Loading [MathJax]/jax/element/mml/optable/Arrows.js
Research article

Existence and concentration of positive solutions for a p-fractional Choquard equation

  • Received: 28 June 2021 Accepted: 06 September 2021 Published: 13 September 2021
  • MSC : 35A15, 35B38, 35J60

  • In this work, we study the existence, multiplicity and concentration behavior of positive solutions for the following problem involving the fractional p-Laplacian

    εps(Δ)spu+V(x)|u|p2u=εμN(1|x|μK|u|q)K(x)|u|q2uinRN,

    where 0<s<1<p<, N>ps, 0<μ<ps, p<q<ps2(2μN), (Δ)sp is the fractional p-Laplacian and ε>0 is a small parameter. Under certain conditions on V and K, we prove the existence of a positive ground state solution and express the location of concentration in terms of the potential functions V and K. In particular, we relate the number of solutions with the topology of the set where V attains its global minimum and K attains its global maximum.

    Citation: Xudong Shang. Existence and concentration of positive solutions for a p-fractional Choquard equation[J]. AIMS Mathematics, 2021, 6(11): 12929-12951. doi: 10.3934/math.2021748

    Related Papers:

    [1] Liyan Wang, Baocheng Zhang, Zhihui Lv, Kun Chi, Bin Ge . Existence of solutions for the fractional $ p&q $-Laplacian equation with nonlocal Choquard reaction. AIMS Mathematics, 2025, 10(4): 9042-9054. doi: 10.3934/math.2025416
    [2] Meixia Cai, Hui Jian, Min Gong . Global existence, blow-up and stability of standing waves for the Schrödinger-Choquard equation with harmonic potential. AIMS Mathematics, 2024, 9(1): 495-520. doi: 10.3934/math.2024027
    [3] Kexin Ouyang, Yu Wei, Huiqin Lu . Positive ground state solutions for a class of fractional coupled Choquard systems. AIMS Mathematics, 2023, 8(7): 15789-15804. doi: 10.3934/math.2023806
    [4] Giulio Romani . Choquard equations with critical exponential nonlinearities in the zero mass case. AIMS Mathematics, 2024, 9(8): 21538-21556. doi: 10.3934/math.20241046
    [5] Zhiwei Lv, Jiafa Xu, Donal O'Regan . Solvability for a fractional $ p $-Laplacian equation in a bounded domain. AIMS Mathematics, 2022, 7(7): 13258-13270. doi: 10.3934/math.2022731
    [6] Kirti Kaushik, Anoop Kumar, Aziz Khan, Thabet Abdeljawad . Existence of solutions by fixed point theorem of general delay fractional differential equation with $ p $-Laplacian operator. AIMS Mathematics, 2023, 8(5): 10160-10176. doi: 10.3934/math.2023514
    [7] Jinguo Zhang, Dengyun Yang, Yadong Wu . Existence results for a Kirchhoff-type equation involving fractional $ p(x) $-Laplacian. AIMS Mathematics, 2021, 6(8): 8390-8403. doi: 10.3934/math.2021486
    [8] Xianyong Yang, Qing Miao . Asymptotic behavior of ground states for a fractional Choquard equation with critical growth. AIMS Mathematics, 2021, 6(4): 3838-3856. doi: 10.3934/math.2021228
    [9] M. J. Huntul . Inverse source problems for multi-parameter space-time fractional differential equations with bi-fractional Laplacian operators. AIMS Mathematics, 2024, 9(11): 32734-32756. doi: 10.3934/math.20241566
    [10] Lulu Tao, Rui He, Sihua Liang, Rui Niu . Existence and multiplicity of solutions for critical Choquard-Kirchhoff type equations with variable growth. AIMS Mathematics, 2023, 8(2): 3026-3048. doi: 10.3934/math.2023156
  • In this work, we study the existence, multiplicity and concentration behavior of positive solutions for the following problem involving the fractional p-Laplacian

    εps(Δ)spu+V(x)|u|p2u=εμN(1|x|μK|u|q)K(x)|u|q2uinRN,

    where 0<s<1<p<, N>ps, 0<μ<ps, p<q<ps2(2μN), (Δ)sp is the fractional p-Laplacian and ε>0 is a small parameter. Under certain conditions on V and K, we prove the existence of a positive ground state solution and express the location of concentration in terms of the potential functions V and K. In particular, we relate the number of solutions with the topology of the set where V attains its global minimum and K attains its global maximum.



    In this paper, we consider the following nonlinear equation governed by the fractional p-Laplacian

    εps(Δ)spu+V(x)|u|p2u=εμN(1|x|μK|u|q)K(x)|u|q2uinRN, (1.1)

    where ε>0 is a small parameter, N>ps, s(0,1), 1<p<, p<q<ps2(2μN) and V,K are positive functions. (Δ)sp denotes the fractional p-Laplacian defined for all u:RNR smooth enough by

    (Δ)spu(x)=P.V.RN|u(x)u(y)|p2(u(x)u(y))|xy|N+psdy,

    the P.V. stands for the Cauchy principle value (see [10]).

    In the case s=1, p=2 and K(x)1, the Eq (1.1) boils down to the Choquard equation

    ε2Δu+V(x)u=εμN(1|x|μ|u|q)|u|q2uinRN. (1.2)

    When N=3, μ=1 and q=2, the Eq (1.2) has appeared in the several context of quantum physics, such as the description of a Polaron at rest [20] and the model of an electron trapped in its own hole [13] and the coupling of the Schrödinger equation under a classical Newtonian gravitational potential [12]. The pioneering mathematical research goes back to Lieb [13] and Lions [14]. The existence and qualitative of solutions of equation like (1.2) have been extensively studied by variational methods, see for example [16,17,18,19,29] and their references. For the existence of semi-classical solutions to Choquard equation (1.2) were studied in some papers. In [28], Wei and Winter constructed a family of solutions which concentrate to the non-degenerate critical points of the potential V. Moroz and Schaftingen [18] proved the existence of solutions concentrating around the local minimum of V by a nonlocal penalization method. See [3] for the existence and multiplicity for a generalized quasilinear Choquard equation.

    In recent years, a great attention has been given to problems driven by the fractional Laplacian. One of the reasons for this comes from the fact that this operator appears in several applications in different subjects, such as crystal dislocation, thin obstacle problems, optimization and finance, anomalous diffusion and many others, we can see [10,25]. Recently, d'Avenia, Siciliano and Squassina [21] considered the existence, regularity, symmetry as well as decay properties of the following fractional Choquard equation

    (Δ)su+au=εμN(1|x|μ|u|q)|u|q2uinRN. (1.3)

    Shen, Gao and Yang [23] obtained the existence of ground states of (1.3) with the nonlinearity satisfies the generaal Beresty-Lions type assumptions. Zhang and Wu [31] studied the existence of nodal solutions of (1.3). Chen and Liu [5] studied (1.3) with nonconstant linear potential and proved the existence of ground states without any symmetry property. Ambrosio [1] investigated the multiplicity and concentration of positive solutions for a fractional Choquard equation with general nonlinearity. In [7], Chen, Li and Yang obtained the multiplicity and concentration of nontrivial nonnegative solutions for a fractional Choquard equation with critical exponent. For other existence results we refer to [4,11,24,32] and the references therein.

    To the best of our knowledge, there are few results about fractional Choquard equation like (1.3). Belchior et al. [8] investigated the following equation

    (Δ)spu+A|u|p2u=(1|x|μF(u))f(u)inRN, (1.4)

    where F is the primitive of f and A is a positive constant. They showed the existence of ground states and asymptotic of the solutions for (1.4). In [2], Ambrosio studied the following problem

    εps(Δ)spu+V(x)|u|p2u=εμN(1|x|μF(u))f(u)inRN.

    He proved solutions concentrating around global minimum of the potential V.

    Recently, Wang et al. [27] applied a kind of structure introduced by Ding and Liu [9] to study the existence and concentration of positive solutions for semilinear Schrödinger-Poisson system. The similar results for the fractional Schrödinger-Poisson system, we can see [30]. Alves and Yang [3] considered the generalized quasilinear Choquard equation

    εpΔpu+V(x)|u|p2u=εμN(RNQ(y)F(u(y))|xy|μ)Q(x)f(u)inRN,

    where 1<p<N, V and Q are two continuous functions satisfy the structure of [9], they established concentration behavior for the Choquard equation. It is quite natural to ask how the potentials will affect the existence and concentration of solutions for (1.1). In this paper, we shall give an affirmative answers for this question.

    Motivated by the above papers, we will establish the existence, multiplicity and concentration of positive solutions for Eq (1.1). To gain further insight into the effect of potential functions V and K on the concentration process, we give the following assumptions introduced by [9]. Set

    θ=minxRNV(x),V={xRN:V(x)=θ},V=lim inf|x|V(x),
    κ=maxxRNK(x),K={xRN:K(x)=κ},K=lim sup|x|K(x).

    We assume that V and K satisfy:

    (H1) V,KL(RN) are uniformly continuous and θ>0, infxRNK(x)>0.

    (H2) θ<V< and there exist R>0,xV such that

    K(x)K(x)for all|x|R.

    (H3) κ>KinfxRNK(x) and there exist R>0,xK such that

    V(x)V(x)for all|x|R.

    From (H2), we may assume that K(x)=maxxVK(x). Set

    ΩV={xV:K(x)=K(x)}{xV:K(x)>K(x)}.

    From (H3), we may assume that V(x)=minxKV(x). Set

    ΩK={xK:V(x)=V(x)}{xK:V(x)<V(x)}.

    Clearly, ΩV and ΩK are bounded sets. Moreover, if VK, then ΩV=ΩK=VK. In particular, ΩV=V if K(x) is a constant, and ΩK=K if V(x) is a constant.

    We now state our main results.

    Theorem 1.1. Assume that (H1) and (H2) hold, then for all small ε>0, (1.1) has a positive ground state solution uε, and there exists a maximum point xε, such that up to a subsequence, xεx0 as ε0, limε0dist(xε,ΩV)=0, and vε(x)=uε(εx+xε) converges in Ws,p(RN) to a ground state solution of

    (Δ)spu+V(x0)|u|p2u=K2(x0)(RN|u(y)|q|xy|μdy)|u|q2u,xRN.

    In particular, if VK, then limε0dist(xε,VK)=0, and up to a subsequence, vε converges in Ws,p(RN) to a ground state solution of

    (Δ)spu+θ|u|p2u=κ2(RN|u(y)|q|xy|μdy)|u|q2u,xRN.

    If (H1) and (H3) hold, and we replace ΩV by ΩK, then all the conclusions remain true.

    Let VK. Now we denote Λ=VK. It is easy to check that Λ is compact. For any δ>0, set Λδ={xRN:dist(x,Λ)δ}.

    Theorem 1.2. Assume that (H1) and (H2) or (H3) hold, then for all small ε>0, problem (1.1) has at least catΛδ(Λ) solutions, if xε its global maximum, up to a subsequence, such that limε0dist(xε,Λ)=0, uε converges in Ws,p(RN) to a ground state solution of

    (Δ)spu+θ|u|p2u=κ2(RN|u(y)|q|xy|μdy)|u|q2u,xRN.

    Note that our main results are also new for the case p=2. Our main theorem improves the result in [2,8] with both linear potential V and nonlinear potential K of the concentration behavior of positive solutions. There are some difficulties in such a problem. The first one is that there would presumably be competition between the V and K: each would try to attract ground states to their minimum and maximum points, respectively. The second one, the operator (Δ)sp and the convolution term are all nonlocal operators, make our analysis more complicated with respect to [3], so we need more accurate estimates.

    The plan of this paper is the following: In Section 2, we give some preliminary results which will be used later. In Section 3, we show some compactness lemmas of the functional associated to our problem. In Section 4, we consider the existence of ground states of case of (1.1) and the concentration phenomenon. In the final section, we prove Theorem 1.2.

