_n}}$ we have
On account of the facts that $u_n \in \mathcal{N}_{\varepsilon_n}$ and $J_{\varepsilon_n}(u_n) \rightarrow c_{V_0}$ in $\mathbb{R}$ as $n \rightarrow \infty$, thus, for $n \in \mathbb{N}$ large enough, we have that $u_n \in \mathcal{B}$. Consequently, from lemma 2.6, it follows that
Therefore, for $n \in \mathbb{N}$ large enough, we obtain
Recalling that $\hat{u}_n \rightarrow \hat{u}$ in ${X_{V_0}}$ as $n \rightarrow \infty$ and using the Dominated Convergence Theorem, we obtain
So, we have
Using $\hat{u}_n \rightarrow \hat{u} \neq 0$ in ${X_{V_0}}$ as $n \rightarrow \infty$ again, we see that this is a contradiction.
Now, we get the boundedness of the sequence $\left\{y_n\right\}_{n \in \mathbb{N}} \subset \mathbb{R}^N$. Passing to a subsequence (still denoted by $\left\{y_n\right\}_{n \in \mathbb{N}}$), we may assume that there exists $y_0 \in \mathbb{R}^N$ such that $y_n \rightarrow y_0 \in \overline{\Lambda}$ as $n \rightarrow \infty$. In fact, if $y_0 \not \in \overline{\Lambda}$, then we can find some constant r > 0 such that $y_n \in B_{r / 2}(y_0) \subset \overline{\Lambda}^c$. Arguing as before, we can reach a contradiction. So, $y_0 \in \overline{\Lambda}$.
It remains to show that $V(y_0) = V_0$. Arguing by contradiction, again we may assume that $V(y_0) \gt V_0$. From (4.13), together with Fatou’s lemma and the invariance of $\mathbb{R}^N$ by translations, it follows that
This is impossible. So, from hypothesis $(V_2)$, it follows that $y_0 \in M$. This proof is now complete.
Lemma 4.4. For any $\delta \gt 0,$ we have
Proof. Let $\varepsilon_n \rightarrow 0$ in $\mathbb{R}$ as $n \rightarrow \infty$, then we can find a sequence $\left\{u_n\right\}_{n \in \mathbb{N}} \subset \hat{\mathcal{N}}_{\varepsilon_n}$ such that
Noting that $\left\{u_n\right\}_{n \in \mathbb{N}} \subset \hat{\mathcal{N}}_{\varepsilon_n} \subset \mathcal{N}_{\varepsilon_n}$, we infer that
and so
Then, we deduce from lemma 4.3 that for $n \in \mathbb{N}$ large enough, there exists some sequence $\left\{\hat{y}_n\right\}_{n \in \mathbb{N}} \subset \mathbb{R}^N$ such that $y_n = \varepsilon_n \hat{y}_n \in M_\delta$. So, we have
Taking into account the facts that $\left\{\hat{u}_n(\cdot + \hat{y}_n)\right\}_{n \in \mathbb{N}} \subset {X_{V_0}}$ has a convergent subsequence and $\varepsilon_n z + y_n \rightarrow y_0 \in M$ as $n \rightarrow \infty$, we derive that $\beta_{\varepsilon_n}(u_n) = y_n + o_n(1)$ in $\mathbb{R}^N$ as $n \rightarrow \infty$. So, there exists a sequence $\left\{y_n\right\}_{n \in \mathbb{N}} \subset M_\delta$ such that
This ends the proof of the lemma.
5. Multiple solutions for problem (2.2)
In this section, we shall establish a relationship between the topology of M and the number of solutions for problem (2.2). Since $\mathcal{N}_\varepsilon$ is not a C 1 submanifold of the space Xɛ, we cannot use directly the standard Ljusternik–Schnirelmann theory, but we can bypass this difficulty by applying the abstract results in the study by Szulkin & Weth [Reference Szulkin, Weth, Gao and Motreanu55].
Theorem 5.1 Let $0 \lt \mu \lt sp$ and $\nu \lt (N - \mu) q / (N - sq).$ Assume that $(f_1)$–$(f_4)$ and $(V_1)$–$(V_2)$ are fulfilled, then for any δ > 0 satisfying $M_\delta \subset \Lambda,$ there exists $\hat{\varepsilon}_\delta \gt 0$ such that for any $\varepsilon \in (0,\hat{\varepsilon}_\delta)$ problem (2.2) possesses at least $cat_{M_\delta}(M)$ positive solutions.
Proof. For any fixed ɛ > 0, we introduce the mapping $\alpha_\varepsilon : M \mapsto S_\varepsilon^+$ defined by
Consequently, we deduce from lemma 4.1 that
Let us introduce the following function:
According to (5.1), we see that $h'(\varepsilon) \rightarrow 0$ in $\mathbb{R}$ as $\varepsilon \rightarrow 0$. Also, we define the following set:
Clearly, for all $y \in M$ and ɛ > 0, $\psi_\varepsilon(\alpha_\varepsilon(y)) \in \hat{S}_\varepsilon^+$, and so $\hat{S}_\varepsilon^+ \neq \emptyset$.
From lemma 4.1, lemma 2.2 (c), lemma 4.4, and lemma 4.2, it follows that there exists $\hat{\varepsilon} = \hat{\varepsilon}_\delta \gt 0$ such that the following diagram is well-defined:
Using lemma 4.2 and decreasing $\hat{\varepsilon}$ if necessary, for all $y \in M$, we have $\beta_\varepsilon(\Phi_\varepsilon(y)) = y + l(\varepsilon,y)$, where $|l(\varepsilon,y)| \leqslant \delta / 2$ uniformly in $y \in M$ and for all $\varepsilon \in (0,\hat{\varepsilon})$. Hence, $H(t,y) := y + (1- t) l(\varepsilon,y)$ for $(t,y) \in [0,1] \times M$ is homotopy between $\beta_\varepsilon \circ \Phi_\varepsilon = \left(\beta_\varepsilon \circ m_\varepsilon\right) \circ \left(m_\varepsilon^{- 1} \circ \Phi_\varepsilon\right)$ and the inclusion map $id : M \mapsto M_\delta$. This means that
Additionally, let us choose a function $h'(\varepsilon) \gt 0$ such that $h'(\varepsilon) \rightarrow 0$ in $\mathbb{R}$ as $\varepsilon \rightarrow 0$ and such that $c_{V_0} + h'(\varepsilon)$ is not a critical level for the functional Jɛ. Using corollary 2.9 and theorem 27 in the study by Szulkin and Weth [Reference Szulkin, Weth, Gao and Motreanu55], for ɛ > 0 sufficiently small, we can deduce that ψɛ possesses at least $cat_{\alpha_\varepsilon(M)} \alpha_\varepsilon(M)$ critical points on $\hat{S}_\varepsilon^+$. So, lemma 2.3 and (5.2) imply that the functional Jɛ has at least $cat_{M_\delta}(M)$ critical points in $\hat{\mathcal{N}}_\varepsilon$. This proof is now complete.
6. Proof of theorem 1.2
The main idea is to show that the solutions obtained in theorem 5.1 verify the following estimate:
Then, we can deduce from this fact that these solutions are indeed solutions of the original problem (1.3). To this end, we shall treat the regularity of non-negative solutions to problem (2.2). More precisely, we first establish the following result inspired by Moser [Reference Moser48] and Ambrosio and Rădulescu [Reference Ambrosio and Rădulescu11].
Lemma 6.1. Let $0 \lt \mu \lt sp$ and $\nu \lt (N - \mu) q / (N - sq)$, and let $\varepsilon_n \rightarrow 0$ in $\mathbb{R}$ as $n \rightarrow \infty$ and $u_n \in \hat{\mathcal{N}}_{\varepsilon_n}$ be a solution to problem (2.2). Then, $J_{\varepsilon_n}(u_n) \rightarrow c_{V_0}$ in $\mathbb{R}$ as $n \rightarrow \infty,$ and there exists a sequence $\left\{\hat{y}_n\right\}_{n \in \mathbb{N}} \subset \mathbb{R}^N$ such that $\hat{u}_n(\cdot) := u_n(\cdot + \hat{y}_n) \in L^\infty(\mathbb{R}^N)$ and $|\hat{u}_n|_{L^\infty(\mathbb{R}^N)} \leqslant C$ for all $n \in \mathbb{N},$ for some constant C > 0. Furthermore,
Proof. On account of the fact that $u_n \in \hat{\mathcal{N}}_{\varepsilon_n}$, arguing as in the proof of lemma 4.4, we see that $J_{\varepsilon_n}(u_n) \rightarrow c_{V_0}$ in $\mathbb{R}$ as $n \rightarrow \infty$. Then, from lemma 4.3, it follows that there exists a sequence $\left\{\hat{y}_n\right\}_{n \in \mathbb{N}} \subset \mathbb{R}^N$ such that $\hat{u}_n(\cdot) := u_n(\cdot + \hat{y}_n) \rightarrow \hat{u}(\cdot) \in {X_{V_0}}$ and $y_n :=\varepsilon_n \hat{y}_n \rightarrow y_0 \in M$ as $n \rightarrow \infty$. For any L > 0 and β > 1, we introduce the function
Taking $\psi(\hat{u}_n)$ as test function, we have
Additionally, applying the boundedness of the sequence $\left\{\hat{u}_n\right\}_{n \in \mathbb{N}} \subset X_{\varepsilon_n}$ and arguing as in the proof of lemma 2.6, we can deduce that there exists $C_0 \gt 0$ such that
According to the hypotheses on g, we see that for any σ > 0 there exists $C_\sigma \gt 0$ such that
Choosing $\sigma \in (0,V_0/{C_0})$, together with the above inequalities, we can infer that
for some constant C > 0.