    In this paper, we will use the following notations:

    The notations C, C1,C2 are positive (possibly different) constants.

    Br(z0) denotes the ball in RN centered at z0 with radius r.

    on(1) and oε(1) denotes the vanishing quantities as n and ε0.

    We will use q for the norm in Lq(RN), u+=max{u,0} and u=min{u,0}.

    In this section, we recall some known results for the readers convenience and the later use. First, we will give some useful facts for the fractional order Sobolev spaces. Let 0<s<1<p< be real numbers, the homogeneous fractional Sobolev space Ds,p(RN) as the completion of C0(RN) with respect to the Gagliardo seminorm

    [u]ps,p=R2N|u(x)u(y)|p|xy|N+psdxdy.

    The fractional Sobolev space Ws,p(RN) is defined as

    Ws,p(RN)={uLp(RN):R2N|u(x)u(y)|p|xy|N+psdxdy<},

    equipped with the norm

    up=[u]ps,p+upp.

    It is easy to see that the embedding Ws,p(RN)Lr(RN) is continuous for any r[p,ps], and compactly in Lrloc(RN) for any r[p,ps).

    Making the change of variable xεx, Eq (1.1) becomes

    (Δ)spu+Vε(x)|u|p2u=(RNKε(y)|u(y)|q|xy|μdy)Kε(x)|u|q2u,xRN, (2.1)

    where Vε(x)=V(εx) and Kε(x)=K(εx). Eqs (1.1) and (2.1) are equivalent, we shall thereafter focus on Eq (2.1). For any ε>0, let Eε={uWs,p(RN):RNVε(x)|u|pdx<} be the Sobolev space endowed with the norm

    upε=R2N|u(x)u(y)|p|xy|N+psdxdy+RNVε(x)|u|pdx.

    By the assumption of V, we see that ε and are equivalent norms for ε>0. Define the energy functional associated with (2.1) by

    Iε(u)=1pupε12qR2NKε(y)|u(y)|qKε(x)|u(x)|q|xy|μdxdy.

    Note that p<q<ps2(2μN), by the Hardy-Littlewood-Sobolev inequality ([15]) and the boundedness of K, we have

    R2NKε(y)|u(y)|qKε(x)|u(x)|q|xy|μdxdyC1(RN|u|2Nμ2Nμdx)2NμNC2u2qε. (2.2)

    Therefore, the functional Iε is well defined on Eε and belongs to C1(Eε,R).

    Define the solution manifold of (2.1) by

    Nε={uEε{0}:upε=R2NKε(y)|u(y)|qKε(x)|u(x)|q|xy|μdxdy}.

    For any uNε, by (2.2) we have

    uεr, (2.3)

    for some r>0.

    The ground energy associated with (2.1) is defined as

    cε=infuNεIε(u).

    The following vanishing lemma is a version of the Concentration-compactness principle of P. L. Lions. We can see (Lemma 2.1 of [2]).

    Lemma 2.1. Let N>ps. Assume that {un} is bounded in Ws,p(RN) and it satisfies

    limnsupyRNBR(y)|un(x)|pdx=0,

    for some R>0. Then un0 strongly in Lr(RN) for every r(p,ps).

    From (2.2) and q>p, it follows that Iε satisfies the geometry of the mountain pass (see [26]). Hence, there is a sequence {un}Eε such that

    Iε(un)cεandIε(un)0, (2.4)

    where cε is the mountain pass level given by

    cε=infγΓsupt[0,1]Iε(γ(t))>0,

    and Γ={γC1([0,1],Eε):γ(0)=0,Iε(γ(1))<0}.

    We observe that for any uEε{0}, there exists a unique tu>0 such that tuuNε, and the maximum of the function g(t)=Iε(tu) for t0 is achieved at t=tu. By a standard arguments, we have

    cε=cε=infuEε{0}maxt0Iε(tu).

    For any a,b>0, consider the limit problem

    (Δ)spu+a|u|p2u=b2(RN|u(y)|q|xy|μdy)|u|q2u,xRN. (2.5)

    Solutions of (2.5) are critical points of the functional defined by

    Iab(u)=1p[u]ps,p+apRN|u|pdxb22qR2N|u(y)|q|u(x)|q|xy|μdxdy.

    Define the solution manifold of (2.5) by

    Mab={uWs,p(RN){0}:Iab(u),u=0}.

    The ground energy associated with (2.5) is defined as cab=infuMabIab(u). It is easy to check that

    cab=infuWs,p(RN){0}maxt0Iab(tu).

    By [8], we known that (2.5) has a positive ground state solution ω, that is cab=Iab(ω).

    Lemma 2.2. Let a1,a2>0 and b1,b2>0, with a1a2 and b1b2. Then ca1b1ca2b2. In particular, if one of inequalities is strict, then ca1b1<ca2b2.

    Proof. Let uMa2b2 be a ground state solution of (2.5) with coefficients a2,b2 such that

    ca2b2=Ia2b2(u)=maxt0Ia2b2(tu). (2.6)

    It is easy to check that there exists t0>0 such that t0uMa1b1. Then we get

    Ia1b1(t0u)=maxt0Ia1b1(tu). (2.7)

    It follows from (2.6) and (2.7) that

    ca2b2=Ia2b2(u)Ia2b2(t0u)=Ia1b1(t0u)+tp0p(a2a1)RN|u|pdx+t2q02q(b21b22)R2N|u(y)|q|u(x)|q|xy|μdxdyIa1b1(t0u)infvMa1b1Ia1b1(v)=ca1b1.

    The proof is completed.

    In this section we will show some compactness results for the functional Iε.

    Lemma 3.1. {un}Eε is a (PS)c sequence for Iε with un0 weakly in Eε. If un in E_{\varepsilon} , the c\geq c_{\infty}: = c_{V_{\infty}K_{\infty}} .

    Proof. Let \{u_{n}\} be a (PS)_{c} sequence for I_{\varepsilon} , by (2.4), we have

    \begin{eqnarray} c+1+\|u_{n}\|_{\varepsilon}\geq I_{\varepsilon}(u_{n}) - \frac{1}{2q}\langle I'_{\varepsilon}(u_{n}), u_{n}\rangle = (\frac{1}{p} - \frac{1}{2q})\|u_{n}\|^{p}_{\varepsilon} \end{eqnarray} (3.1)

    for n large enough. Therefore \{u_{n}\} is bounded in E_{\varepsilon} .

    For each n , there is a unique t_{n} > 0 such that t_{n}u_{n}\in \mathcal{M}_{V_{\infty}K_{\infty}} . We now show that the sequence \{t_{n}\} satisfies \limsup_{n\rightarrow \infty} t_{n} \leq 1 . By contradiction we assume that there exist \sigma > 0 and a subsequence (still denoted by \{t_{n}\} ) such that t_{n} \geq 1+ \sigma for all n . From the boundedness of \{u_{n}\} , we have \langle I'_{\varepsilon}(u_{n}), u_{n}\rangle = o_{n}(1) . That is

    \begin{eqnarray} [u_{n}]^{p}_{s, p} + \int_{\mathbb{R}^{N}}V_{\varepsilon}( x)|u_{n}|^{p}dx = \int_{\mathbb{R}^{2N}} \frac{K_{\varepsilon}(y)|u_{n}(y)|^{q}K_{\varepsilon}( x)|u_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy +o_{n}(1). \end{eqnarray} (3.2)

    Since t_{n}u_{n}\in \mathcal{M}_{V_{\infty}K_{\infty}} , we obtain

    \begin{eqnarray} t^{p}_{n}([u_{n}]^{p}_{s, p} + \int_{\mathbb{R}^{N}}V_{\infty}|u_{n}|^{p}dx) = t^{2q}_{n}K^{2}_{\infty}\int_{\mathbb{R}^{2N}} \frac{|u_{n}(y)|^{q}|u_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy. \end{eqnarray} (3.3)

    We deduce from (3.2) and (3.3) that

    \begin{eqnarray} \int_{\mathbb{R}^{N}}(V_{\infty}-V_{\varepsilon}(x))|u_{n}|^{p}dx = \int_{\mathbb{R}^{2N}}\frac{(t^{2q-p}_{n}K^{2}_{\infty}-K_{\varepsilon}(y)K_{\varepsilon}(x))|u_{n}(y)|^{q}|u_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy. \end{eqnarray} (3.4)

    By the definition of V_{\infty} and K_{\infty} , for any \nu > 0 , there exist a constant \rho > 0 sufficiently large such that for |x| > \rho ,

    \begin{eqnarray} V(x) > V_{\infty} - \nu, \hskip0.5cm K(x) < K_{\infty} + \nu. \end{eqnarray} (3.5)

    Since \{u_{n}\} is bounded and u_{n}\rightharpoonup 0 in E_{\varepsilon} , by (3.5) we have

    \begin{eqnarray} \int_{\mathbb{R}^{N}}(V_{\infty}-V_{\varepsilon}(x))|u_{n}|^{p}dx & = &\int_{|x|\leq \frac{\rho}{\varepsilon}}(V_{\infty}-V_{\varepsilon}(x))|u_{n}|^{p}dx\\ &&+ \int_{|x| > \frac{\rho}{\varepsilon}}(V_{\infty}-V_{\varepsilon}(x))|u_{n}|^{p}dx\\ &\leq & o_{n}(1) +C\nu. \end{eqnarray} (3.6)

    On the other hand,

    \begin{eqnarray} &&\int_{\mathbb{R}^{2N}} \frac{K_{\varepsilon}( y)|u_{n}(y)|^{q}K_{\varepsilon}( x)|u_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy\\ && = \int_{|x| > \frac{\rho}{\varepsilon}}\int_{|y| > \frac{\rho}{\varepsilon}} \frac{K_{\varepsilon}(y)|u_{n}(y)|^{q}K_{\varepsilon}( x)|u_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy\\ &&+\int_{|x| > \frac{\rho}{\varepsilon}}\int_{|y|\leq \frac{\rho}{\varepsilon}} \frac{K_{\varepsilon}( y)|u_{n}(y)|^{q}K_{\varepsilon}(x)|u_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy\\ &&+\int_{|x|\leq \frac{\rho}{\varepsilon}}\int_{|y| > \frac{\rho}{\varepsilon}} \frac{K_{\varepsilon}( y)|u_{n}(y)|^{q}K_{\varepsilon}( x)|u_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy\\ &&+\int_{|x|\leq \frac{\rho}{\varepsilon}}\int_{|y|\leq \frac{\rho}{\varepsilon}} \frac{K_{\varepsilon}( y)|u_{n}(y)|^{q}K_{\varepsilon}( x)|u_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy \\ && = I+II+III+IV. \end{eqnarray} (3.7)

    By (2.2), (3.5) and the boundedness of \{u_{n}\} , we obtain

    \begin{eqnarray} I & < & (K_{\infty} + \nu)^{2}\int_{|x| > \frac{\rho}{\varepsilon}}\int_{|y| > \frac{\rho}{\varepsilon}} \frac{|u_{n}(y)|^{q}|u_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy\\ &\leq &(K_{\infty} + \nu)^{2}\int_{\mathbb{R}^{2N}} \frac{|u_{n}(y)|^{q}|u_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy\\ &\leq &K^{2}_{\infty}\int_{\mathbb{R}^{2N}} \frac{|u_{n}(y)|^{q}|u_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy + C\nu +C\nu^{2}. \end{eqnarray} (3.8)

    From the boundedness of K(x) and \{u_{n}\} , there is a constant C > 0 such that

    \begin{eqnarray} \int_{\mathbb{R}^{N}} \frac{K_{\varepsilon}( y)|u_{n}(y)|^{q}}{|x-y|^{\mu}}dxdy \leq C. \end{eqnarray} (3.9)