Let us define the following functions:
We first observe that ψ is an increasing function, and hence
Then, we can infer from (6.3) and the Jensen inequality that
We also point out that
So, putting together with (6.2), (6.3), (6.4), and (6.5) and using the Sobolev embedding, we infer that there exists some constant C > 0 such that
Choose $\beta = \frac{q_s^*}{q}$, and let R > 0 large enough. Combining $\hat{u}_n \rightarrow \hat{u}$ in ${X_{V_0}}$ as $n \rightarrow \infty$ with the Hölder inequality, we can conclude that there exists some constant C > 0 such that
Then, we choose a fixed $\epsilon \in (0,1 / C)$ and infer that
In the above inequality, we pass to the limit as $L \rightarrow + \infty$ and we obtain $\hat{u}_n \in L^{\frac{{q_s^*}^2}{q}}(\mathbb{R}^N)$.
Thanks to $0 \leqslant \hat{u}_{n,L} \leqslant \hat{u}_n$, then in (6.6), we pass to the limit as $L \rightarrow + \infty$ and we have
This means that
For $1 \leqslant m \in \mathbb{N}$, let us define
It follows that
and so
Let us define
Then, we have
Consequently, using a standard iteration argument, we have
In the above inequality, we pass to the limit as $m \rightarrow \infty$ and then we infer that $|\hat{u}_n|_{L^\infty(\mathbb{R}^N)} \leqslant C$ uniformly in $n \in \mathbb{N}$.
Next, let us define
We point out that $\hat{u}_n$ satisfies the following equation:
From the growth hypotheses on g, corollary 2.1 in the study by Ambrosio & Rădulescu [Reference Ambrosio and Rădulescu11], $\hat{u}_n \rightarrow \hat{u}$ in ${X_{V_0}}$ as $n \rightarrow \infty$, and the uniformly boundedness of the sequence $\left\{\hat{u}_n\right\}_{n \in \mathbb{N}}$ in $L^\infty(\mathbb{R}^N) \cap {X_{V_0}}$, we can conclude that $\hat{u}_n(x) \rightarrow 0$ in $\mathbb{R}$ as $|x| \rightarrow +\infty$ uniformly with respect to $n \in \mathbb{N}$. This ends the proof of the lemma.
Proof of theorem 1.2 completed
We first choose δ > 0 small enough such that $M_\delta \subset \Lambda$. Then, we claim that there exists $\bar{\varepsilon}_\delta \gt 0$ such that for any $\varepsilon \in (0,\bar{\varepsilon}_\delta)$ and any solution $u_\varepsilon \in \hat{\mathcal{N}}_\varepsilon$ of problem (2.2), we have
Otherwise, we may assume that there exists a subsequence $\left\{\varepsilon_n\right\}_{n \in \mathbb{N}} \subset \mathbb{R}$ such that $\varepsilon_n \rightarrow 0$ as $n \rightarrow \infty$, $u_{\varepsilon_n} \in \hat{\mathcal{N}}_{\varepsilon_n}$ such that $J_{\varepsilon_n}'(u_{\varepsilon_n})=0$ and
But we see that $J_{\varepsilon_n}(u_{\varepsilon_n}) \rightarrow c_{V_0}$ in $\mathbb{R}$ as $n \rightarrow \infty$. Thus, from lemma 4.3, it follows that there exists a sequence $\left\{\hat{y}_n\right\}_{n \in \mathbb{N}} \subset \mathbb{R}^N$ such that $\hat{u}_n(\cdot) :=u_{\varepsilon_n}(\cdot + \hat{y}_n) \rightarrow \hat{u}(\cdot)$ in ${X_{V_0}}$ and $\varepsilon_n \hat{y}_n \rightarrow y_0 \in M$ as $n \rightarrow \infty$.
Next, we choose r > 0 such that $B_r(y_0) \subset B_{2r}(y_0) \subset \Lambda$, and so $B_{\frac{r}{\varepsilon_n}}\left(\frac{y_0}{\varepsilon_n}\right) \subset \Lambda_{\varepsilon_n}$. Furthermore, for n large enough, we can deduce that $\Lambda_{\varepsilon_n}^c \subset B_{\frac{r}{\varepsilon_n}}^c(\hat{y}_n)$. In addition, from (6.1), we see that $\hat{u}_n(x) \rightarrow 0$ as $|x| \rightarrow + \infty$ uniformly in $n \in \mathbb{N}$. Therefore, we can find R > 0 such that $\hat{u}_n(x) \lt a$ for any $|x| \geqslant R$, $n \in \mathbb{N}$. Consequently, $u_{\varepsilon_n}(x) \lt a$ for any $x \in B_R^c(\hat{y}_n)$, $n \in \mathbb{N}$. Moreover, for $n \in \mathbb{N}$ sufficiently large, we know that
Thus, we infer that $u_{\varepsilon_n}(x) \lt a$ for any $x \in \Lambda_{\varepsilon_n}^c$ and for all $n \in \mathbb{N}$ large enough, which contradicts relation (6.8).
Fix $\varepsilon \in (0,\varepsilon_\delta)$, where $\varepsilon_\delta := \min\{\hat{\varepsilon}_\delta, \bar{\varepsilon}_\delta\}$. From theorem 5.1, we can see that problem (2.2) has at least $\text{cat}_{M_\delta}(M)$ nontrivial solutions. Now, we use uɛ to denote one of these solutions, and so $u_\varepsilon \in \hat{\mathcal{N}}_\varepsilon$. Then, using (6.7) and recalling that the definitions of g and G, we can also infer that uɛ is a solution of problem (1.3). So, problem (1.3) possesses at least $\text{cat}_{M_\delta}(M)$ nontrivial solutions.
Finally, we establish the behaviour of the maximum points of solutions to problem (1.3). Let us choose $\varepsilon_n \rightarrow 0$ and consider a sequence $\left\{u_n\right\}_{n \in \mathbb{N}} \subset X_{\varepsilon_n}$ of solutions for problem (1.3) as before. From $(g_1)$, it follows that there exists a positive constant $\iota \lt a$ such that
since $u_n \in \mathcal{N}_{\varepsilon_n}$ and $J_{\varepsilon_n}(u_n) \rightarrow c_{V_0}$ in $\mathbb{R}$ as $n \rightarrow \infty$. Then, for $n \in \mathbb{N}$ sufficiently large, we have that $u_n \in \mathcal{B}$. Consequently, from lemma 2.6, it follows that
Arguing as before, there exists R > 0 such that
Furthermore, up to a subsequence, we may assume that
Otherwise, if relation (6.12) does not hold, we deduce from (6.11) that $|u_n|_{L^\infty(\mathbb{R}^N)} \lt \iota$. Using $u_n \in \mathcal{N}_{\varepsilon_n}$ again and (6.9)–(6.10), we have
This implies that $\|u_n\|_{X_{\varepsilon_n}} = 0$ for all $n \in \mathbb{N}$ sufficiently large, which is a contradiction. Consequently, relation (6.12) holds true. On account of (6.11) and (6.12), we can infer that if pn is a global maximum point of un and $p_n = \hat{y}_n + q_n$ for some $q_n \in B_R$, $\varepsilon_n p_n \rightarrow y_0 \in M$ as $n \rightarrow \infty$; then using the continuity of the potential V we see that $V(\varepsilon_n p_n) \rightarrow V(y_0) = V_0$ in $\mathbb{R}$ as $n \rightarrow \infty$.
This proof is nowcomplete. $\square$
Declarations
Conflict of interest
The authors declare that they have no competing interests.
Data availability
Data sharing is not applicable to this article as no datasets were generated or analysed during the current study.
Funding
The research of V.D. Rădulescu was supported by the grant ‘Nonlinear Differential Systems in Applied Sciences’ of the Romanian Ministry of Research, Innovation and Digitalization, within PNRR-III-C9-2022-I8/22. This research was partially supported by the Scientific Research Fund of Hunan Provincial Education Department (No. 20B243), the National Natural Science Foundation of China (Nos. 12301219 and 12171486), the Science and Technology Innovation Program of Hunan Province (No. 2024RC3021), the Project for Young Backbone Teachers of Hunan Province (No.10900-150220002), and the Natural Science Foundation for Excellent Young Scholars of Hunan Province (No. 2023JJ20057).