    By (3.9) and u_{n}\rightharpoonup 0 , we have

    \begin{eqnarray} II &\leq& \int_{|y|\leq \frac{\rho}{\varepsilon}}\int_{\mathbb{R}^{N}} \frac{K(\varepsilon y)|u_{n}(y)|^{q}K(\varepsilon x)|u_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy\\ &\leq & C \int_{|y|\leq \frac{\rho}{\varepsilon}}|u_{n}|^{q}dy = o_{n}(1). \end{eqnarray} (3.10)

    Similarly, we have

    \begin{eqnarray} III = o_{n}(1)\hskip0.5cm \text{and}\hskip0.5cm IV = o_{n}(1). \end{eqnarray} (3.11)

    By (3.7), (3.8), (3.10) and (3.11), we deduce that

    \begin{eqnarray} \int_{\mathbb{R}^{2N}} \frac{K_{\varepsilon}( y)|u_{n}(y)|^{q}K_{\varepsilon} (x)|u_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy &\leq & K^{2}_{\infty}\int_{\mathbb{R}^{2N}} \frac{|u_{n}(y)|^{q}|u_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy\\ && + C\nu +C\nu^{2} + o_{n}(1). \end{eqnarray} (3.12)

    Combining (3.6), (3.12) with (3.4), we obtain

    \begin{eqnarray} K^{2}_{\infty}(t^{2q-p}_{n} -1)\int_{\mathbb{R}^{2N}} \frac{|u_{n}(y)|^{q}|u_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy \leq C\nu +C\nu^{2} + o_{n}(1). \end{eqnarray} (3.13)

    From u_{n}\not\rightarrow 0 in E_{\varepsilon} , there exists a sequence \{z_{n}\} \subset \mathbb{R}^{N} and constant R, \beta > 0 such that

    \begin{eqnarray} \liminf\limits_{n\rightarrow \infty}\int_{B_{R}(z_{n})}|u_{n}|^{p}dx \geq \beta > 0. \end{eqnarray} (3.14)

    Indeed, if (3.14) does not true, Lemma 2.1 implies that u_{n}\rightarrow 0 in L^{r}(\mathbb{R}^{N}) for every r\in (p, p_{s}^{*}) . It follows from (2.2) that

    \begin{eqnarray*} \int_{\mathbb{R}^{2N}} \frac{K_{\varepsilon}( y)|u_{n}(y)|^{q}K_{\varepsilon}( x)|u_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy = o_{n}(1). \end{eqnarray*}

    This and (3.2) implies \|u_{n}\|_{\varepsilon}\rightarrow 0 as n\rightarrow \infty , which contradicts to u_{n}\not\rightarrow 0 in E_{\varepsilon} .

    Now we set v_{n}(x) = u_{n}(x + z_{n}) . We known that \{v_{n}\} is bounded. Then there exists v\in W^{s, p}(\mathbb{R}^{N}) such that v_{n} \rightharpoonup v weakly in W^{s, p}(\mathbb{R}^{N}) . By (3.14), we see that v\neq 0 . Hence, there is a set \Omega \subset \mathbb{R}^{N} with |\Omega| > 0 such that v(x) > 0 in \Omega . Then from (3.13) and t_{n} \geq 1+ \sigma , we have

    \begin{eqnarray*} 0 < K^{2}_{\infty}(( 1+ \sigma)^{2q-p} -1)\int_{\mathbb{R}^{2N}} \frac{|v_{n}(y)|^{q}|v_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy \leq C\nu +C\nu^{2} + o_{n}(1). \end{eqnarray*}

    Taking limit in the above inequality and by Fatou's lemma, we get

    \begin{eqnarray*} 0 < K^{2}_{\infty}(( 1+ \sigma)^{2q-p} -1)\int_{\mathbb{R}^{2N}} \frac{|v(y)|^{q}|v(x)|^{q}}{|x-y|^{\mu}}dxdy \leq C\nu +C\nu^{2}, \end{eqnarray*}

    for any \nu > 0 . It's a contradiction. Therefore, \limsup_{n\rightarrow \infty} t_{n} \leq 1 .

    We next consider the following two cases:

    Case 1. \limsup_{n\rightarrow \infty} t_{n} = 1 . We assume that there exists a subsequence, still denoted by \{t_{n}\} such that \lim_{n\rightarrow \infty} t_{n} = 1 . Recalling t_{n}u_{n}\in \mathcal{M}_{V_{\infty}K_{\infty}} , then

    \begin{eqnarray} c +o_{n}(1) = I_{\varepsilon}(u_{n}) & = &I_{\varepsilon}(u_{n})-I_{V_{\infty}K_{\infty}}(t_{n}u_{n}) +I_{V_{\infty}K_{\infty}}(t_{n}u_{n}) \\ &\geq & c_{\infty} + I_{\varepsilon}(u_{n})-I_{V_{\infty}K_{\infty}}(t_{n}u_{n}). \end{eqnarray} (3.15)

    We observe that

    \begin{eqnarray} I_{\varepsilon}(u_{n})-I_{V_{\infty}K_{\infty}}(t_{n}u_{n})& = & \frac{1-t^{p}_{n}}{p}[u_{n}]^{p}_{s, p} +\frac{1}{p}\int_{\mathbb{R}^{N}}(V_{\varepsilon}(x)-t^{p}_{n}V_{\infty})|u_{n}|^{p}dx\\ &+& \frac{1}{2q}\int_{\mathbb{R}^{2N}}\frac{(K^{2}_{\infty}t^{2q}_{n}-K_{\varepsilon}( y)K_{\varepsilon}( x))|u_{n}(y)|^{q}|u_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy. \end{eqnarray} (3.16)

    From the boundedness of \{u_{n}\} , \lim_{n\rightarrow \infty} t_{n} = 1 , u_{n} \rightharpoonup 0 in E_{\varepsilon} and (3.5), one has

    \begin{eqnarray} \frac{1-t^{p}_{n}}{p}[u_{n}]^{p}_{s, p} = o_{n}(1) \end{eqnarray} (3.17)

    and

    \begin{eqnarray} \int_{\mathbb{R}^{N}}(V_{\varepsilon}(x)-t^{p}_{n}V_{\infty})|u_{n}|^{p}dx \geq o_{n}(1) -C\nu. \end{eqnarray} (3.18)

    By (3.12), we have

    \begin{eqnarray} &&\int_{\mathbb{R}^{2N}}\frac{(K^{2}_{\infty}t^{2q}_{n}-K_{\varepsilon}( y)K_{\varepsilon}( x))|u_{n}(y)|^{q}|u_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy \\ &&\geq (t^{2q}_{n}-1)K^{2}_{\infty}\int_{\mathbb{R}^{2N}}\frac{|u_{n}(y)|^{q}|u_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy-o_{n}(1)-C\nu -C\nu^{2}\\ && = o_{n}(1)-C\nu -C\nu^{2}. \end{eqnarray} (3.19)

    It follows from (3.15)–(3.19) that

    \begin{eqnarray*} c+o_{n}(1) \geq c_{\infty} + o_{n}(1)-C\nu -C\nu^{2}. \end{eqnarray*}

    Letting n\rightarrow \infty and \nu \rightarrow 0 , we get c \geq c_{\infty} .

    Case 2. \limsup_{n\rightarrow \infty} t_{n} = t_{0} < 1 . In this case, without loss of generality, we assume that t_{n} < 1 for all n . Recalling that t_{n}u_{n}\in \mathcal{M}_{V_{\infty}K_{\infty}} , then by I'_{\varepsilon}(u_{n}) \rightarrow 0 , (3.6) and the boundedness of \{u_{n}\} , we have

    \begin{eqnarray*} c_{\infty}&\leq &I_{V_{\infty}K_{\infty}}(t_{n}u_{n}) = I_{V_{\infty}K_{\infty}}(t_{n}u_{n})-\frac{1}{2q}\langle I'_{V_{\infty}K_{\infty}}(t_{n}u_{n}) t_{n}u_{n}\rangle\\ & = &(\frac{1}{p} - \frac{1}{2q})t^{p}_{n}\left( [u_{n}]^{p}_{s, p} + \int_{\mathbb{R}^{N}}V_{\infty}|u_{n}|^{p}dx \right)\\ & < &(\frac{1}{p} - \frac{1}{2q})\left( [u_{n}]^{p}_{s, p} + \int_{\mathbb{R}^{N}}V_{\infty}|u_{n}|^{p}dx \right)\\ & = & I_{\varepsilon}(u_{n}) - \frac{1}{2q}\langle I'_{\varepsilon}(u_{n}), u_{n}\rangle +(\frac{1}{p} - \frac{1}{2q})\int_{\mathbb{R}^{N}}(V_{\infty}-V_{\varepsilon}(x))|u_{n}|^{p}dx+ o_{n}(1)\\ &\leq & I_{\varepsilon}(u_{n})+ o_{n}(1) + C\nu. \end{eqnarray*}

    Letting n\rightarrow \infty and \nu \rightarrow 0 , we get c_{\infty} \leq c . The proof is completed.

    Lemma 3.2. The functional I_{\varepsilon} satisfies (PS)_{c} condition with c < c_{\infty} .

    Proof. Let \{u_{n}\}\subset E_{\varepsilon} be a sequence such that I_{\varepsilon}(u_{n}) \rightarrow c and I'_{\varepsilon}(u_{n}) \rightarrow 0 as n \rightarrow \infty . By (3.1) we have that \{u_{n}\} is bounded in E_{\varepsilon} . Then, up to a subsequence, there exists u\in E_{\varepsilon} such that

    \begin{eqnarray} \left\{ \begin{aligned} &u_{n}\rightharpoonup u, \hskip0.5cm\text{weakly} \hskip0.2cm \text{in}\hskip0.2cm E_{\varepsilon}, \\ & u_{n}\rightarrow u, \hskip0.5cm \text{strongly} \hskip0.2cm \text{in} \hskip0.2cm L^{r}_{loc}(\mathbb{R}^{N}), \hskip0.2cm p\leq r < p^{*}_{s}, \\ & u_{n}\rightarrow u, \hskip0.5cm \text{a.e.} \hskip 0.2cm \text{in} \hskip0.2cm \mathbb{R}^{N}. \end{aligned} \right. \end{eqnarray} (3.20)

    By (3.20), p < q < \frac{p^{*}_{s}}{2}(2-\frac{\mu}{N}) , and Hardy-Littlewood-Sobolev inequality, we obtain that

    \begin{eqnarray*} \int_{\mathbb{R}^{N}}\frac{K_{\varepsilon}(y)|u_{n}(y)|^{q}}{|x-y|^{\mu}}dy\rightharpoonup \int_{\mathbb{R}^{N}}\frac{K_{\varepsilon}(y)|u(y)|^{q}}{|x-y|^{\mu}}dy\hskip0.4cm \text{in} \hskip0.1cm L^{\frac{2N}{\mu}}(\mathbb{R}^{N}). \end{eqnarray*}

    Then, for any \phi \in C^{\infty}_{0}(\mathbb{R}^{N}) , we have

    \begin{eqnarray*} \int_{\mathbb{R}^{2N}} \frac{K_{\varepsilon}( y)|u_{n}(y)|^{q}K_{\varepsilon}( x)|u_{n}(x)|^{q-2}u_{n}(x)\phi}{|x-y|^{\mu}}dxdy = \int_{\mathbb{R}^{2N}} \frac{K_{\varepsilon}(y)|u(y)|^{q}K_{\varepsilon}( x)|u(x)|^{q-2}u(x)\phi}{|x-y|^{\mu}}dxdy+ o_{n}(1). \end{eqnarray*}

    Then, we have I'_{\varepsilon}(u) = 0. Set w_{n} = u_{n} -u . By Brezis-Lieb lemma, we have

    \begin{eqnarray} \|w_{n}\|^{p}_{\varepsilon} = \|u_{n}\|^{p}_{\varepsilon}-\|u\|^{p}_{\varepsilon}+o_{n}(1). \end{eqnarray} (3.21)

    From a Brezis-Lieb lemma for the nonlocal term of the functional ([17]), we obtain

    \begin{eqnarray} \int_{\mathbb{R}^{N}}(\frac{1}{|x|^{\mu}}*K_{\varepsilon}|w_{n}|^{q})K_{\varepsilon}(x)|w_{n}|^{q}dx & = &\int_{\mathbb{R}^{N}}(\frac{1}{|x|^{\mu}}*K_{\varepsilon}|u_{n}|^{q})K_{\varepsilon}(x)|u_{n}|^{q}dx\\ &&-\int_{\mathbb{R}^{N}}(\frac{1}{|x|^{\mu}}*K_{\varepsilon}|u|^{q})K_{\varepsilon}(x)|u|^{q}dx+o_{n}(1). \end{eqnarray} (3.22)

    It follows from (3.21) and (3.22) that

    \begin{eqnarray*} I_{\varepsilon}(w_{n}) = I_{\varepsilon}(u_{n})-I_{\varepsilon}(u) +o_{n}(1) = c-I_{\varepsilon}(u) +o_{n}(1) \end{eqnarray*}

    and I'_{\varepsilon}(w_{n}) \rightarrow 0 as n \rightarrow \infty . Since I'_{\varepsilon}(u) = 0 , we get

    \begin{eqnarray*} I_{\varepsilon}(u) = I_{\varepsilon}(u)-\frac{1}{2q}\langle I'_{\varepsilon}(u), u\rangle = (\frac{1}{p}-\frac{1}{2q})\|u\|^{p}_{\varepsilon}\geq 0. \end{eqnarray*}

    Hence, I_{\varepsilon}(w_{n}) \rightarrow c-I_{\varepsilon}(u) < c_{\infty} . By Lemma 3.1, w_{n} \rightarrow 0 in E_{\varepsilon} . Then u_{n}\rightarrow u in E_{\varepsilon} . The proof is completed.

    Lemma 3.3. Let \{u_{n}\} be a (PS)_{c} sequence restricted in \mathcal{N}_{\varepsilon} and assume that c < c_{\infty} . Then \{u_{n}\} has a convergent subsequence in E_{\varepsilon} .

    Proof. Let \{u_{n}\} be a (PS)_{c} sequence for I_{\varepsilon} on \mathcal{N}_{\varepsilon} at level c , namely

    \begin{eqnarray*} I_{\varepsilon}(u_{n}) \rightarrow c \hskip0.2cm \text{and}\hskip0.2cm I'_{\varepsilon}|_{\mathcal{N}_{\varepsilon}}(u_{n}) \rightarrow 0. \end{eqnarray*}

    It's easy to check that \{u_{n}\} is bounded in E_{\varepsilon} . We assume that

    \begin{eqnarray*} I'_{\varepsilon}(u_{n}) = o_{n}(1) + \lambda_{n}g'(u_{n}), \end{eqnarray*}

    where g(u) = \langle I'_{\varepsilon}(u), u\rangle , and

    \begin{eqnarray} \langle g'(u), u\rangle = p\|u\|^{p}_{\varepsilon}-2q\int_{\mathbb{R}^{2N}} \frac{K_{\varepsilon}(y)|u(y)|^{q}K_{\varepsilon}( x)|u(x)|^{q}}{|x-y|^{\mu}}dxdy. \end{eqnarray} (3.23)

    Since \{u_{n}\} is bounded, we have

    \begin{eqnarray} 0 = g(u_{n}) = \langle I'_{\varepsilon}(u_{n}), u_{n}\rangle = o_{n}(1) + \lambda_{n}\langle g'(u_{n}), u_{n}\rangle. \end{eqnarray} (3.24)

    Since u_{n}\in \mathcal{N}_{\varepsilon} , by (2.3) and (3.23) we have

    \begin{eqnarray*} \langle g'(u_{n}), u_{n}\rangle = (p-2q)\|u_{n}\|^{p}_{\varepsilon}\leq (p-2q)r^{*}, \end{eqnarray*}

    where r^{*} is defined in (2.3). Then,

    \begin{eqnarray*} |\lambda_{n}\langle g'(u_{n}), u_{n}\rangle|\geq |\lambda_{n}|(2q-p)r^{*}. \end{eqnarray*}

    Thus \lambda_{n} \rightarrow 0 and I'_{\varepsilon}(u_{n}) \rightarrow 0 as n\rightarrow \infty . Therefore, \{u_{n}\} is a (PS)_{c} sequence for I_{\varepsilon} in E_{\varepsilon} . By Lemma 3.2, \{u_{n}\} has a convergent subsequence.

    We only give the details proof under the assumptions ( \text{H}_{1} ) and ( \text{H}_{2} ). The arguments of ( \text{H}_{3} ) is similar. Under the assumption ( \text{H}_{2} ), we may suppose that x_{*} = 0\in\mathcal{V} or x_{*} = 0\in\mathcal{V}\cap \mathcal{K} if \mathcal{V}\cap \mathcal{K} \neq \emptyset . Then

    \begin{eqnarray*} \theta = V(0)\hskip0.5cm \text{and}\hskip0.5cm \alpha: = K(0)\geq K(x) \hskip0.2cm \text{for} \hskip0.1cm \text{ all} \hskip0.2cm |x|\geq R. \end{eqnarray*}

    Lemma 4.1. \limsup_{\varepsilon\rightarrow 0}c_{\varepsilon} \leq c_{\theta\alpha} . In particular, if \mathcal{V}\cap \mathcal{K} \neq \emptyset , then \limsup_{\varepsilon\rightarrow 0}c_{\varepsilon} = c_{\theta\kappa} .

    Proof. Let w\in \mathcal{M}_{\theta\alpha} be such that

    \begin{eqnarray*} c_{\theta\alpha} = I_{\theta\alpha}(w) = \max\limits_{t\geq 0}I_{\theta\alpha}(tw). \end{eqnarray*}

    Then there exists a unique t_{\varepsilon} > 0 such that t_{\varepsilon}w \in \mathcal{N}_{\varepsilon} . Thus

    \begin{eqnarray} c_{\varepsilon} \leq I_{\varepsilon}(t_{\varepsilon}w) = \max\limits_{t\geq 0}I_{\varepsilon}(tw). \end{eqnarray} (4.1)

    Observe that

    \begin{eqnarray} I_{\varepsilon}(t_{\varepsilon}w)& = &I_{\theta\alpha}(t_{\varepsilon}w) +\frac{t^{p}_{\varepsilon}}{p}\int_{\mathbb{R}^{N}}(V_{\varepsilon}(x)-\theta)|w|^{p}dx\\ &&+\frac{t^{2q}_{\varepsilon}}{2q}\int_{\mathbb{R}^{2N}} \frac{(\alpha^{2}-K_{\varepsilon}( y)K_{\varepsilon}( x))|w(y)|^{q}|w(x)|^{q}}{|x-y|^{\mu}}dxdy. \end{eqnarray} (4.2)

    Since t_{\varepsilon}w \in \mathcal{N}_{\varepsilon} , by the boundedness of K(x) , we get that there exist T_{2} > T_{1} > 0 such that T_{1}\leq t_{\varepsilon} < T_{2} . We may assume that t_{\varepsilon} \rightarrow t_{0} as \varepsilon \rightarrow 0 . Then by the boundedness of V , K , and the Lebesgue's theorem, we have

    \begin{eqnarray*} \frac{t^{p}_{\varepsilon}}{p}\int_{\mathbb{R}^{N}}(V_{\varepsilon}(x)-\theta)|w|^{p}dx = o_{\varepsilon}(1), \end{eqnarray*}

    and

    \begin{eqnarray*} \frac{t^{2q}_{\varepsilon}}{2q}\int_{\mathbb{R}^{2N}} \frac{(\alpha^{2}-K_{\varepsilon}(y)K_{\varepsilon}( x))|w(y)|^{q}|w(x)|^{q}}{|x-y|^{\mu}}dxdy = o_{\varepsilon}(1), \end{eqnarray*}

    Thus, by (4.2) we obtain

    \begin{eqnarray*} I_{\varepsilon}(tw) = I_{\theta\alpha}(t_{0}w) +o_{\varepsilon}(1). \end{eqnarray*}

    It follows from (4.1) that

    \begin{eqnarray*} c_{\varepsilon} &\leq& I_{\theta\alpha}(t_{0}w) +o_{\varepsilon}(1) \leq \max\limits_{t\geq 0}I_{\theta\alpha}(tw) = I_{\theta\alpha}(w) = c_{\theta\alpha}. \end{eqnarray*}

    The proof is completed.

    Proposition 1. Assume that ( \mathit{\text{H}}_{1} ) and ( \mathit{\text{H}}_{2} ) hold. Then for any \varepsilon > 0 small enough, problem (2.1) has a positive ground state solution.

    Proof. Let \{u_{n}\} denotes the (PS) sequence for I_{\varepsilon} given in (2.4). Recall that \theta < V_{\infty} and \alpha \geq K_{\infty} . It follows from Lemma 2.2 that c_{\theta\alpha} < c_{\infty} . Then, by Lemmas 3.2 and 4.1, we obtain I_{\varepsilon} satisfies the (PS)_{c_{\varepsilon}} condition for \varepsilon > 0 small enough. Hence, by mountain pass lemma we have problem (2.1) has a nontrivial ground state solution u_{\varepsilon} . We note that all the calculations above can be repeated by word by word, replacing I^{+}_{\varepsilon} with the functional

    \begin{eqnarray*} I^{+}_{\varepsilon}(u) = \frac{1}{p}\|u\|^{p}_{\varepsilon} - \frac{1}{2q} \int_{\mathbb{R}^{2N}}\frac{K_{\varepsilon}( y)|u^{+}(y)|^{q}K_{\varepsilon}( x)|u^{+}(x)|^{q}}{|x-y|^{\mu}}dxdy. \end{eqnarray*}

    Then we get a ground state solution u_{\varepsilon} of the equation

    \begin{eqnarray*} (-\Delta )^{s}_{p}u + V_{\varepsilon}( x)|u|^{p-2}u = (\int_{\mathbb{R}^{N}}\frac{K_{\varepsilon}(y)|u^{+}(y)|^{q}}{|x-y|^{\mu}}dy)K_{\varepsilon}(x)|u^{+}|^{q-2}u^{+}, \hskip0.1cm x\in\mathbb{R}^{N}. \end{eqnarray*}

    Taking u^{-}_{\varepsilon} as a test function in above equation, we have

    \begin{eqnarray} \int_{\mathbb{R}^{2N}}\frac{|u_{\varepsilon}(x)-u_{\varepsilon}(y)|^{p}}{|x-y|^{N+ps}}(u_{\varepsilon}(x)-u_{\varepsilon}(y)) (u^{-}_{\varepsilon}(x)-u^{-}_{\varepsilon}(y))dxdy +\int_{\mathbb{R}^{N}} V_{\varepsilon}(x)|u^{-}_{\varepsilon}(x)|^{p}dx = 0. \end{eqnarray} (4.3)

    For any p\geq 1 , we have

    \begin{eqnarray} |u_{\varepsilon}(x)-u_{\varepsilon}(y)|^{p-2} (u_{\varepsilon}(x)-u_{\varepsilon}(y)) (u^{-}_{\varepsilon}(x)-u^{-}_{\varepsilon}(y))\geq |u^{-}_{\varepsilon}(x)-u^{-}_{\varepsilon}(y)|^{p}. \end{eqnarray} (4.4)

    Combining (4.3) with (4.4) yields

    \begin{eqnarray*} \|u^{-}_{\varepsilon}\|^{p}_{\varepsilon} = [u^{-}_{\varepsilon}]^{p}_{s, p} +\int_{\mathbb{R}^{N}} V_{\varepsilon}(x)|u^{-}_{\varepsilon}(x)|^{p}dx \leq 0. \end{eqnarray*}

    Thus, we have u^{-}_{\varepsilon}(x)\equiv 0 and u_{\varepsilon}\geq 0 . It follows from the maximum principle ([22]) that u_{\varepsilon} > 0 in \mathbb{R}^{N} .

    Lemma 4.2. Let u_{\varepsilon_{n}} be a solution of (2.1) given in Proposition 1. Then, there exists a sequence \{z_{\varepsilon_{n}}\} \subset \mathbb{R}^{N} with \varepsilon_{n}z_{\varepsilon_{n}} \rightarrow z_{0}\in \Omega_{V} such that v_{\varepsilon_{n}} = u_{\varepsilon_{n}}(x + z_{\varepsilon_{n}}) converges strongly in W^{s, p}(\mathbb{R}^{N}) to a ground state solution of

    \begin{eqnarray*} (-\Delta )^{s}_{p}u + V(z_{0})|u|^{p-2}u = K^{2}(z_{0})(\int_{\mathbb{R}^{N}}\frac{|u(y)|^{q}}{|x-y|^{\mu}}dy)|u|^{q-2}u, \hskip0.1cm x\in\mathbb{R}^{N}. \end{eqnarray*}

    In particular, if \mathcal{V}\cap \mathcal{K} \neq \emptyset , then z_{0}\in \mathcal{V}\cap \mathcal{K} , and up to a subsequence, v_{\varepsilon_{n}} converges in W^{s, p}(\mathbb{R}^{N}) to a ground state solution of

    \begin{eqnarray*} (-\Delta )^{s}_{p}u + \theta|u|^{p-2}u = \kappa^{2}(\int_{\mathbb{R}^{N}}\frac{|u(y)|^{q}}{|x-y|^{\mu}}dy)|u|^{q-2}u, \hskip0.1cm x\in\mathbb{R}^{N}. \end{eqnarray*}

    Proof. Let \varepsilon_{n} \rightarrow 0 as n \rightarrow \infty , u_{n} : = u_{\varepsilon_{n}} \in \mathcal{N}_{\varepsilon_{n}} be a solution of (2.1) . Then I_{\varepsilon_{n}}(u_{n}) = c_{\varepsilon_{n} } and I'_{\varepsilon_{n}}(u_{n}) = 0 . It is easy to check that \{u_{n}\} is bounded. Then, there exist R^{*}, \beta > 0 and a sequence \{z_{n}\} \subset \mathbb{R}^{N} such that

    \begin{eqnarray} \liminf\limits_{n\rightarrow \infty}\int_{B_{R^{*}}(z_{n})}|u_{n}|^{p}dx \geq \beta > 0. \end{eqnarray} (4.5)

    Now we set v_{n} = u_{n}(x + z_{n}) . Then v_{n} is a solution of the following equation

    \begin{eqnarray} (-\Delta )^{s}_{p}u + V_{n}(x)|u|^{p-2}u = (\int_{\mathbb{R}^{N}}\frac{K_{n}(y)|u(y)|^{q}}{|x-y|^{\mu}}dy)K_{n}( x)|u|^{q-2}u, \hskip0.1cm x\in\mathbb{R}^{N}, \end{eqnarray} (4.6)

    with the energy

    \begin{eqnarray*} J_{\varepsilon_{n}}(v_{n}) & = & \frac{1}{p}([v_{n}]^{p}_{s, p} +\int_{\mathbb{R}^{N}}V_{n}(x)|v_{n}|^{p} dx)\\ &&-\frac{1}{2q}\int_{\mathbb{R}^{2N}}\frac{K_{n}(y)|v_{n}(y)|^{q}K_{n}(x)|v_{n}(x)|^{q}}{|x-y|^{\mu}}dxdy\\ & = & I_{\varepsilon_{n}}(u_{n}) = c_{\varepsilon_{n}}, \end{eqnarray*}

    where V_{n}(x) = V(\varepsilon_{n}x +\varepsilon_{n}z_{n}) and K_{n}(x) = K(\varepsilon_{n}x +\varepsilon_{n}z_{n}) . We see that \{v_{n}\} is bounded, then there exists v \in W^{s, p}(\mathbb{R}^{N}) satisfying, after passing to a subsequence if necessary

    \begin{eqnarray} \left\{ \begin{aligned} &v_{n}\rightharpoonup v, \hskip0.5cm\text{weakly} \hskip0.2cm \text{in}\hskip0.2cm W^{s, p}(\mathbb{R}^{N}), \\ & v_{n}\rightarrow v, \hskip0.5cm \text{strongly} \hskip0.2cm \text{in} \hskip0.2cm L^{r}_{loc}(\mathbb{R}^{N}), \hskip0.2cm p\leq r < p^{*}_{s}, \\ & v_{n}\rightarrow v, \hskip0.5cm \text{a.e.} \hskip 0.2cm \text{in} \hskip0.2cm \mathbb{R}^{N}. \end{aligned} \right. \end{eqnarray} (4.7)

    It follows from (4.5) that v\neq 0 .

    We next show that \{\varepsilon_{n}z_{n}\} is bounded. Assume by contradiction that |\varepsilon_{n}z_{n}|\rightarrow \infty as n \rightarrow \infty . By the boundedness of V and K , we may assume that

    \begin{eqnarray} V(\varepsilon_{n}z_{n}) \rightarrow V^{\infty} \hskip0.5cm\text{and}\hskip0.5cm K(\varepsilon_{n}z_{n}) \rightarrow K^{\infty}. \end{eqnarray} (4.8)

    By the definition of V_{\infty} and K_{\infty} , we have that

    \begin{eqnarray} V^{\infty} \geq V_{\infty} > \theta, \hskip0.5cm \alpha\geq K^{\infty}. \end{eqnarray} (4.9)

    Since V and K are uniformly continuous, by (4.8) one has

    \begin{eqnarray*} |V_{n}(x) - V^{\infty}|\leq |V_{n}(x) - V(\varepsilon_{n}z_{n})| + |V(\varepsilon_{n}z_{n}) - V^{\infty}| = o_{n}(1), \end{eqnarray*}

    and

    \begin{eqnarray*} |K_{n}(x) - K^{\infty}|\leq |K_{n}(x) - K(\varepsilon_{n}z_{n})| + |K(\varepsilon_{n}z_{n}) - K^{\infty}| = o_{n}(1), \end{eqnarray*}

    uniformly on bounded sets of \mathbb{R}^{N} . Then we have

    \begin{eqnarray} V_{n} \rightarrow V^{\infty} \hskip0.5cm\text{and}\hskip0.5cm K_{n} \rightarrow K^{\infty}, \end{eqnarray} (4.10)

    as n \rightarrow \infty uniformly on bounded sets of \mathbb{R}^{N} . From (4.7) and (4.10), for each \varphi \in C^{\infty}_{0}(\mathbb{R}^{N}) , we have

    \begin{eqnarray} \int_{\mathbb{R}^{N}}V_{n}(x)|v_{n}|^{p-2}v_{n}\varphi dx = \int_{\mathbb{R}^{N}}V^{\infty}|v|^{p-2}v\varphi dx +o_{n}(1), \end{eqnarray} (4.11)

    and

    \begin{eqnarray} &&\int_{\mathbb{R}^{2N}}\frac{|v_{n}(x)-v_{n}(y)|^{p-2}}{|x-y|^{N+ps}}(v_{n}(x)-v_{n}(y)) (\varphi(x)-\varphi(y))dxdy\\ && = \int_{\mathbb{R}^{2N}}\frac{|v(x)-v(y)|^{p-2}}{|x-y|^{N+ps}}(v(x)-v(y)) (\varphi(x)-\varphi(y))dxdy+o_{n}(1). \end{eqnarray} (4.12)

    Moreover, by (4.7), (4.10) and Hardy-Littlewood-Sobolev inequality, we infer that

    \begin{eqnarray*} \int_{\mathbb{R}^{N}}\frac{K_{n}(y)|v_{n}(y)|^{q}}{|x-y|^{\mu}} dy\rightharpoonup \int_{\mathbb{R}^{N}}\frac{K^{\infty}|v(y)|^{q}}{|x-y|^{\mu}} dy\hskip0.2cm \text{in}\hskip0.2cm L^{\frac{2N}{\mu}}(\mathbb{R}^{N}), \end{eqnarray*}

    and

    \begin{eqnarray*} K_{n}(x)|v_{n}|^{q}v_{n}\rightarrow K^{\infty}|v|^{q-2}v \hskip0.2cm \text{in}\hskip0.2cm L^{r}(\mathbb{R}^{N}), \hskip0.2cm r\in[1, \frac{p^{*}_{s}}{q-1}). \end{eqnarray*}

    Then, for each \varphi \in C^{\infty}_{0}(\mathbb{R}^{N})

    \begin{eqnarray} \int_{\mathbb{R}^{2N}}\frac{K_{n}(y)|v_{n}(y)|^{q}K_{n}(x)|v_{n}(x)|^{q-2}v_{n}\varphi}{|x-y|^{\mu}}dxdy = \int_{\mathbb{R}^{2N}}\frac{(K^{\infty})^{2}|v(y)|^{q}|v(x)|^{q-2}v\varphi}{|x-y|^{\mu}}dxdy+o_{n}(1). \end{eqnarray} (4.13)

    Since v_{n} satisfies (4.6), by (4.11)–(4.13), for any \varphi \in C^{\infty}_{0}(\mathbb{R}^{N}) , we have

    \begin{eqnarray*} 0& = &\lim\limits_{n\rightarrow \infty}(\int_{\mathbb{R}^{2N}}\frac{|v_{n}(x)-v_{n}(y)|^{p-2}}{|x-y|^{N+ps}}(v_{n}(x)-v_{n}(y)) (\varphi(x)-\varphi(y))dxdy\\ &&+\int_{\mathbb{R}^{N}}V_{n}(x)|v_{n}|^{p-2}v_{n}\varphi dx -\int_{\mathbb{R}^{2N}}\frac{K_{n}(y)|v_{n}(y)|^{q}K_{n}(x)|v_{n}(x)|^{q-2}v_{n}\varphi}{|x-y|^{\mu}}dxdy)\\ & = & \int_{\mathbb{R}^{2N}}\frac{|v(x)-v(y)|^{p-2}}{|x-y|^{N+ps}}(v(x)-v(y)) (\varphi(x)-\varphi(y))dxdy + \int_{\mathbb{R}^{N}}V^{\infty}|v|^{p-2}v\varphi dx\\ &&-\int_{\mathbb{R}^{2N}}\frac{(K^{\infty})^{2}|v(y)|^{q}|v(x)|^{q-2}v\varphi}{|x-y|^{\mu}}dxdy, \end{eqnarray*}

    which implies v is a solution of

    \begin{eqnarray*} (-\Delta )^{s}_{p}u + V^{\infty}|u|^{p-2}u = (K^{\infty})^{2}(\int_{\mathbb{R}^{N}}\frac{|u(y)|^{q}}{|x-y|^{\mu}}dy)|u|^{q-2}u, \hskip0.1cm x\in\mathbb{R}^{N}. \end{eqnarray*}

    By (4.9) and Lemma 2.2, we have

    \begin{eqnarray} I_{V^{\infty}K^{\infty}} (v)\geq c_{V^{\infty}K^{\infty}} > c_{\theta\alpha}. \end{eqnarray} (4.14)

    Using Fatou's lemma and Lemma 4.1, we get

    \begin{eqnarray} c_{V^{\infty}K^{\infty}}&\leq &I_{V^{\infty}K^{\infty}}(v) = I_{V^{\infty}K^{\infty}}(v) -\frac{1}{2q}\langle I'_{V^{\infty}K^{\infty}} (v), v\rangle\\ & = & (\frac{1}{p} - \frac{1}{2q})(\int_{\mathbb{R}^{2N}}\frac{|v(x)-v(y)|^{p}}{|x-y|^{N+ps}}dxdy + \int_{\mathbb{R}^{N}}V^{\infty}|v|^{p}dx)\\ &\leq & \liminf\limits_{n\rightarrow \infty}(\frac{1}{p} - \frac{1}{2q})(\int_{\mathbb{R}^{2N}}\frac{|v_{n}(x)-v_{n}(y)|^{p}}{|x-y|^{N+ps}}dxdy + \int_{\mathbb{R}^{N}}V_{n}(x)|v|^{p}dx)\\ & = & \liminf\limits_{n\rightarrow \infty}(J_{\varepsilon_{n}}(v_{n}) -\frac{1}{2q}\langle J'_{\varepsilon_{n}} (v_{n}), v_{n}\rangle)\\ & = & \liminf\limits_{n\rightarrow \infty} I_{\varepsilon_{n}}(u_{n}) = \liminf\limits_{n\rightarrow \infty} c_{\varepsilon_{n}}\\ &\leq & \limsup\limits_{n\rightarrow \infty} c_{\varepsilon_{n}} \leq c_{\theta\alpha}, \end{eqnarray} (4.15)

    which contradicts to (4.14). Hence, \{\varepsilon_{n}z_{n}\} is bounded.

    Passing to a subsequence, still denoted by \{\varepsilon_{n}z_{n}\} , we may assume that \varepsilon_{n}z_{n} \rightarrow z_{0} as n \rightarrow \infty . Then V_{n}(x)\rightarrow V(z_{0}) and K_{n}(x)\rightarrow K(z_{0}) as n \rightarrow \infty . Hence, v is a solution of

    \begin{eqnarray} (-\Delta )^{s}_{p}u + V(z_{0})|u|^{p-2}u = K^{2}(z_{0})(\int_{\mathbb{R}^{N}}\frac{|u(y)|^{q}}{|x-y|^{\mu}}dy)|u|^{q-2}u, \hskip0.1cm x\in\mathbb{R}^{N}. \end{eqnarray} (4.16)

    Next, we claim that z_{0}\in \Omega_{V} . Suppose by contradiction that z_{0}\not\in \Omega_{V} , then by ( \text{H}_{2} ) and Lemma 2.2, we have c_{V(z_{0})K(z_{0})} > c_{\theta\alpha} . It follows from Lemma 4.1 and the proof of (4.15) that

    \begin{eqnarray*} \limsup\limits_{n\rightarrow \infty}c_{\varepsilon_{n}} \leq c_{\theta\alpha} < c_{V(z_{0})K(z_{0})} \leq \liminf\limits_{n\rightarrow \infty}c_{\varepsilon_{n}}, \end{eqnarray*}

    which is absurd. Hence, z_{0}\in \Omega_{V} , and then \lim_{n\rightarrow \infty}dist(\varepsilon_{n}z_{n}, \Omega_{V}) = 0 . Moreover, v is a ground state solution of (4.16). In particular, if \mathcal{V}\cap \mathcal{K} \neq \emptyset , we see that \lim_{n\rightarrow \infty}dist(\varepsilon_{n}z_{n}, \mathcal{V}\cap \mathcal{K}) = 0 and v is a ground state solution of

    \begin{eqnarray} (-\Delta )^{s}_{p}u + \theta|u|^{p-2}u = \kappa^{2}(\int_{\mathbb{R}^{N}}\frac{|u(y)|^{q}}{|x-y|^{\mu}}dy)|u|^{q-2}u, \hskip0.1cm x\in\mathbb{R}^{N}. \end{eqnarray} (4.17)

    Finally, we shall prove that v_{n} \rightarrow v in W^{s, p}(\mathbb{R}^{N}) . Since v is a ground state solution of (4.16), by Fatou's lemma, we have

    \begin{eqnarray*} c_{V(z_{0})K(z_{0})}& = &I_{V(z_{0})K(z_{0})} (v) = I_{V(z_{0})K(z_{0})} (v) -\frac{1}{2q}\langle I'_{V(z_{0})K(z_{0})} (v), v\rangle\notag\\ & = & (\frac{1}{p} - \frac{1}{2q})(\int_{\mathbb{R}^{2N}}\frac{|v(x)-v(y)|^{p}}{|x-y|^{N+ps}}dxdy + \int_{\mathbb{R}^{N}}V(z_{0})|v|^{p}dx)\notag\\ &\leq & \liminf\limits_{n\rightarrow \infty}(\frac{1}{p} - \frac{1}{2q})(\int_{\mathbb{R}^{2N}}\frac{|v_{n}(x)-v_{n}(y)|^{p}}{|x-y|^{N+ps}}dxdy + \int_{\mathbb{R}^{N}}V_{n}(x)|v_{n}|^{p}dx)\notag\\ & = & \liminf\limits_{n\rightarrow \infty}(J_{\varepsilon_{n}}(v_{n}) -\frac{1}{2q}\langle J'_{\varepsilon_{n}} (v_{n}), v_{n}\rangle)\\ & = &\liminf\limits_{n\rightarrow \infty} c_{\varepsilon_{n}} \leq \limsup\limits_{n\rightarrow \infty} c_{\varepsilon_{n}} \leq c_{V(z_{0})K(z_{0})}. \end{eqnarray*}

    Then

    \begin{eqnarray*} \lim\limits_{n\rightarrow \infty}([v_{n}]^{p}_{s, p} + \int_{\mathbb{R}^{N}}V_{n}(x)|v_{n}|^{p}dx) = \int_{\mathbb{R}^{2N}}\frac{|v(x)-v(y)|^{p}}{|x-y|^{N+ps}}dxdy + \int_{\mathbb{R}^{N}}V(z_{0})|v|^{p}dx. \end{eqnarray*}

    Thus by V_{n}(x)\rightarrow V(z_{0}) and Brezis-Lieb lemma, we get v_{n} \rightarrow v in W^{s, p}(\mathbb{R}^{N}) .

    The proof is completed.

    Lemma 4.3. v_{n}\in L^{\infty}(\mathbb{R}^{N}) and there exists C > 0 such that |v_{n}|_{\infty} \leq C for all n . Furthermore, \lim_{|x|\rightarrow \infty} v_{n}(x) = 0 uniformly in n , where v_{n} is given in Lemma 4.2.

    Proof. Given T > 0 and \beta > 0 . For each n , we denote v_{T, n} = v_{n}-(v_{n}-T)^{+} and g(v_{n}) = v_{n}v^{p\beta}_{T, n}\in W^{s, p}(\mathbb{R}^{N}) . Taking g(v_{n}) as the test function in (4.6), we have

    \begin{eqnarray} &&\int_{\mathbb{R}^{2N}}\frac{K_{n}(y)|v_{n}(y)|^{q}K_{n}(x)|v_{n}(x)|^{q-2}v_{n}g(v_{n})}{|x-y|^{\mu}}dxdy\\ && = \int_{\mathbb{R}^{2N}}\frac{|v_{n}(x)-v_{n}(y)|^{p-2}}{|x-y|^{N+ps}}(v_{n}(x)-v_{n}(y)) (g(v_{n})(x)-g(v_{n})(y))dxdy\\ &&+ \int_{\mathbb{R}^{N}}V_{n}(x)|v_{n}|^{p-2}v_{n}g(v_{n}) dx. \end{eqnarray} (4.18)

    By the boundedness of K and \{v_{n}\} , we have for each n ,

    \begin{eqnarray*} \int_{\mathbb{R}^{N}}\frac{K_{n}(y)|v_{n}(y)|^{q}}{|x-y|^{\mu}}dxdy \leq C. \end{eqnarray*}

    Therefore,

    \begin{eqnarray} \int_{\mathbb{R}^{2N}}\frac{K_{n}(y)|v_{n}(y)|^{q}K_{n}(x)|v_{n}(x)|^{q-2}v_{n}g(v_{n})}{|x-y|^{\mu}}dxdy\leq C \int_{\mathbb{R}^{N}}|v_{n}|^{q}v^{p\beta}_{T, n}dx. \end{eqnarray} (4.19)

    Set

    \begin{eqnarray*} G(t) = \int^{t}_{0}(g'(\tau))^{\frac{1}{p}}d\tau. \end{eqnarray*}

    We have for any a, b \in \mathbb{R} ,

    \begin{eqnarray*} |G(a)-G(b)|^{p} \leq |a-b|^{p-2}(a-b)(g(a)-g(b)). \end{eqnarray*}

    It follows from (4.18) and (4.19) that

    \begin{eqnarray} \int_{\mathbb{R}^{2N}}\frac{|G(v_{n}(x))-G(v_{n}(y))|^{p}}{|x-y|^{N+ps}}dxdy \leq C \int_{\mathbb{R}^{N}}|v_{n}|^{q}v^{p\beta}_{T, n}dx. \end{eqnarray} (4.20)

    By the definition of g , we can get G(v_{n}) \geq \frac{1}{\beta +1}v_{n}v^{\beta}_{T, n} . From (4.20) and the Sobolev inequality we obtain

    \begin{eqnarray} \left( \int_{\mathbb{R}^{N}}|v_{n}v^{\beta}_{T, n}|^{p^{*}_{s}}dx \right)^{\frac{p}{p^{*}_{s}}} \leq C(\beta +1)^{p} \int_{\mathbb{R}^{N}}|v_{n}|^{q}v^{p\beta}_{T, n}dx. \end{eqnarray} (4.21)

    Choosing T_{0} > 1 , by the definition of v_{T, n} , we obtain

    \begin{eqnarray} \int_{\mathbb{R}^{N}}|v_{n}|^{q}v^{p\beta}_{T, n}dx & = &\int_{\{v_{n}\leq T_{0}\}}|v_{n}|^{q}v^{p\beta}_{T, n}dx +\int_{\{v_{n} > T_{0}\}}|v_{n}|^{q}v^{p\beta}_{T, n}dx\\ &\leq &T^{p\beta}_{0}\int_{\mathbb{R}^{N}}|v_{n}|^{q}dx +\int_{\{v_{n} > T_{0}\}}|v_{n}|^{p^{*}_{s}}v^{p\beta}_{T, n}dx\\ &\leq &\left(\int_{\{v_{n} > T_{0}\}}|v_{n}|^{p^{*}_{s}}dx\right)^{\frac{p^{*}_{s}-p}{p^{*}_{s}}} \left(\int_{\mathbb{R}^{N}}|v^{p}_{n}v^{p\beta}_{T, n}|^{\frac{p^{*}_{s}}{p}}dx\right)^{\frac{p}{p^{*}_{s}}}\\ && + T^{p\beta}_{0}\int_{\mathbb{R}^{N}}|v_{n}|^{q}dx. \end{eqnarray} (4.22)

    Since v_{n} \rightarrow v in W^{s, p}(\mathbb{R}^{N}) , for T_{0} large enough, we conclude that

    \begin{eqnarray} \left(\int_{\{v_{n} > T_{0}\}}|v_{n}|^{p^{*}_{s}}dx\right)^{\frac{p^{*}_{s} -p}{p^{*}_{s}}} \leq \frac{1}{2C(\beta +1)^{p}}. \end{eqnarray} (4.23)

    By (4.22), (4.23) and the boundedness of \{v_{n}\} , we get

    \begin{eqnarray*} \int_{\mathbb{R}^{N}}|v_{n}|^{q}v^{p\beta}_{T, n}dx \leq CT^{p\beta}_{0} \int_{\mathbb{R}^{N}}|v_{n}|^{q}dx \leq C. \end{eqnarray*}

    Letting T \rightarrow \infty , by Fatou's lemma, we conclude that v_{n} \in L^{q+p\beta}(\mathbb{R}^{N}) , and (4.21) implies

    \begin{eqnarray*} \left(\int_{\mathbb{R}^{N}}|v_{n}|^{(\beta+1)p^{*}_{s}}dx\right)^{\frac{p}{p^{*}_{s}}} \leq C\int_{\mathbb{R}^{N}}|v_{n}|^{q+p\beta}dx . \end{eqnarray*}

    Now, from an iterative procedure, in a finite number of steps, we can show that v_{n}\in L^{\infty}(\mathbb{R}^{N}) and there exists C > 0 such that \|v_{n}\|_{\infty} \leq C . Moreover, by v_{n} \rightarrow v in W^{s, p}(\mathbb{R}^{N}) , we infer that \lim_{|x|\rightarrow \infty} v_{n}(x) = 0 uniformly in n .

    The proof is completed.

    Proof of Theorem 1.1. From proposition 1, problem (2.1) has a positive ground state solution u_{\varepsilon} . Then the function w_{\varepsilon}(x) = u_{\varepsilon}(\frac{x}{\varepsilon}) is a positive ground state solution of (1.1). Now we show the concentration of the maximum points. Let u_{\varepsilon_{n}} be a solution of (2.1). By Lemma 4.2, we have that v_{\varepsilon_{n}} = u_{\varepsilon_{n}}(x + z_{\varepsilon_{n}}) is a solution of the problem (4.6). Furthermore, v_{\varepsilon_{n}} \rightarrow v in W^{s, p}(\mathbb{R}^{N}) and \varepsilon_{n}z_{\varepsilon_{n}} \rightarrow z_{0}\in \Omega_{V} . Furthermore, from Lemma 4.3 we have v_{n}\in L^{\infty}(\mathbb{R}^{N}) for all n . It follows from (4.5) that

    \begin{eqnarray*} \beta \leq \int_{B_{R^{*}}(0)}|v_{\varepsilon_{n}}|^{p}dx \leq |B_{R^{*}}(0)||v_{\varepsilon_{n}}|^{p}_{\infty}. \end{eqnarray*}

    This implies there exist \iota > 0 such that

    \begin{eqnarray} \|v_{\varepsilon_{n}}\|_{\infty} \geq \iota > 0. \end{eqnarray} (4.24)

    Set p_{n} be a global maximum of v_{n} , by Lemma 4.3 and (4.24), we have that p_{n} \in B_{R}(0) for some R > 0 . Hence, the global maximum of u_{\varepsilon_{n}} given by y_{\varepsilon_{n}} = p_{n} + z_{\varepsilon_{n}} . Since \{p_{n}\} is bounded and \varepsilon_{n}z_{\varepsilon_{n}} \rightarrow z_{0} , we have \varepsilon_{n}y_{\varepsilon_{n}} \rightarrow z_{0} , thus the continuity of V and K gives \lim_{n\rightarrow \infty} V(\varepsilon_{n}z_{\varepsilon_{n}}) = V(z_{0}) and \lim_{n\rightarrow \infty} K(\varepsilon_{n}z_{\varepsilon_{n}}) = K(z_{0}) .

    The proof is completed.

    Let \delta > 0 be fixed. Define a smooth cut-off function \eta: \mathbb{R}^{+}\rightarrow \mathbb{R}^{+} satisfying \eta(t) = 1 if 0 \leq t \leq \frac{\delta}{2} , \eta(t) = 0 if t \geq \delta . For any \xi\in \Lambda , define

    \begin{eqnarray*} W_{\varepsilon, \xi}(x) = \eta(|\varepsilon x-\xi|)\omega(\frac{\varepsilon x-\xi}{\varepsilon}), \end{eqnarray*}

    where \omega(x) is the positive ground state solution of (4.17). By definition, W_{\varepsilon, \xi} has compact support for any \xi\in \Lambda , and hence it belongs to E_{\varepsilon} . It's easy to check that there exists a unique t_{\varepsilon} > 0 such that t_{\varepsilon}W_{\varepsilon, \xi} \in \mathcal{N}_{\varepsilon} . Now, we define the map \phi_{\varepsilon}: N\rightarrow \mathcal{N}_{\varepsilon} by \phi_{\varepsilon}(\xi) = t_{\varepsilon}W_{\varepsilon, \xi} .

    Lemma 5.1. Uniformly in \xi\in \Lambda , we have

    \begin{eqnarray*} \lim\limits_{\varepsilon \rightarrow 0}I_{\varepsilon}(\phi_{\varepsilon}(\xi)) = c_{\theta\kappa}. \end{eqnarray*}

    Proof. Let \xi\in \Lambda . By the definition of \phi_{\varepsilon}(\xi) and a simple change of variable, one has

    \begin{eqnarray} &&\int_{\mathbb{R}^{2N}}\frac{|\eta(\varepsilon|x|)\omega(x)-\eta(\varepsilon|y|)\omega(y)|^{p}}{|x-y|^{N+ps}}dxdy + \int_{\mathbb{R}^{N}}V(\varepsilon x +\xi)|\eta(\varepsilon|x|)\omega|^{p}dx\\ && = t^{2q-p}_{\varepsilon}\int_{\mathbb{R}^{2N}}\frac{K(\varepsilon y +\xi)|\eta(\varepsilon|y|)\omega(y)|^{q}K(\varepsilon x +\xi)|\eta(\varepsilon|x|)\omega(x)|^{q}}{|x-y|^{\mu}}dxdy. \end{eqnarray} (5.1)

    By Lemma 2.2 of [2] and Lebesgue's theorem, we get

    \begin{eqnarray} \lim\limits_{\varepsilon \rightarrow 0}\int_{\mathbb{R}^{2N}}\frac{|\eta(\varepsilon|x|)\omega(x)-\eta(\varepsilon|y|)\omega(y)|^{p}}{|x-y|^{N+ps}}dxdy = \int_{\mathbb{R}^{2N}}\frac{|\omega(x)-\omega(y)|^{p}}{|x-y|^{N+ps}}dxdy, \end{eqnarray} (5.2)
    \begin{eqnarray} \lim\limits_{\varepsilon \rightarrow 0}\int_{\mathbb{R}^{N}}V(\varepsilon x +\xi)|\eta(\varepsilon|x|)\omega|^{p}dx = \theta \int_{\mathbb{R}^{N}}|\omega|^{p}dx, \end{eqnarray} (5.3)

    and

    \begin{eqnarray} \lim\limits_{\varepsilon \rightarrow 0}\int_{\mathbb{R}^{2N}}\frac{K(\varepsilon y +\xi)|\eta(\varepsilon|y|)\omega(y)|^{q}K(\varepsilon x +\xi)|\eta(\varepsilon|x|)\omega(x)|^{q}}{|x-y|^{\mu}}dxdy = \kappa^{2}\int_{\mathbb{R}^{2N}}\frac{|\omega(y)|^{q} |\omega(x)|^{q}}{|x-y|^{\mu}}dxdy. \end{eqnarray} (5.4)

    On the other hand, since \omega is a ground state solution of (4.17), we have

    \begin{eqnarray} \int_{\mathbb{R}^{2N}}\frac{|\omega(x)-\omega(y)|^{p}}{|x-y|^{N+ps}}dxdy +\theta \int_{\mathbb{R}^{N}}|\omega|^{p}dx = \kappa^{2}\int_{\mathbb{R}^{2N}}\frac{|\omega(y)|^{q} |\omega(x)|^{q}}{|x-y|^{\mu}}dxdy. \end{eqnarray} (5.5)

    Then, by (5.1)–(5.5), we deduce that \lim_{\varepsilon \rightarrow 0}t_{\varepsilon} = 1 . At this point, by the same change of variable as before and (5.2), (5.3), we have

    \begin{eqnarray*} I_{\varepsilon}(\phi_{\varepsilon}(\xi)) & = &(\frac{1}{p} - \frac{1}{2q})t^{p}_{\varepsilon}(\int_{\mathbb{R}^{2N}}\frac{|\eta(\varepsilon|x|)\omega(x)-\eta(\varepsilon|y|)\omega(y)|^{p}}{|x-y|^{N+ps}}dxdy\\ &&+ \int_{\mathbb{R}^{N}}V(\varepsilon x +\xi)|\eta(\varepsilon|x|)\omega|^{p}dx)\\ & = &(\frac{1}{p} - \frac{1}{2q})(\int_{\mathbb{R}^{2N}}\frac{|\omega(x)-\omega(y)|^{p}}{|x-y|^{N+ps}}dxdy + \theta\int_{\mathbb{R}^{N}}|\omega|^{p}dx)+ o_{\varepsilon}(1)\\ & = &c_{\theta\kappa}+o_{\varepsilon}(1). \end{eqnarray*}

    Moreover, the limit is uniformly in \xi .

    The proof is completed.

    Let R > 0 be such that \Lambda_{\delta} \subset B_{R}(0) . Define \chi: \mathbb{R}^{N}\rightarrow \mathbb{R}^{N} by \chi(x) = x for x\in B_{R}(0) and \chi(x) = \frac{Rx}{|x|} for x\in \mathbb{R}^{N}\backslash B_{R}(0) . Now we define \rho_{\varepsilon}: \mathcal{N}_{\varepsilon} \rightarrow \mathbb{R}^{N} by

    \begin{eqnarray*} \rho_{\varepsilon}(u) = \frac{\int_{\mathbb{R}^{N}}\chi(\varepsilon x)|u|^{p}dx}{\int_{\mathbb{R}^{N}}|u|^{p}dx}. \end{eqnarray*}

    By the definition of \chi and the Lebesgue's theorm, we have that

    \begin{eqnarray} \rho_{\varepsilon}(\phi_{\varepsilon}(\xi)) & = & \frac{\int_{\mathbb{R}^{N}}\chi(\varepsilon x)|\eta(|\varepsilon x-\xi|)\omega(\frac{\varepsilon x-\xi}{\varepsilon})|^{p}dx}{\int_{\mathbb{R}^{N}}|\eta(|\varepsilon x-\xi|)\omega(\frac{\varepsilon x-\xi}{\varepsilon})|^{p}dx}\\ & = &\xi +\frac{\int_{\mathbb{R}^{N}}(\chi(\varepsilon x + \xi)-\xi)|\eta(|\varepsilon x|)\omega(x)|^{p}dx}{\int_{\mathbb{R}^{N}}|\eta(|\varepsilon x|)\omega(x)|^{p}dx}\\ & = &\xi +o_{\varepsilon}(1), \end{eqnarray} (5.6)

    uniformly for \xi \in \Lambda_{\delta} .

    Let f(\varepsilon) be any positive function tending to 0 as \varepsilon \rightarrow 0 . Set

    \begin{eqnarray*} \overline{\mathcal{N}}_{\varepsilon} = \{u\in \mathcal{N}_{\varepsilon}: I_{\varepsilon}(u) \leq c_{\theta\kappa} +f(\varepsilon) \}. \end{eqnarray*}

    By Lemma 5.1, we see that \overline{\mathcal{N}}_{\varepsilon}\neq \emptyset for \varepsilon > 0 small enough.

    Lemma 5.2.

    \begin{eqnarray*} \lim\limits_{\varepsilon \rightarrow 0}\sup\limits_{u\in \overline{\mathcal{N}}_{\varepsilon} }\inf\limits_{\xi \in \Lambda_{\delta}} |\rho_{\varepsilon}(u) -\xi| = 0. \end{eqnarray*}

    Proof. Let \varepsilon_{n} \rightarrow 0 as n \rightarrow \infty , by the definition, there exists a sequence \{u_{n}\} \subset \overline{\mathcal{N}}_{\varepsilon_{n}} such that

    \begin{eqnarray*} \inf\limits_{\xi \in \Lambda_{\delta}} |\rho_{\varepsilon_{n}}(u_{n}) -\xi| = \sup\limits_{u\in \overline{\mathcal{N}}_{\varepsilon} }\inf\limits_{\xi \in\Lambda_{\delta}} |\rho_{\varepsilon}(u) -\xi| + o_{n}(1). \end{eqnarray*}

    So it suffices to find a sequence \{\xi_{n}\} \subset \Lambda_{\delta} satisfying

    \begin{eqnarray} |\rho_{\varepsilon_{n}}(u_{n}) -\xi_{n}| = o_{n}(1). \end{eqnarray} (5.7)

    Since \{u_{n}\} \subset \overline{\mathcal{N}}_{\varepsilon_{n}} \subset \mathcal{N}_{\varepsilon_{n}} , by Lemma 2.2 we have that

    \begin{eqnarray*} c_{\theta\kappa} \leq c_{\varepsilon_{n}} \leq I_{\varepsilon_{n}}(u_{n}) \leq c_{\theta\kappa} +f(\varepsilon_{n}), \end{eqnarray*}

    this implies I_{\varepsilon_{n}}(u_{n}) \rightarrow c_{\theta\kappa} as n \rightarrow \infty . By the proof of Lemma 4.2, we obtain that there exists a sequence \{ \overline{\xi_{n}}\} \subset \mathbb{R}^{N} such that \varepsilon_{n}\overline{\xi_{n}} \rightarrow \overline{z_{0}} \in \Lambda and v_{n}(x) = u_{n}(x+\overline{\xi_{n}}) converges strongly in W^{s, p}(\mathbb{R}^{N}) to v , a positive ground state of (4.17). Set \xi_{n} = \varepsilon_{n}\overline{\xi_{n}} . Then by Lebesgue's theorem,

    \begin{eqnarray*} \rho_{\varepsilon_{n}}(u_{n}) & = & \frac{\int_{\mathbb{R}^{N}}\chi(\varepsilon_{n} x +\xi_{n} )|v_{n}(x )|^{p}dx}{\int_{\mathbb{R}^{N}}|v_{n}(x )|^{p}dx}\notag\\ & = &\xi_{n} +\frac{\int_{\mathbb{R}^{N}}(\chi(\varepsilon_{n} x +\xi_{n} )-\xi_{n})|v_{n}(x)|^{p}dx}{\int_{\mathbb{R}^{N}}|v_{n}(x )|^{p}dx}\\ & = &\xi_{n} +o_{n}(1). \end{eqnarray*}

    Thus (5.7) holds.

    The proof is completed.

    Proof of Theorem 1.2. For a fixed \delta > 0 , by Lemma 5.1 there exists \varepsilon_{\delta} > 0 such that I_{\varepsilon}(\phi_{\varepsilon}(\xi))\leq c_{\theta\kappa} + f(\varepsilon) for any \varepsilon \in (0, \varepsilon_{\delta}) and \xi \in \Lambda . Using Lemma 5.2, we have dist(\rho_{\varepsilon}(u), \Lambda_{\delta}) < \frac{\delta}{2} for such \varepsilon and u\in \overline{\mathcal{N}}_{\varepsilon} . It follows that the map \rho_{\varepsilon}\circ \phi_{\varepsilon}: \Lambda\rightarrow \Lambda_{\delta} is well defined. Then, by (5.6) the map \rho_{\varepsilon}\circ \phi_{\varepsilon} is homotopic to the inclusion map: Id: \Lambda\rightarrow \Lambda_{\delta} . Applying homotopic and by the same arguments of [6], we obtain that cat_{\overline{\mathcal{N}_{\varepsilon}}}(\overline{\mathcal{N}_{\varepsilon}}) \geq cat_{\Lambda_{\delta}}(\Lambda) . By the definition of \overline{\mathcal{N}_{\varepsilon}} and choosing \varepsilon_{\delta} small, from Lemma 3.3, we have that I_{\varepsilon} satisfies the (PS) condition in \overline{\mathcal{N}_{\varepsilon}} . Therefore, standard Ljusternik-Schnirelmann theory implies that I_{\varepsilon} has at least cat_{\overline{\mathcal{N}_{\varepsilon}}}(\overline{\mathcal{N}_{\varepsilon}}) critical points on \mathcal{N}_{\varepsilon} . It is easy to check that I_{\varepsilon} has at least cat_{\Lambda_{\delta}}(\Lambda) critical points in E_{\varepsilon} . The concentration behavior of these solutions as \varepsilon \rightarrow 0 are similar as in the proof of Theorem 1.1.

    The authors would like to express sincere thanks to the anonymous referees for their carefully reading this paper and valuable useful comments. This work was supported by NSFC (No.11601234, 11571176), Natural Science Foundation of Jiangsu Province of China for Young Scholar (No.BK20160571).

    The author declares no conflicts of interest in this paper.



    [1] V. Ambrosio, Multiplicity and concentration resuts for a fractional Choquard equation via penalization method, Potential Anal., 50 (2019), 55–82. doi: 10.1007/s11118-017-9673-3
    [2] V. Ambrosio, On the multiplicity and concentration of positive solutions for a p-fractional Choquard equation in \mathbb{R}^{N}, Comput. Math. Appl., 78 (2019), 2593–2617. doi: 10.1016/j.camwa.2019.04.001
    [3] C. O. Alves, M. B. Yang, Existence of semiclassical ground state solutions for a generalized Choquard equation, J. Differ. Equations., 257 (2014), 4133–4164. doi: 10.1016/j.jde.2014.08.004
    [4] S. Bhattarai, On fractional Schrödinger systems of Choquard type, J. Differ. Equations., 263 (2017), 3197–3229. doi: 10.1016/j.jde.2017.04.034
    [5] Y. H. Chen, C. G. Liu, Ground state solutions for non-autonomous fractional Choquard equations, Nonlinearity, 29 (2016), 1827–1842. doi: 10.1088/0951-7715/29/6/1827
    [6] S. Cingolani, M. Lazzo, Multiple positive solutions to nonlinear Schrödinger equations with competing potential functions, J. Differ. Equations, 160 (2000), 118–138. doi: 10.1006/jdeq.1999.3662
    [7] S. X. Chen, Y. Li, Z. P. Yang, Multiplicity and concentration of nontrivial nonnegative solutions for a fractional Choquard equation with critical exponent, RACSAM, 114 (2020), 33. doi: 10.1007/s13398-019-00768-4
    [8] P. Belchior, H. Bueno, O. H. Miyagaki, G. A. Pereira, Remarks about a fractional Choquard equation: Ground state, regularity and polynomial decay, Nonlinear Anal. 164 (2017), 38–53.
    [9] Y. H. Ding, X. Y. Liu, Semiclassical solutions of Schrödinger equations with magnetic fields and critical nonlinearities, Manuscripta Math., 140 (2013), 51–82. doi: 10.1007/s00229-011-0530-1
    [10] E. Di Nezza, G. Palatucci, E. Valdinoci, Hitchhiker's guide to the fractional Sobolev spaces, Bull. Sci. Math., 136 (2012), 521–573. doi: 10.1016/j.bulsci.2011.12.004
    [11] Z. Gao, X. H. Tang, S. T. Chen, On existence and concentration behavior of positive ground state solutions for a class of fractional Schrödinger-Choquard equations, Z. Angew. Math. Phys., 69 (2018), 122. doi: 10.1007/s00033-018-1016-8
    [12] K. R. W. Jones, Gravitational self-energy as the litmus of reality, Mod. Phys. Lett. A., 10 (1995), 657–667. doi: 10.1142/S0217732395000703
    [13] E. H. Lieb, Existence and uniqueness of the minimizing solution of Choquard's nonlinear equation, Stud. Appl. Math., 57 (1977), 93–105. doi: 10.1002/sapm197757293
    [14] P. L. Lions, The Choquard equation and related questions, Nonlinear Anal.-Theor., 4 (1980), 1063–1072. doi: 10.1016/0362-546X(80)90016-4
    [15] E. H. Lieb, M. Loss, Analysis, Graduate studies in mathematics, American Mathematical Society, 2001.
    [16] L. Ma, L. Zhao, Classification of positive solitary solutions of the nonlinear Choquard equation, Arch. Rational. Mech. Anal., 195 (2010), 455–467. doi: 10.1007/s00205-008-0208-3
    [17] V. Moroz, J. V. Schaftingen, Groundstates of nonlinear Choquard equations: Existence, qualitative properties and decay asymptotics, J. Funct. Anal., 265 (2013), 153–184. doi: 10.1016/j.jfa.2013.04.007
    [18] V. Moroz, J. V. Schaftingen, Semi-classical states for the Choquard equation, Calc. Var., 52 (2015), 199–235. doi: 10.1007/s00526-014-0709-x
    [19] V. Moroz, J. V. Schaftingen, A guide to the Choquard equation, J. Fixed Point Theory Appl., 19 (2017), 773–813. doi: 10.1007/s11784-016-0373-1
    [20] S. Pekar, Untersuchungen Über die Elektronentheorie der kristalle, Berlin: Akademie-Verlag, 1954.
    [21] P. d'Avenia, G. Siciliano, M. Squassina, On fractional Choquard equations, Math. Mod. Meth. Appl. S., 25 (2015), 1447–1476. doi: 10.1142/S0218202515500384
    [22] L. M. Del Pezzo, A. Quaas, A Hopf's lemma and a strong minimum principle for the fractional p-Laplacian, J. Differ. Equations., 263 (2017), 765–778. doi: 10.1016/j.jde.2017.02.051
    [23] Z. Shen, F. S. Gao, M. B. Yang, Ground states for nonlinear fractional Choquard equations with general nonlinearities, Math. Methods Appl. Sci., 39 (2016), 4082–4098. doi: 10.1002/mma.3849
    [24] G. Singh, Nonlocal perturbations of fractional Choquard equation, Adv. Nonlinear Anal., 8 (2019), 1184–1212.
    [25] L. Silvestre, Regularity of the obstacle problem for a fractional power of the Laplace operator, Commun. Pure Appl. Math., 60 (2007), 67–112. doi: 10.1002/cpa.20153
    [26] M. Willem, Minimax theorems, Basel, Berlin: Birkhaüser Boston, 1996.
    [27] J. Wang, L. X. Tian, J. X. Xu, F. B. Zhang, Existence and concentration of positive solutions for semilinear Schrödinger-Poisson systems in \mathbb{R}^{3}, Calc. Var., 48 (2013), 243–273. doi: 10.1007/s00526-012-0548-6
    [28] J. C. Wei, M. Winter, Strongly interacting bumps for the Schröinger-Newton equations, J. Math. Phys., 50 (2009), 012905. doi: 10.1063/1.3060169
    [29] J. K. Xia, Z. Q. Wang, Saddle solutions for the Choquard equation, Calc. Var., 58 (2019), 85. doi: 10.1007/s00526-019-1546-8
    [30] Y. Y. Yu, F. K. Zhao, L. G. Zhao, The concentration behavior of ground state solutions for a fractional Schrödinger-Poisson systems, Calc. Var., 56 (2017), 116. doi: 10.1007/s00526-017-1199-4
    [31] W. Zhang, X. Wu, Nodal solutions for a fractional Choquard equation, J. Math. Anal. Appl., 464 (2018), 1167–1183. doi: 10.1016/j.jmaa.2018.04.048
    [32] H. Zhang, J. Wang, F. B. Zhang, Semiclassical states for fractional Choquard equations with critical growth, Commun. Pure Appl. Anal., 18 (2019), 519–538. doi: 10.3934/cpaa.2019026
  • This article has been cited by:

    1. Die Hu, Xianhua Tang, Jiuyang Wei, Existence of Semiclassical Ground State Solutions for a Class of N-Laplace Choquard Equation with Critical Exponential Growth, 2024, 34, 1050-6926, 10.1007/s12220-024-01780-w
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2381) PDF downloads(86) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog