Skip to content
BY 4.0 license Open Access Published by De Gruyter December 1, 2020

Multiplicity and concentration behaviour of solutions for a fractional Choquard equation with critical growth

  • Zhipeng Yang and Fukun Zhao EMAIL logo

Abstract

In this paper, we study the singularly perturbed fractional Choquard equation

ε2s(Δ)su+V(x)u=εμ3(R3|u(y)|2μ,s+F(u(y))|xy|μdy)(|u|2μ,s2u+12μ,sf(u))inR3,

where ε > 0 is a small parameter, (−)s denotes the fractional Laplacian of order s ∈ (0, 1), 0 < μ < 3, 2μ,s=6μ32s is the critical exponent in the sense of Hardy-Littlewood-Sobolev inequality and fractional Laplace operator. F is the primitive of f which is a continuous subcritical term. Under a local condition imposed on the potential V, we investigate the relation between the number of positive solutions and the topology of the set where the potential attains its minimum values. In the proofs we apply variational methods, penalization techniques and Ljusternik-Schnirelmann theory.

MSC 2010: 35Q40; 35J50; 58E05

1 Introduction and the main results

In the present paper we are interested in the existence, multiplicity and concentration behavior of the semi-classical solutions of the singularly perturbed nonlocal elliptic equation

(1.1) ε2s(Δ)su+V(x)u=εμN(R3G(u(y))|xy|μdy)g(u)inRN,

where ε > 0 is a small parameter, 0 < μ < N, V, g = G' are real continuous functions on ℝN and the fractional Laplacian (−)s is defined by

(Δ)sΨ(x)=CN,sP.V.RNΨ(x)Ψ(y)|xy|N+2sdy,ΨS(RN),

P.V. stands for the Cauchy principal value, CN,s is a normalized constant, S(ℝN) is the Schwartz space of rapidly decaying functions, s ∈ (0, 1). As ε goes to zero in (1.1), the existence and asymptotic behavior of the solutions of the singularly perturbed equation (1.1) is known as the semi-classical problem. It was used to describe the transition between Quantum Mechanics and Classical Mechanics.

Our motivation to study (1.1) mainly comes from the fact that solutions u(x) of (1.1) correspond to standing wave solutions Ψ(x,t)=eiEt/εu(x) of the following time-dependent fractional Schrödinger equation

(1.2) iεΨt=ε2s(Δ)sΨ+(V(x)+E)Ψ(K(x)|G(Ψ)|)g(Ψ)(x,t)N×

where i is the imaginary unit, ε is related to the Planck constant. Equations of the type (1.2) was introduced by Laskin (see [25, 26]) and come from an expansion of the Feynman path integral from Brownian-like to Lévy-like quantum mechanical paths. It also appeared in several areas such as optimization, finance, phase transitions, stratified materials, crystal dislocation, flame propagation, conservation laws, materials science and water waves (see [11]).

When s = 1, the equation (1.1) turns out to be the Choquard equation

(1.3) ε2Δu+V(x)u=εμN(RNG(u(y))|xy|μdy)g(u)inRN,

The existence, multiplicity and concentration of solutions for (1.3) has been widely investigated. On one hand, some people have studied the classical problem, namely ε = 1 in (1.3). When V = 1 and G(u)=|u|qq, (1.3) covers in particular the Choquard-Pekar equation

(1.4) Δu+u=(RN1|χ|μ|u|qdy)|u|q2u in N.

The case N = 3, q = 2 and μ = 1 came from Pekar [38] in 1954 to describe the quantum mechanics of a polaron at rest. In 1976 Choquard used (1.4) to describe an electron trapped in its own hole, in a certain approximation to Hartree-Fock theory of one component plasma [27]. In this context (1.4) is also known as the nonlinear Schrödinger-Newton equation. By using critical point theory, Lions [29] obtained the existence of infinitely many radialy symmetric solutions in H1(ℝN) and Ackermann [1] prove the existence of infinitely many geometrically distinct weak solutions for a general case. For the properties of the ground state solutions, Ma and Zhao [30] proved that every positive solution is radially symmetric and monotone decreasing about some point for the generalized Choquard equation (1.4) with q ≥ 2. Later, Moroz and Van Schaftingen [32, 33] eliminated this restriction and showed the regularity, positivity and radial symmetry of the ground states for the optimal range of parameters, and also derived that these solutions decay asymptotically at infinity.

On the other hand, some people have focused on the semiclassical problem, namely, ε → 0 in (1.3). The question of the existence of semiclassical solutions for the non-local problem (1.3) has been posed in [5]. Note that if v is a solution of (1.3) for x0 N, then u = v(εx + x0) verifies

(1.5) Δu+V(εx+x0)u=(RNG(u(y))|xy|μdy)g(u)inRN,

which means some convergence of the family of solutions to a solution u0 of the limit problem

(1.6) Δu+V(x0)u=(RNG(u(y))|xy|μdy)g(u)inRN.

For this case when N = 3, μ = 1 and G(u) = |u|2, Wei and Winter [49] constructed families of solutions by a Lyapunov-Schmidt-type reduction when infxNV>0. This method of construction depends on the existence, uniqueness and non-degeneracy up to translations of the positive solution of the limiting equation (1.6), which is a difficult problem that has only been fully solved in the case when N = 3, μ = 1 and G(u) = |u|2. Moroz and Van Schaftingen [34] used variational methods to develop a novel non-local penalization technique to show that equation (1.3) with G(u) = |u|q has a family of solutions concentrated at the local minimum of V, with V satisfying some additional assumptions at infinity. In addition, Alves and Yang [4] investigated the multiplicity and concentration behaviour of solutions for a quasi-linear Choquard equation via the penalization method. Very recently, in an interesting paper, Alves et al. [2] study (1.3) with a critical growth, they consider the critical problem with both linear potential and nonlinear potential, and showed the existence, multiplicity and concentration behavior of solutions when the linear potential has a global minimum or maximum.

On the contrary, the results about fractional Choquard equation (1.1) are relatively few. Recently, d’Avenia, Siciliano and Squassina [17] studied the existence, regularity and asymptotic of the solutions for the following fractional Choquard equation

(1.7) (Δ)su+ωu=(RN|u(y)|q|xy|μdy)|u|q2uinRN,

where ω>0,2NμN<q<2NμN2s. Shen, Gao and Yang [42] obtain the existence of ground states for (1.7) with general nonlinearities by using variational methods. Chen and Liu [14] studied (1.7) with nonconstant linear potential and proved the existence of ground states without any symmetry property. For critical problem, Wang and Xiang [47] obtain the existence of infinitely many nontrivial solutions and the Brezis-Nirenberg type results can be founded in [36]. For the critical Choquard equations in the sense of Hardy-Littlewood-Sobolev, Cassani and Zhang [12] developed a robust method to get the existence of ground states and qualitative properties of solutions, where they do not require the nonlinearity to enjoy monotonicity nor Ambrosetti-Rabinowitz-type conditions. For other existence results we refer to [6, 8, 23, 24, 31, 48, 52] and the references therein.

It seems that the only works concerning the concentration behavior of solutions are due to [13, 51]. Assuming the global condition on V:

0<infxRNV(x)<liminf|x|V(x)=V,

which was firstly introduced by Rabinowitz [39] in the study of the nonlinear Schrödinger equations. By using the method of Nehari manifold developed by Szulkin and Weth [46], authors in [13, 51] obtained the multiplicity and concentration of positive solutions for the following fractional Choquard equation

(1.8) ε2s(Δ)su+V(x)u=εμ3(R3|u(y)|2μ,s+F(u(y))|xy|μdy)(|u|2μ,s2u+12μ,sf(u))inR3,

where ε > 0, 0 < μ < 3, F is the primitive function of f.

Different to [13, 51], in this paper, we are devote to establishing the existence and concentration of positive solutions for the fractional Choquard equation (1.8) when the potential function satisfies the following local conditions [18]:

V1VCR3,R and 0<infxR3V(x).

(V2)There is a bounded open domain Ω such that

V0:=infΩV(x)<minΩV(x).

Without loss of generality, we may assume that M={xΩ:V(x)=V0}andV(0)=minxR3V(x)=V0.

To go on studying the problem (1.8), the following Hardy-Littlewood-Sobolev inequality [28] is the starting point.

Lemma 1.1

Let t, r > 1 and 0 < μ < 3 with

1t+μ3+1r=2,

f ∈ Lt(ℝ3) and h ∈ Lr(ℝ3). There exists a sharp constant C(t, μ, r), independent of f , h such that

(1.9) R3R3f(x)h(y)|xy|μdydxC(t,μ,r)|f|t|h|r.

In particular, if t=r=66μ, then

C(t,μ,r)=C(μ)=πμ2Γ(32μ2)Γ(3μ2)(Γ()32Γ(3))μ31.

In this case there is equality in (1.9) if and only if f ≡ Ch and

h(x)=A(a2+|xb|2)6μ2

for some A ∈, a ∈\{0} and b ∈3.

Notice that, by the Hardy-Littlewood-Sobolev inequality, the integral

R3R3|u(x)|q|u(y)|q|xy|μdydx

ie well defined if uq ∈ Lt(ℝ3) satisfies 2t+μ3=2. Therefore, for u ∈ Hs(ℝ3) we will require that tq[2,2s], where 2s=632s is fractional critical Sobolev exponent for dimension 3. Then we have

6μ3q6μ32s.

Thus, 6μ3 is called the lower critical exponent and 2μ,s:=6μ32s is the upper critical exponent in the sense of Hardy-Littlewood-Sobolev inequality and the fractional Laplace operator.

For the nonlinearity term, we assume that the continuous function f vanishes in (−∞, 0) and satisfies:

(f1) |f(u)|c(|u|q11+|u|q21) for some c > 0 and 6μ3<q1q2<2μ,s.

(f2) The function u ⟼ f (u) is increasing in (0,∞).

(f3) (i)

lim|u|+F(u)|u|2μ,s1=+fors(34,1);

(ii)

lim|u|+F(u)|u|2μ,s2s32s(log|u|)12=+fors=34;

(iii)

lim|u|+F(u)|u|2μ,s2s32s=+fors(0,34).

Note that there is no (AR) type assumption on f. Then it is difficult to show that the functional satisfies the (PS) condition even for the autonomous case, which is necessary to use Ljusternik-Schnirelmann category theory. We shall investigate the (PS) sequence carefully and restore the compactness for (PS) sequence via some compactness Lemmas.

In order to describe the multiplicity, we first recall that, if Y is a closed subset of a topological space X, the Ljusternik-Schnirelmann category catXY is the least number of closed and contractible sets in X which cover Y. Then we state our main result as follows.

Theorem 1.1

If 0 < μ < 2s, assume that V satisfies (V1) and (V2) and the function f satisfies (f1)−(f3). Then for any δ > 0 such that

Mδ={xR3:dist(x,M)δ}Ω,

there exists εδ > 0 such that the problem (1.8) has at least catMδ(M) positive solutions for any ε ∈ (0, εδ). Moreover, if uε denotes one of these positive solutions and ηε3 its global maximum, then

limε0Vηε=V0.

Remark 1.1

Here, we make a few observations about the restriction on the parameter 0 < μ < 2s. In order to adapt the penalization method introduced by del Pino and Felmer in [18], we will proposed some control conditions on the non-local term (1|x|μ(|u|2μ,s+F(u))), which need some regularity (see Lemma 2.7 and 2.8), where we introduce the assumption 0 < μ < 2s.

We shall use the method of Nehari manifold, concentration compactness principle and category theory to prove the main results. There are some difficulties in proving our theorems. The first difficulty is that the non-linearity f is only continuous, we can not use standard arguments on the Nehari manifold. To overcome the nondifferentiability of the Nehari manifold, we shall use some variants of critical point theorems from Szulkin and Weth [46]. The second one is the lack of compactness of the embedding of Hs(ℝ3) into the space L2s*(3). We shall borrow the idea in [2, 12] to deal with the difficulties brought by the critical exponent. However, we require some new estimates, which are complicated because of the appearance of fractional Laplacian and the convolution-type nonlinearity. Moreover, the potential V satisfies (V1) and (V2) instead of the global condition. Since we have no information on the potential V at infinity, we adapt the truncation trick explored in [18]. It consists in making a suitable modification on the nonlinearity, solving a modified problem and then check that, for ε small enough, the solutions of the modified problem are indeed solutions of the original one. It is worthwhile to remark that in the arguments developed in [18], one of the key points is the existence of estimates involving the L-norm of the modified problem. But for the critical nonlocal problem (1.8), this kind of estimates are more delicate.

This paper is organized as follows. In section 2, besides describing the functional setting to study problem (1.8), we give some preliminary Lemmas which will be used later. In section 3, influenced by the work [18] and [45], we introduce a modified functional and show it satisfies the Palais-Smale condition. In section 4, we study the autonomous problem associated. This study allows us to show that the modified problem has multiple solutions. Finally, we show the critical point of the modified functional which satisfies the original problem, and investigate its concentration behavior, which completes the proof Theorem 1.1.

2 Variational settings and preliminary results

Throughout this paper, we denote | · |r the usual norm of the space Lr(ℝ3), 1 ≤ r < 1, Br(x) denotes the open ball with center at x and radius r, C or Ci(i = 1, 2, · · · ) denote some positive constants may change from line to line. *and mean the weak and strong convergence. Let E be a Hilbert space, the Fréchet derivative of a functional Φ at u, Φ'(u), is an element of the dual space E* and we shall denote Φ'(u) evaluated at v ∈ E by 〈Φ'(u), v〉.

2.1 The functional space setting

Firstly, fractional Sobolev spaces are the convenient setting for our problem, so we will give some skrtchs of the fractional order Sobolev spaces and the complete introduction can be found in [19]. We recall that, for any s ∈ (0, 1), the fractional Sobolev space H s(ℝ3) = Ws,2(ℝ3) is defined as follows:

Hs(R3)={uL2(R3):R3(|ξ|2s|F(u)|2+|F(u)|2)dξ<},

whose norm is defined as

uHs(R3)2=R3(|ξ|2s|F(u)|2+|F(u)|2)dξ,

where F denotes the Fourier transform. We also define the homogeneous fractional Sobolev space Ds,2(ℝ3) as the completion of C0R3 with respect to the norm

uDs,2(R3):=(R3×R3|u(x)u(y)|2|xy|3+2sdxdy)12=[u]Hs(R3).

The embedding Ds,2(3)L2s*(3) is continuous and for any s ∈ (0, 1), there exists a best constant Ss > 0 such that

SS:=infuDs,2(3)uDs,2(3)2|u|2s2

According to [16], Ss is attained by

(2.1) u0(x)=C(bb2+|xa|2)32s2,x3,

where C ∈ ℝ, b > 0 and a ∈3 are fixed parameters. We use SH,L to denote the best constant defined by

(2.2) SH,L:=infuDs,2(R3){0}R3|(Δ)s2u|2dx(R3R3|u(y)|2μ,s|u(x)|2μ,s|xy|μdydx)12μ,s.

The fractional Laplacian, (−Δ)su, of a smooth function u : ℝ3 ℝ, is defined by

F((Δ)su)(ξ)=|ξ|2sF(u)(ξ),ξ3.

Also (−Δ)su can be equivalently represented [19] as

(Δ)su(x)=12C(s)R3u(x+y)+u(xy)2u(x)|y|3+2sdy,xR3

where

C(s)=R31cosξ1|ξ|3+2sdξ1,ξ=ξ1,ξ2,ξ3.

Also, by the Plancherel formular in Fourier analysis, we have

[u]Hs(3)2=2C(s)|(Δ)s2u|22.

For convenience, we will omit the normalization constant in the following. As a consequence, the norms on Hs(ℝ3) defined below

uR3|u|2dx+R3×R3|u(x)u(y)|2|xy|3+2sdxdy12;uR3|ξ|2s|F(u)|2+|F(u)|2dξ12;uR3|u|2dx+(Δ)s2u2212.

are equivalent.

Making the change of variable x ⟼ εx, we can rewrite the equation (1.8) as the following equivalent form

(2.3) (Δ)su+V(εx)u=(R3|u(y)|2μ,s+F(u(y))|xy|μdy)(|u|2μ,s2u+12μ,sf(u))inR3,

If u is a solution of the equation (2.3), then v(x):=u(xε) is a solution of the equation (1.8). Thus, to study the equation (1.8), it suffices to study the equation (2.3). In view of the presence of potential V(x), we introduce the subspace

Hε={uHS(3):R3V(εx)u2dx<+},

which is a Hilbert space equipped with the inner product

(u,v)Hε=R3(Δ)s2u(Δ)s2vdx+R3V(εx)uvdx,

and the norm

uHε2=R3(Δ)s2u2dx+R3V(εx)u2dx.

We denote · by · ε in the sequel for convenience.

For the reader’s convenience, we review some useful result for this class of fractional Sobolev spaces:

Lemma 2.1

[19] Let 0 < s < 1, then there exists a constant C = C(s) > 0, such that

|u|2s2C[u]Hs(3)2

for every u ∈ Hs(ℝ3). Moreover, the embedding Hs(ℝ3) ↪ Lr(ℝ3) is continuous for any r[2,2S] and is locally compact whenever r[2,2S).

Lemma 2.2

[40] If {un} is bounded in Hs(ℝ3) and for some R > 0 we have

limnsupyR3BR(y)un2dx=0,

then un 0 in Lr(ℝ3) for any 2<r<2s.

Lemma 2.3

[37] Let uDs,2(3),φC0(3) and for each r>0,φr(x)=φ(xr). Then

uφr0 in Ds,2(3) as r0.

If, in addition, φ ≡ 1 in a neighbourhood of the origin, then

uφru in Ds,2(3) as r+.

2.2 Preliminary lemmas

Set G(u)=|u|2μ,s+F(u),g(u)=dG(u)du. In the sequel, set r0=66μ, then 2<r0q1r0q2<2s. So for any u ∈ Hs(ℝ3), we have

(2.4) |F(u)|r0C(|u|q1r0q1+|u|q2r0q2)

and

(2.5) |G(u)|r0C(|u|q1r0q1+|u|q2r0q2+|u|2s*2μ,s*).

Therefore, the Hardy-Littlewood-Sobolev inequality implies that

(2.6) |R3R3G(u(y))G(u(x))|xy|μdydx|C|G(u)|r02C(|u|q1r02q1+|u|q2r02q2+|u|2s22μ,s)

and

(2.7) |R3R3G(u(y))g(u(x))u(x)|xy|μdydx|C(|u|q1r02q1+|u|q2r02q2+|u|2s22μ,s).

It is clear that problem (2.3) is the Euler-Lagrange equations of the functional I : Hε ℝ defined by

(2.8) I(u)=12uε2122μ,SR3R3G(u(y))G(u(x))|xy|μdydx.

From (2.6) we know that I(u) is well defined on Hε and belongs to C1, with its derivative given by

(2.9) I(u),v=R3((Δ)s2u(Δ)s2v+V(εx)uv)dx12μ,s*R3R3G(u(y))g(u(x))v(x)|xy|μdydx

for all u, v ∈ Hε. Hence the critical points of I in Hε are weak solutions of problem (2.3). In the following, we will consider critical points of I using variational methods.

Firstly, we give the following Lemma, whose simple proof is omit.

Lemma 2.4

If (f1) and (f2) are satisfied, then

(2.10) 0<F(u)<f(u)u,0<G(u)<g(u)u,u0.

In addition, (f2) and (2.10) imply

(2.11) F(u)uandG(u)uareincreasingon(0,+).

For the derivative of the functional I we have the following Lemma.

Lemma 2.5

Let (V1) and (f1) hold, then

(i) I' maps bounded sets in Hs(ℝ3) into bounded sets in (Hs(ℝ3))*.

(ii) I' is weakly sequentially continuous. Namely, if un * u in Hs(ℝ3), then I'(un) * I'(u) in (Hs(ℝ3))*.

Proof. (i). Let {un} be a bounded sequence in Hs(ℝ3). For any v ∈ Hs(ℝ3), from the Hardy-Littlewood-Sobolev inequality and (2.5) it follows that

(2.12) |R3R3G(un(y))g(un(x))v(x)|xy|μdydx|C|G(un)|r0(|un|q1r0q11|v|q1r0+|un|q2r0q21|v|q2r0+|un|2s2μ,s1|v|2s)Cvε.

Then |I(un),v|Cvε. Hence, {I0(un)} is bounded in (Hs(ℝ3))*.

(ii). Assume that un * u in Hs(ℝ3). For any vC0(3) with support Ω, by Lemma 2.1 we may assume that un(x) → u(x) a.e. in ℝ3 and un → u in Lp(Ω),p<2s. We first check that if un ⇀ u in Hs(ℝ3), then

(2.13) R3R3|un(y)|2μ,s*|un(x)|2*,s2μun(x)v(x)|xy|μdydxR3R3|u(y)|2μ,s*|u(x)|2μ,82μu(x)v(x)|xy|μdydx,

as n → ∞. By the Hardy-Littlewood-Sobolev inequality, we have

|h|xy|μ|6μC|h|r0, for all hLr0(3),

and it is a linear bounded operator from Lr0(3) to L6μ(3). Choosing hn(y):=|un(y)|2μ,sLr0(3), we have

||un(y)|2μ,s|xy|μ|6μC|un|2sC.

Therefore, by Hölder inequality we can prove the sequence

(|un(y)|2μ,s|xy|μ)|un(x)|2μ,s2μun(x)

is bounded in L2s2s1(3). Then, by duality we have

R3|un(y)|2μ,s|un(x)|2μ,s2μun(x)|xy|μdyR3|u(y)|2μ,s|u(x)|2μ,s2μu(x)|xy|μdy,inL2s2s1(R3)

as n → ∞. Then (2.13) follows for every vHS(3)L2s(3).

By (f1) we have |f(u)|C(1+|u|q21) for all u ∈+, and then f (un) → f(u) in Lq2rq21(Ω). Then Hardy-Littlewood-Sobolev inequality implies that

(2.14) R3R3G(un(y))(f(un(x))f(u(x)))v(x)|xy|μdydxC|G(un)|r|f(un)f(u)|G2rq21,Ω|v|q2r0

Since F(un) is bounded in Lr0(3). we may assume that F(un) ⇀ F(u) in Lr0(3). Consequently,

R3F(un(y))|xy|μdyR3F(u(y))|xy|μdy in L6μ(3).

Moreover, as (2.12) we get R3|f(u(x))v(x)|rdx<, we infer

R3R3F(un(y))|xy|μf(u(x))v(x)dydxR3F(u(y))|xy|μf(u(x))v(x)dydxinL6μ(3),

which combining with (2.14) we have

R3R3F(un(y))|xy|μf(un(x))v(x)dydxR3F(u(y))|xy|μf(u(x))v(x)dydxinL6μ(3)

Notice that

R3R3G(u(y))g(u(x))v(x)|xy|μdydx=R3(R3|u(y)|2μ,s+F(u(y))|xy|μdy)(|u(x)|2μ,s2μu(x)+12μ,sf(u(x)))v(x)dx,

from our argument above it is then easy to prove

R3R3G(un(y))g(un(x))v(x)|xy|μdydxR3G(u(y))G(u(x))v(x)|xy|μdydxinL6μ(3).

Hence, for any vC0(3),

(2.15) I(un),vI(u),v.

Since {I0(un)} is bounded in (HS(3)) and Cn(3) is dense in Hs(ℝ3), we conclude that (2.15) holds for any v ∈ Hs(ℝ3), and so I(un)I(u)in(HS(3)).

2.3 Regularity of solutions and Pohožaev identity

The assumption (f1) is too weak for the standard bootstrap method as in [4, 15, 32]. Therefore, in order to prove regularity of solutions of (2.3) we shall rely on a nonlocal version of the Brezis-Kato estimate. Note that a special case of the regularity result of Brezis and Kato [10, Theorem 2.3] states that if u ∈ H1(ℝN) is a solution of the linear elliptic equation

Δu+u=Vu in N,

and VL(N)+LN2(N), then uLp(N) for every p ≥ 1. Similar to [33, 42], we extent this result to the fractional Choquard equation with critical growth. For the convenience of readers, we give here a short proof. We first have the following useful inequality.

Lemma 2.6

[33] Let p, q, r, t ∈ [1, +∞) and λ ∈ [0, 2] such that

1+NμN1p1t=λq+2λr.

If θ ∈ (0, 2) satisfies

min(q,r)(NμN1p)<θ<max(q,r)(11p)

and

min(q,r)(NμN1t)<2θ<max(q,r)(11t),

then for every HLpRN,KLtRNanduLqRNLrRN,

RNRNH|u(y)|θK|u(x)|2θ|xy|μdydxC(RN|H|p)1p(RN|K|t)1t(RN|u|q)λq(RN|u|r)2λr.

Applying Lemma 2.6,we have the following result, which is a nonlocal counterpart of the estimate [10, Lemma 2.1]: If VL(N)+LN2(N). then for every ε > 0, there exists Cε such that

RNV|u|2dxε2RN|u|2dx+CεRN|u|2dx.

Lemma 2.7

Let N ≥ 2s, μ ∈ (0, 2s) and θ ∈ (0, 2). If H,KL2NNμRN+L2NN+2sμRNandNμN<θ<2NμN, then for every ε > 0, there exists Cε, such that for every u ∈ Hs(ℝN),

RNRNH|u(y)|θK|u(x)|2θ|xy|μdydxε2RN|(Δ)s2u|2dx+Cε,θRN|u|2dx.

Proof. Since 0 < μ < 2s, we may assume that H = H* + H* and K = K* + K* with H,KL2NNμ(3) and H,KL2NN+2sμ(N). Applying Lemma 2.6 with q=r=2s,p=t=2NN+2sμ and λ = 0, we have

RNRNH|u(y)|θK|u(x)|2θ|xy|μdydxC|H|2NN+2sμ|K|2NN+2sμ|u|2s2,

where we use |θ1|<μN2s. Taking p=t=2NNμ,q=r=2 and λ = 2, we have |θ1|<μN and

RNRNH|u(y)|θK|u(x)|2θ|xy|μdydxC|H|2NNμ|K|2NNμ|u|22.

Similarly, with p=2NN+2sμ,t=2NNμ,q=2,r=2S and λ = 1, we have

RNRNH|u(y)|θK|u(x)|2θ|xy|μdydxC|H|2NN+2sμ|K|2NNμ|u|2|u|2s

and with p=2NNμ,t=2NN+2sμ,q=2,r=2S and λ = 1,

RNRNH|u(y)|θK|u(x)|2θ|χy|μdydxC|K|2NN2sμ|H|2NNμ|u|2|u|2s*.

By the Sobolev inequality, we have thus prove that for every u ∈ Hs(ℝN),

RNRNH|u(y)|θK|u(x)|2θ|xy|μdydxC(RN|H|2NN+2sμRN|K|2NN+2sμ)N+2sμ2NRN|(Δ)s2u|2dx+(RN|H*|2NNμRN|K|2NNμ)Nμ2NRN|u|2dx).

The conclusion follows by choosing H* and K* such that

C(RN|H|2NN+2sμRN|K|2NN+2sμ)N+2sμ2Nε2.

Now, we have the following result, which is a nonlocal Brezis-Kato type regularity estimate.

Lemma 2.8

Let N ≥ 2s and 0 < μ < 2s. If H,KL2NNμ(N)+L2NN+2sμ(N) and u ∈ Hs(ℝN) solves

(2.16) (Δ)su+u=(RNH(u(y))u(y)|xy|μdy)K(u(x)),

then u ∈ Lp(ℝN) for every p[2,NNμ2NN2s). Moreover, there exists a constant Cp independent of u such that

(RN|u|pdx)1pCp(RN|u|2dx)12.

Proof. By Lemma 2.7 with θ = 1, there exists λ > 0 such that for every φHs(N),

RNRN|HφKφ||xy|μdydx12RN|(Δ)s2φ|2dx+λ2RN|φ|2dx.

Choose sequences {Hk}k2N and {Kk}kinL2NNμ(N) such that |Hk||H|, |Kk||K|, and Hk → H and Kk → K almost everywhere in ℝN. For each k ∈ ℕ, for φ,ψHS(N), the form ak : Hs(ℝN) × Hs(ℝN) ℝ defined by

ak(φ,ψ)=RN(Δ)s2φ(Δ)s2ψ+λφψRNRNHkφKkψ|xy|μdydx

is bilinear and coercive. Therefore, applying the Lax-Milgram theorem [9, Corollary 5.8], there exists a unique solution uk ∈ Hs(ℝN) satisfies

(2.17) (Δ)suk+λuk=RN(Hkuk|xy|μdy)Kkuk+(λ1)u,

where u ∈ Hs(ℝN) is the given solution of (2.16). Moreover, we can prove that the sequences {uk}k2N converges weakly to u in Hs(ℝN) as k → ∞.

For μ > 0, we define the truncation uk,μ by

uk,μ(x)={μ if uk(x)μuk(x) if μ<uk(x)<μμ if uk(x)μ

For p ≥ 2, we have |uk,μ|p2uk,μHS(N), so we can take it as a test function in (2.17), we have

4(p1)p2RN(|(Δ)s2(uk,μ)p2|2+λuk,μ|p2|2)RNRN(Hkuk)(Kk|uk,μ|p2uk,μ)|xy|μdydx+(λ1)u|uk.μ|p2uk,μ.

By Lemma 2.7 with θ = 1, there exists C > 0 such that

RNRN|Hk||uk,μ||Kk||uk,μ|p2|uk,μ||xy|μdydx2(p1)p2RN|(Δ)s2(uk,μ)p2|2dx+CRN||uk,μ|p2|2dx.

We have thus

2(p1)p2RN|(Δ)s2(uk,μ)p2|2dxC1RN(|uk|2+|u|2)dx+Ak,μ(|Hkuk||xy|μdy)|Kkuk|,

where

Ak,μ={x3:|uk(x)|>μ}.

Since p<2NNμ, by the Hardy-Littlewood-Sobolev inequality with 1r=2Nμ2N+11p and 1t=2Nμ2N+1p,

Ak,μ(1|x|μ|Kkuk|p1dy)|Hkuk|C(RNKkuk|p1|rdx)1r(Ak,μ|Hkuk|tdx)1t

By Hölder inequality, if uk ∈ Lp(ℝN), then |Kk||uk|p1Lr(N),|Hk||uk|Lt(N), thus by Lebesgue’s dominated convergence theorem we have

limμAk,μ(1|x|μ|Kk||uk|p1dy)|Hkuk|=0.

In view of the Sobolev estimate, we have proved the inequality

limsupk(RN|uk|2s*dx)22ClimsupkRN|uk|2dx.

By iterating over p a finite number of times we cover the range p[2,NNμ2NN2s).

3 The penalized problem

In this section, we will adapt for our case an argument explored by the penalization method introduction by del Pino and Felmer [18] to overcome the lack of compactness. Let K > 2 to be determined later, and take a > 0 to be the unique number such that G(a)a=V0K where V0 is given by (V1). We define

G˜(u)={G(u) if ua,V0Ku if u>a,

and

H(x,u)=χΩ(x)G(u)+(1χΩ(x))G˜(u),

where χ is characteristic function of set Ω. From hypotheses (f1)−(f3)we get that H is a Carathéodory function and satisfies the following properties:

(g1) H(x,u)G(u)C(|u|q1+|u|q2+|u|2μ,s).

(g2) The function H(x,u)u is increasing for u > 0.

(g3)(i)

lim|u|+H(x,u)|u|2,s1=+ for s(34,1);

(ii)

lim|u|+H(x,u)|u|21μ,s2s32s(log|u|)12=+ for s=34;

(iii)

lim|u|+H(x,u)|u|2μ,s232S=+ for s(0,34).

Moreover, in order to find positive solutions, we shall henceforth consider H(x, u) = 0 for all u ≤ 0. It is easy to check that if u is a positive solution of the equation

ε2s(Δ)su+V(x)u=εμ312μ,sR3H(εx,u)|xy|μdyh(εx,u) in R3uCloc0,αR3HsR3

such that u(x) ≤ a for all x3\Ω, then H (x, u) = G(u) and therefore u is also a solution of problem (1.8). In view of this argument above, we shall deal with in the following with the penalized problem

(3.1) (Δ)su+V(εx)u=12μ,s(R3H(εx,u)|xy|μdy)h(εx,u) in 3,

and we will look for solutions uε of problem (3.1) verifying

uε(x)a for all x3\Ωε,

where Ωε={x3:εxΩ}.

The energy functional associated with (3.1) is

Jε(u)=12uε2122μ,sR3R3H(εx,u(y))H(εx,u(x))|xy|μdydx.

which is of C1 class and whose derivative is given by

Jε(u),v=R3((Δ)s2u(Δ)s2v+V(εx)uv)dx12μ,s*R3R3H(εx,u(y))h(εx,u(x))v(x)|xy|μdydx

for all u, v ∈ Hε. Hence the critical points of Jε in Hε are weak solutions of problem (3.1).

Now, we denote the Nehari manifold associated to Jε by

Nε=uHε{0}:Jε(u),u=0.

Obviously, Nε contains all nontrivial critical points of Iε. But we do not know whether Nε is of class C1 under our assumptions and therefore we cannot use minimax theorems directly on Nε. To overcome this difficulty, we will adopt a technique developed in [45, 46] to show that Nε is still a topological manifold, naturally homeomorphic to the unit sphere of Hε, and then we can consider a new minimax characterization of the corresponding critical value for Iε.

For this we denote by Hε+ the subset of Hε given by

Hε+={uHε:|supp(u+)Ωε|>0}

and Sε+=SεHε+, where Sε is the unit sphere of Hε.

Lemma 3.1

The set Hε+ is open in Hε.

Proof. Suppose by contradiction there are a sequence {un}Hε\Hε+and uHε+ such that un → u in Hε. Hence |supp(un+)Ωε|=0 for all n and un+(x)u+(x) a.e. in xΩε. So,

u+(x)=limnun+(x)=0, a.e. in xΩε.

But, this contradicts the fact that uHε+ Therefore Hε+ is open.

From definition of Sε+ and Lemma 3.1 it follows that Sε+ is a incomplete C 1,1-manifold of codimension 1, modeled on Hε and contained in the open Hε+. Hence, Hε=TuSε+u for each uSε+, where TuSε+={vHε: (u, v)ε = 0}.

In the rest of this section, we show some Lemmas related to the function Jε and the set Hε+. First, we show the functional Jε satisfying the Mountain Pass geometry.

Lemma 3.2

The functional Jε satisfies the following conditions:

(i) There exist α, ρ > 0 such that Jε(u) ≥ α with ‖u‖ε = ρ;

(ii) There exists e ∈ Hε satisfying ‖e‖ε > ρ such that Jε(e) < 0.

Proof. (i). For any u ∈ Hε\{0}, it follows from (g1) and the Hardy-Littlewood-Sobolev inequality that

(3.2) |R3R3H(εx,u(y))H(εx,u(x))|xy|μdydx|C(|u|q1r02q1+|u|q2r02q2+|u|2s22μ,s)

Hence,

Jε(u)=12R3(|(Δ)s2u|2+V(εx)u2)dx122μ,sR3R3H(εx,u(y))H(εx,u(x))|xy|μdydx.12uε2C1u2q1C2u2q2C3u22μ,s*.

Therefore, we can choose positive constants α, ρ such that

Jε(u)α with uε=ρ.

(ii). Fix a positive function u0Hε+ with suppu0Ωε, and we set

ψ(t):=Σε(tu0u0ε)>0,

where

Σε(u)=122μ,sR3R3H(εx,u(y))H(εx,u(x))|xy|μdydx.

Since H(εx, u0) = F(u0) and by using Lemma 2.4, we deduce that

(3.3) ψ(t)=Σε(tu0u0ε)u0u0ε=R3R3(F(tu0u0ε)f(tu0u0ε)u0u0ε|xy|μ)dxdy=22μ,stR3R3122μ,s(F(tu0u0ε)f(tu0u0ε)tu0u0ε|xy|μ)dxdy22μ,stψ(t).

Integrating (3.3) on [1, t‖u0ε] with t>1u0ε, we have

Σε(tu0)Σε(u0u0ε)u0ε22μ,st22μ,s

Therefore, we have

Jε(tu0)=t22uε2Σε(tu0)C1t2C2t22μ,sfort>1u0ε.

Taking e = tu0 with t sufficiently large, we can see (ii) holds.

Since f is only continuous, the next two results are very important because they allow us to overcome the non-differentiability of Nε and the incompleteness of Sε+.

Lemma 3.3

Assume that the potential V satisfies (V1)−(V2) and the functional f satisfies (f1)−(f3). Then the following properties hold:

(A1)For each uSε+. let φu : ℝ+ be given by φu(τ) = Jε(τu). Then there exists a unique τu > 0 such that φu(τ)>0in(0,τu)andφu(τ)<0in(τu,)

(A2)There is a σ > 0 independent on u such that τu > σ for all uSε+. Moreover, for each compact setWSε+ there is CW > 0 such that τuCW for all u ∈ W.

(A3)The map m^ε:Hε+Nε given by mε(u) = τuu is continuous and mε:=m^ε|Sε+ is a homeomorphism between Sε+ and Nε. Moreover, mε1(u)=uuε.

Proof. (A1) From Lemma 3.2, it is sufficient to note that, φu(0) = 0, φu(τ) > 0 when τ > 0 is small and φu(τ) < 0 when τ > 0 is large. Since φu ∈ C1(ℝ+, ℝ), there is τu > 0 global maximum point of φu and φu(τu)=0. Thus,Jε(τuu)(τuu)=0 and τuu ∈ Nε. We see that τu > 0 is the unique positive number such that φu(τu)=0. Indeed, suppose by contradiction that there are τ1 > τ2 > 0 with φu(τ1)=φu(τ2)=0. Then, for i = 1, 2 we have that

(3.4) 0=R3R31|xy|μ(H(εx,τ1u(u))τ1h(εx,τ1u(x))u(x)H(εx,τ2u(u))τ2h(εx,τ2u(x))u(x))dydx.

By Lemma 2.4 and (h2) we know that h(εx,τu(x))u(x) andH(εx,τu(x))τ are positive and increasing in τ. Then (3.4) in impossible and (A1) is proved.

(A2) Suppose uSε+, then as (2.5) we have

τuR3R3H(εx,u(y))h(εx,u(x))u(x)|xy|μdydxC(τu2q1+τu2q2+τu22μ,s).

From previous inequality we obtain σ > 0 independent on u, such that τu > σ.

Finally, if WSε+ is compact, and suppose by contradiction that there is {un} ⊂ W such that τn := τn:=τun. Since W is compact, there is u ∈ W such that un → u in Hε. Then uWSε+. By (h2) we obtain

0Jε(τnun)τn2=12122μ,sR3R3H(εx,τnun(y))|xy|μτnH(εx,τnun(x))τndydx,

which yields a contradiction. Therefore (A2) is true.

(A3) First of all we observe that m^ε,mε andmε1 are well defined. In fact, by (A1), for each uHε+. there exists a unique τu > 0 such that τuu ∈ Nε, hence there is a unique m^ε(u)=τuuNε. On the other hand, if uNε thenuHε+. Therefore, mε1(u)=uuεSε+, is well defined and it is a continuous function. Since

mε1(mε(u))=mε1(tuu)=τuuτuuε=u,uSε+,

we conclude that mε is a bijection.

To prove m^ε:Hε+Nε is continuous, let {un}Hε+anduHε+ be such that un → u in Hε. By (A2), there is a τ0 > 0 up to a subsequence such that τn:=τunτ0. Since τnun Nε we obtain

τn2unε2=12μ,sR3R3H(εx,τnun(y))h(εx,τnu(x))τnun(x)|xy|μdydx,nN.

By Lemma 2.5 and passing to the limit as n → ∞, it follows that

τ02uε2=12μ,sR3R3H(εx,τ0un(y))h(εx,τ0u(x))τ0un(x)|xy|μdydx,

which means that τ0uNεandτu=τ0. This proves m^ε(un)m^ε(u) inHε+. So, m^ε,mε are continuous functions and (A3) is proved.

Now we define the functions

Ψ^ε:Hε+RandΨε:Sε+R,

by Ψ^ε(u)=Iε(m^ε(u)) and Ψε:=Ψ^ε|Sε+. The next result is a direct consequence of Lemma 3.3. The details can be seen in the relevant material from [46]. For the convenience of the reader, here we do a sketch of the proof.

Lemma 3.4

Assume that (V1)− (V2) and (f1)− (f3) are satisfied. Then: (B1)Ψ^εC1(Hε+,R) and

Ψ^ε(u)v=m^ε(u)εuεJε(m^ε(u))v,uHε+andvHε.

(B2)ΨεC1(Sε+,R) and

Ψε(u)v=mε(u)εJε(mε(u))v,vTuSε+.

(B3)If {un} is a (PS)c sequence of Ψε, then {mε(un)} is a (PS)c sequence of Jε. If {un} ⊂ Nε is a bounded (PS)c sequence for Jε,then{mε1(un)} is a (PS)c sequence of Ψε.

(B4)u is a critical point of Ψε if and only if, mε(u) is a critical point of Jε. Moreover, corresponding critical values coincide and

infSε+Ψε=infNεIε.

Proof. (B1) Let uHε+ and v ∈ Hε. From definition of Ψ^ε andtu and the mean value theorem, we obtain

Ψ^ε(u+hv)Ψ^ε(u)=Jε(τu+hv(u+hv))Jε(τuu)Jε(τu+hv(u+hv))Jε(τu+hvu)=Jε(τu+hv(u+θhv))τu+hvhv,

where |h| is small enough and θ ∈ (0, 1). Similarly,

Ψ^ε(u+hv)Ψ^ε(u)Jε(τu(u+hv))Jε(τuu)=Jε(τu(u+ςhv))τuhv,

where ζ ∈ (0, 1). Since the mapping u ⟼ τu is continuous according to Lemma 3.3, we see combining these two inequalities that

limh0Ψ^ε(u+hv)Ψ^ε(u)h=τuJε(τuu)v=m^ε(u)εuεJε(m^ε(u))v.

Since Jε ∈ C1, it follows that the Gâteaux derivative of Ψ^ε is bounded linear in v and continuous on u. From [50] we know that Ψ^εC1(Hε+,R) and

Ψ^ε(u)v=m^ε(u)εuεJε(m^ε(u))v,uHε+andvHε.

The item (B1) is proved.

(B2) The item (B2) is a direct consequence of the item (B1).

(B3) We first note that Hε=TuSε+Ru for every uSε+ and the linear projection P:HεTuSε+ defined by P(v + τu) = v is continuous, namely, there is C > 0 such that

(3.5) vεCv+τuε,uSε+,vTuSε+andτR.

Moreover, by (B1) we have

(3.6) Ψε=supvTuSε+,vε=1Ψε(u)v=wεsupvTuSε+,vε=1Jε(w)v,

where w=mε(u).SincewNε, we conclude that

(3.7) Jε(w)u=Jε(w)wwε=0.

Hence, from (3.5) and (3.7) we have

Ψε(u)wεJε(w)CwεsupvTuSε+{0}Jε(w)vvε=CΨε(u),

which showing that

(3.8) Ψε(u)wεJε(w)CΨε(u),uSε+.

Since w ∈ Nε, we have ‖w‖ ≥ > 0. Therefore, the inequality in (3.8) together with Jε(w)=Ψε(u) imply the item (B3).

(B4) It follow from (3.8) that Ψε(u)=0 if and only if Jε(w)=0. The remainder follows from definition of Ψε.

As in [46], using the mountain pass theorem without the (PS) condition, we get the existence of a (PS) sequence {un} ⊂ Hε with

cε=infuNεIε(u)=infuHε+maxτ>0Iε(τu)=infuSε+maxτ>0Iε(τu)>0.

Lemma 3.5

Suppose that (f1) − (f3) hold. Assume that {un}Nεisa(PS)c sequence with

0<cεc<2μ,s122μ,sSH,L2μ,s2μ,s1.

Then {un} is bounded in Hε. Moreover, {un} cannot be vanishing, namely there exist r, δ > 0 and a sequence {yn} ⊂3 such that

lim infnBr(yn)|un|2dxδ.

Proof. We first prove the boundedness of {un}. Argue by contradiction we suppose that {un} is unbounded in Hε. Without loss of generality, assume that unε.Letvn:=ununε up to a subsequence, then there exists v ∈ Hε such that

unuinHε,unuinLlocr(R3),2r<2s,un(x)u(x)a.e.inR3.

If vn is vanishing, i.e.

limnsupyR3Br(y)vn2(x)dx=0,

then Lemma 2.2 implies that vn0inLr0q1(R3)andLr0q2(R3). By (2.4) and (2.5) we get

(3.9) R3R3F(τvn(x))F(τvn(y))|xy|μdxdy0,R3R3F(τvn(y))|xy|μ|τvn(x)|2μ,sdxdy0.

Then for sufficiently large n we have

cε+on(1)=Iε(un)supτ0Iε(τvn)supτ0(τ22τ22μ,s22μ,sSH,L2μ,s)+on(1)=2μ,s122μ,sSH,L2μ,s2μ,s1+on(1),

which is a contradiction. Therefore, {vn} is non-vanishing, namely there exists yn3 and δ > 0 such that

(3.10) Br(yn)vn2(x)dx>δ.

Denote vn~()=vn(+yn), then we can assume that

vn~v~inHε,vn~v~inLlocr(R3),2r<2s,vn~(x)v~(x)a.e.inR3.

With the use of (3.10), we have v~0. Then there exists a measure set 𝛬 such that v~(x)0 for x ∈ 𝛬. Let |un¯|:=|vn~|unε. Then |un¯(x)|+ for xΛ. By Lemma 2.4, we have

ΛΛ1|xy|μF(un¯(y))|un¯(y)||vn~(y)|F(un¯(x))|un¯(x)||vn~(x)|dydx+.

Therefor, Lemma 2.4 and Fatou Lemma imply that

limnR3R31|xy|μF(un¯(y))|un¯(y)||vn~(y)|F(un¯(x))|un¯(x)||vn~(x)|dydx=+.

Namely

limnR3R31|xy|μF(un(y))unεF(un(x))unεdydx=+.

Then,

cεunε+on(1)=Iε(un)unε,

which is a contradiction. Therefore, {un} is bounded in Hs(ℝ3).

Next we show the second conclusion. We argue by contradiction, if {un} is vanishing, then similar to (3.9)

we have

(3.11) cε+on(1)=Iε(un)=12unε2122μ,sR3R3|un(y)|2μ,s|un(x)|2μ,s|xy|μdydx+on(1),

and

(3.12) 0=unε2R3R3|un(y)|2μ,s|un(x)|2μ,s|xy|μdydx+on(1).

If unε0, then it follows from (3.11) and (3.12) that cε = 0, which is impossible. Then unε0 and by virtue of (3.12) we get

unε2SH,L2μ,sunε22μ,s+on(1).

Hence,

lim infnunε2SH,L2μ,s2μ,s1.

From (3.11) and (3.12) we deduce that

(3.13) cε+on(1)=Iε(un)2μ,s122μ,sSH,L2μ,s2μ,s1,

which is a contradiction. Therefore, {un} is non-vanishing.

Lemma 3.6

Assume (V1) − (V2) and (f1) − (f3) hold, let {un} be a (PS)c sequence for Jε with c ∈ [cε,2μ,s122μ,sSH,L2μ,s2μ,s1) Then, for each η > 0 there exists R = R(η) > 0 such that

lim supnR3BR|(Δ)s2u|2+V(εx)un2dx<η.

Proof. By Lemma 3.5, we can have {un} is bounded in Hε. Therefore, we may assume that unu in Hε and unuinLlocr(R3) for any r[2,2s). Fix R > 0 and let ψRC(R3) be such that ψR=0inBR2(0),ψR=1 in BRc,ψR[0,1] and |ψR|CR, where C is a constant independent of R. Since {unψR} is bounded we can see that

R3((Δ)s2un(Δ)s2(unψR)+V(εx)ψRun2)dx=Jε(un),unψR+12μ,sR3(1|x|μH(εx,un))h(εx,un)unψRdx=on(1)+12μ,sR3(1|x|μH(εx,un))h(εx,un)unψRdx.

By Lemma 2.8, taking

H(u):=|u|2μ,s+F(u)u,K(u):=|u|2μ,s2μ+12μ,sf(u)L23μ(R3)+L63+2sμ(R3),

we have

R3(1|x|μH(εx,un))h(εx,un)undxR3(1|x|μH(un)un)K(un)undxε2RN|(Δ)s2u|2dx+CεRN|u|2dx.

For nn0 and ε > 0 fixed, take R > 0 big enough such that ΩεBR/2. Then we have

R3BR/2(|(Δ)s2un|2+V(εx)un2)dx12μ,sR3BR/2(1|x|μH(εx,un))h(εx,un)undx+on(1)R3R3(un(x)un(y))(ψR(x)ψR(y))|xy|3+2sun(y)dxdy.

which means

12R3BR/2(|(Δ)s2un|2+V(εx)un2)dxon(1)R3R3(un(x)un(y))(ψR(x)ψR(y))|xy|3+2sun(y)dxdy

Now, we note that the Hölder inequality and the boundedness of {un} imply that

|R3R3(un(x)un(y))(ψR(x)ψR(y))|xy|3+2sun(y)dxdy|(R3R3|un(x)un(y)|2|xy|3+2sdxdy)12(R3R3|ψR(x)ψR(y)|2|xy|3+2sun2(y)dxdy)12C(R3R3|ψR(x)ψR(y)|2|xy|3+2sun2(y)dxdy)12.

Therefore, it is enough to prove that

limRlim supnR3R3|ψR(x)ψR(y)|2|xy|3+2sun2(y)dxdy=0

to conclude our result.

Let us note that ℝ3 × ℝ3 can be written as

R3×R3=((R3B2R)×(R3B2R))((R3B2R)×B2R)(B2R×R3):=X1X1X3.

Then

(3.14) R3×R3|ψR(x)ψR(y)|2|xy|3+2sun2(x)dxdy=X1|ψR(x)ψR(y)|2|xy|3+2sun2(x)dxdy+X2|ψR(x)ψR(y)|2|xy|3+2sun2(x)dxdy+X3|ψR(x)ψR(y)|2|xy|3+2sun2(x)dxdy.

Now, we estimate each integral in (3.14). Since ψR=1inR3B2R, we have

(3.15) X1|ψR(x)ψR(y)|2|xy|3+2sun2(x)dxdy=0.

Let k > 4, we have

X2=(R3B2R)×B2R=(R3BkR)×B2R(BkRB2R)×B2R.

Let us note that, if (x,y)(R3BkR)×B2R, then

|xy||x||y||x|2R>|x|2.

Therefore, taking into account 0ψR1,|ψR|CR and applying Hölder inequality, we can see

(3.16) X2|ψR(x)ψR(y)|2|xy|3+2sun2(x)dxdy=R3BkRB2R|ψR(x)ψR(y)|2|xy|3+2sun2(x)dxdy+BkRB2RB2R|ψR(x)ψR(y)|2|xy|3+2sun2(x)dxdy25+2sR3BkRB2R|un(x)|2|x|3+2sdxdy+CR2BkRB2RB2R|un(x)|2|xy|3+2(s1)dxdyCR3R3BkR|un(x)|2|x|3+2sdxdy+CR2(kR)2(1s)BkRB2Run2(x)dxdyCR3(R3BkR|un(x)|2sdx)22s(R3BkR1|x|92s+3)2s3+Ck2(1s)R2sBkRB2Run2(x)dxCk3(R3BkR|un(x)|2sdx)22s+Ck2(1s)R2sBkRB2Run2(x)dxCk3+Ck2(1s)R2sBkRB2Run2(x)dx.

Now, fix ε(0,12), and we note that

(3.17) X3|ψR(x)ψR(y)|2|xy|3+2sun2(x)dxdyB2RBεRR3|ψR(x)ψR(y)|2|xy|3+2sun2(x)dxdy+BεRR3|ψR(x)ψR(y)|2|xy|3+2sun2(x)dxdy.

Let us estimate the first integral in (3.17). First, we have

B2RBεRR3{y:|xy|<R}|ψR(x)ψR(y)|2|xy|3+2sun2(x)dxdyCR2sB2RBεRun2(x)dx

and

B2RBεRR3{y:|xy|R}|ψR(x)ψR(y)|2|xy|3+2sun2(x)dxdyCR2sB2RBεRun2(x)dx.

Then

(3.18) B2RBεRR3|ψR(x)ψR(y)|2|xy|3+2sun2(x)dxdyCR2sB2RBεRun2(x)dx.

By using the definition of ψR,ε(0,1) and ψR1, we have

(3.19) BεRR3|ψR(x)ψR(y)|2|xy|3+2sun2(x)dxdy=BεRR3BR|ψR(x)ψR(y)|2|xy|3+2sun2(x)dxdy4BεRR3BRun2(x)|xy|3+2sdxdyCBεRun2(x)dx(12ε)R1r2+2sdr=C((12ε)R)2sBεRun2(x)dx

where we use the fact that if (x,y)BεR×(R3BR), then |xy|>(12ε)R. Taking into account (3.17)-(3.19) we deduce

(3.20) X3un2(x)|ψR(x)ψR(y)|2|xy|μdxdyCR2sB2RBϵR|un(x)|2dx+C((1ϵ)R)2sBϵRun2(x)dx.

Putting together (3.14),(3.15),(3.16) and (3.20), we can infer

(3.21) R3×R3un2(x)|ψR(x)ψR(y)|2|xy|μdxdyCk3+Ck2(1s)R2sBkRB2Run2(x)dx+CR2sB2RBεRun2(x)dx+C((1ϵ)R)2sBϵRun2(x)dx.

Since {un} is bounded in Hε, we may assume that unuinLloc2(R3) for some uHε. Then, taking the limit as n → ∞ in (3.21), we have

lim supnR3×R3un2(x)|ψR(x)ψR(y)|2|xy|μdxdyCk3+Ck2(1s)R2sBkRB2Ru2(x)dx+CR2sB2RBεRu2(x)dx+C((1ϵ)R)2sBϵRu2(x)dxCk3+Ck2(BkRB2R|u|2s(x)dx)22s+C(B2RBεR|u|2s(x)dx)22s+C(ε1ε)2s(BεR|u(x)|2s)22s,

where in the last passage we use Hölder inequality.

Since uL2s(R3),k>4andε(0,12), we obtain

lim supRBkRB2R|u(x)|2sdx=lim supRB2RBεR|u(x)|2sdx=0.

Choosing ε=1k, we get

lim supRlim supnR3×R3un2(x)|ψR(x)ψR(y)|2|xy|μdxdylimklim supR(Ck3+Ck2(BkRB2R|u(x)|2sdx)22s+C(B2RBRk|u(x)|2sdx)22s+C(1k1)2s(BRk|u(x)|2sdx)22s)limkCk3+C(1k1)2s(BRk|u(x)|2sdx)22s=0,

which complete our proof.

Lemma 3.7

Under the conditions of Lemma 3.6, the functional Jε satisfies the (PS)c condition for all c ∈ [cε,2μ,s122μ,sSH,L2μ,s2μ,s1).

Proof. Since {un} is bounded in Hε, we may assume

unuinHε,unuinLlocr(R3),2r<2s,un(x)u(x)a.e.inR3.

Let us prove that unuinHε asn. Setting ωn=unuε2, we have

(3.22) ωn=Jε(un),unJε(un),u+12μ,sR3(1|x|μH(εx,un))h(εx,un)(unu)dx+on(1).

Note that Jε(un),un=Jε(un),u=on(1) so we only need to show that

(3.23) R3(1|x|μH(εx,un))h(εx,un)(unu)dx=on(1).

Similar the proof in Lemma 2.7, we can see that

(3.24) 1|x|μH(εx,un)1|x|μH(εx,u)inL6μ(R3).

By using Lemma 2.1, we have

(3.25) limnBR1|x|μH(εx,un)h(εx,un)(unu)dx=0.

By Lemma 2.1 and 3.6, for any η > 0 there exists R=R(η)>0 such that

lim supnR3BR|1|x|μH(εx,un)h(εx,un)un|dxC1η

and

lim supnR3BR|1|x|μH(εx,un)h(εx,un)u|dxC2η

Taking into account the above limits we can deduce that

limnR3(1|x|μH(εx,un))h(εx,un)(unu)dx=0.

Lemma 3.8

The functional Ψε verifies the (PS)c condition in Sε+forallc[cε,2μ,s122μ,sSH,L2μ,s2μ,s1).

Proof. Let {un}Sε+ be a (PS)c sequence for Ψε. Thus, Ψε(un)candΨε0, where is the norm in the dual space (TunSε+). It follows from Lemma 3.4-(B3) that {mε(un)}isa(PS)c sequence for Iε in Hε. From Lemma 3.7 we see that there is a uSε+ such that mε(un)mε(u) in Hε. From Lemma 3.3-(A3), it follows that unuinSε+.

4 The autonomous problem

Since we are interestd in giving a multiplicity result for the modified problem, we start by considering the limit problem associated to (1.8), namely, the problem

(4.1) (Δ)su+V0u=(R3|u(y)|2μ,s+F(u(y))|xy|μdy)(|u|2μ,s2u+12μ,sf(u))inR3.

Set G(u)=|u|2μ,s+F(u),g(u)=dG(u)du, then the equation (4.1) changes into

(4.2) (Δ)su+V0u=12μ,s(R3G(u(y))|xy|μdy)g(u)inR3.

which has the following associated functional

I0(u)=12R3(|(Δ)s2u|2+V0u2)dx122μ,sR3R3G(u(y))G(u(x))|xy|μdydx.

The functional I0 is well defined on the Hilbert space H0 = Hs(ℝ3) with the inner product

(u,v)0=R3(Δ)s2u(Δ)s2vdx+R3V0uvdx,

and the norm

u02=R3|(Δ)s2u|2dx+R3V0u2dx.

We denote the Nehari manifold associated to I0 by

N0={uH0{0}:I0(u),u=0},

and by H0+ the open subset of H0 given by

H0+={uH0:|supp(u+)|>0},

and S0+=S0H0+, where S0 is the unit sphere of H0.

As in section 3, S0+ a incomplete C1,1. manifold of codimension 1, modeled on H0 and contained in the open H0+. Thus, H0=TuS0+Ru for each uS0+, where TuS0+={vH0:(u,v)0=0}.

Next we have the following Lemmas and the proofs follow from a similar argument used in the proofs of Lemma 3.3 and Lemma 3.4.

Lemma 4.1

Let V0 be given in (V1) and the functional f satisfies (f1)−(f3). Then the following properties hold:

(a1)For each uH0+, let ϕu:R+R be given by ϕu(τ)=I0(τu). Then there exists a unique τu > 0 such that ϕu(τ)>0in(0,τu)andϕu(τ)<0in(τu,).

(a2)There is a σ > 0 independent on u such that τu > σ for all uS0+. Moreover, for each compact setWS0+ there is CW > 0 such that τuCW for all u ∈ W.

(a3) The map m^:H0+N0 given by m^(u)=τuu is continuous and m:=m^|S0+ is a homeomorphism between S0+ and N0. Moreover, m1(u)=uu0.

We define the applications

Ψ^0:H0+RandΨ0:S0+R,
byΨ^0(u)=I0(m^(u))andΨ0:=Ψ^0|S0+.

Lemma 4.2

Let V0 be given in (V1) and (f1) − (f3) are satisfied. Then: (b1)Ψ^0C1(H0+,R) and

Ψ^0(u)v=m^(u)0u0I0(m^(u))v,uH0+andvH0.

(b2)Ψ0C1(S0+,R) and

Ψ0(u)v=m(u)0I0(m(u))v,vTuS0+.

(b3)If {un} is a (PS)c sequence of Ψ0, then {m(un)} is a (PS)c sequence of I0. If fung ⊂ N0 is a bounded (PS)c sequence for I0, then {m−1(un)} is a (PS)c sequence of Ψ0.

(b4)u is a critical point of Ψ0 if and only if, m(u) is a critical point of I0. Moreover, corresponding critical values coincide and

infS0+Ψ0=infN0I0.

As in the previous section, we have the following variational characterization of the infimum of I0 over N0:

cV0=infuN0I0(u)=infuH0+maxτ>0I0(τu)=infuS0+maxτ>0I0(τu)>0.

The next Lemma allows us to assume that the weak limit of a (PS)c sequence is non-trivial.

Lemma 4.3

Let {un} ⊂ H0 be a (PS)c sequence with c[cV0,2μ,s122μ,sSH,L2μ,s2μ,s1) for I0. Then, only one of the following conclusions holds.

(i) un 0 in H0, or

(ii) There exist a sequence {yn} ⊂3 and constants R, β > 0 such that

lim infn+Br(yn)un2dxβ>0.

Proof. Suppose (ii) does not occur. Then, for any R > 0, we have

limn+supyR3Br(y)un2dx=0.

Similarly to Lemma 3.5, we have {un} is bounded in H0, then by Lemma 2.2, we have

un0inLr(R3)forr(2,2s).

Thus, by (f1) we have

R3R3G(u(y))g(u(x))u(x)|xy|μdydx=on(1).

Recalling that I0(un)un0, we get

un02=on(1).

Therefore the conclusion follows.

From Lemma 4.3 we can see that, if u is the weak limit of a (PS)cV0 sequence {un} for the functional I0, then we can assume u0. Otherwise we would have un 0 and once it doesn’t occur un 0, we conclude from Lemma 4.3 that there exist {yn} ⊂3 and R, β > 0 such that

lim infn+Br(yn)un2dxβ>0.

Then set vn(x)=un(x+yn), making a change of variable, we can prove that {vn} ia also a (PS)cV0 sequence for the functional I0, it is bounded in H0 and there is v ∈ H0 such that vn → v in H0 with v0.

Next we devote to estimating the least energy cV0. Recall that the best Sobolev constant Ss of the embedding Ds,2(R3)L2s(R3) is defined by

Ss:=infuDs,2(R3)uDs,2(R3)2|u|2s2

In particular, we consider the following family of functions Uε defined as

Uε(x)=ε2s32u¯(xεSs12s),u¯=un(x)|u0|2s,

for ε > 0 and x ∈3, the minimizer of Ss (see,[41]), which satisfies

(Δ)su=|u|2s2u,xR3.

Then, by a simple calculation, we know

U~(x)=Ss(3μ)(2s3)4(2+2sμ)C(μ)2s32(3+2sμ)Uε(x)

is the unique minimizer for SH,L that satisfies

(Δ)su=(R3|u(y)|2μ,s|xy|μ)|u|2μ,su,inR3.

Moreover,

R3|(Δ)s2U~|2dx=R3R3|U~(x)|2μ,s|U~(y)|2μ,s|xy|μdxdy=SH,L6μ3μ+2s.

Let φC0(R3, [0, 1]) and small δ > 0 be such that φ ≡ 1 in B(0) and ϕ0 in R3B2δ(0). For any ε > 0, define the best function by uε=φUε.

Similar to [22, Lemma 1.2], we can easily draw the following conclusion.

Lemma 4.4

The constant SH,L defined in (2.2) is achieved if and only if

u=C(bb2+|xa|2)32s2,

where C > 0 is a fixed constant, a ∈3 and b > 0 are parameters. Furthermore,

SH,L=SsC(μ)12μ,s.

Proof

We sketch the proof for the completeness of this paper. By the Hardy-Littlewood-Sobolev inequality,we have

SH,L1C(μ)12μ,suDs,2(R3){0}R3|(Δ)s2u|2dx|u|2s2=SsC(μ)12μ,s.

On the other hand, the equality in the Hardy-Littlewood-Sobolev (1.6) holds if and only if

f(x)=h(x)=C(bb2+|xa|2)6μ2,

where C > 0 is a fixed constant, aR3andb>0 are parameters. Thus

(R3R3|u(x)|2μ,s|u(y)|2μ,s|xy|μdxdy)12μ,s=C(μ)12μ,s|u|2s2,

if and only if

u=C(bb2+|xa|2)32s2.

Then, by the definition of SH,L, we get

(4.3) SH,LR3|(Δ)s2u|dx(R3R3|u(y)|2μ,s|u(x)|2μ,s|xy|μdydx)12μ,s=1C(μ)12μ,sR3|(Δ)s2u|2dx|u|2s2

an thus we have

SH,LSsC(μ)12μ,s.

From the arguments above, we know that SH,L is achieved if and only if u=C(bb2+|xa|2)32s2 and

SH,L=SsC(μ)12μ,s.

Next. repeat the proof in [22, Lemma 1.3],we have the following important information about the best constant SH,L.

Lemma 4.5

For every open subset ΩR3 , we have

(4.4) SH,L(Ω):=infuD0s,2(Ω){0}Ω|(Δ)s2u|2dx(ΩΩ|u(y)|2μ,s|u(x)|2μ,s|xy|μdydx)12μ,s=SH,L

where SH,L(Ω) is never achieved except when Ω=R3.

Proof

It is clear that SH,LSH,L(Ω) by D0s,2(Ω)Ds,2(R3). Let {un}C0(R3) be a minimizing sequence for SH,L. We make translations and dilations for {un} by choosing yn3 and τn > 0 such that

vn:=τn32s2un(τnx+yn)C0(Ω),

which satisfies

R3|(Δ)s2vn|2dx=R3|(Δ)s2un|2dx

and

ΩΩ|vn(y)|2μ,s|vn(x)|2μ,s|xy|μdxdy=R3R3|un(y)|2μ,s|un(x)|2μ,s|xy|μdxdy.

Hence SH,L(Ω)SH,L. Moreover, since U~(x) is the only class of functions such that the equality holds in the Hardy-Littlewood-Sobolev inequality, we know that SH,L(Ω) is never achieved except for Ω=R3.

Lemma 4.6

(4.5) R3R3|uε(x)uε(y)|2|xy|3+2sdxdySs32s+O(ε32s),
(4.6) |uε|22=Cε2s+O(ε32s)if4s<3,Cε2slog(1ε)+O(ε2s)if4s=3,Cε32s+O(ε2s)if4s>3,
(4.7) R3R3|uε(x)|2μ,s|uε(y)|2μ,s|xy|μdxdyC(μ)32sSH,L6μ2sO(ε6μ2),

In addition, if q<2μ,s, then there holds

(4.8) BδBδ|Uε(x)|q|Uε(y)|q|xy|μdxdy=O(ε6μq(32s)).

Proof

For the proof of (4.5) and (4.6), we can see that in [41]. So we only need to estimate (4.7) and (4.8). Concerning (4.7), similar to [2, Lemma 7.1], we have

(4.9) R3R3|uε(x)|2μ,s|uε(y)|2μ,s|xy|μdxdyBδBδ|uε(x)|2μ,s|uε(y)|2μ,s|xy|μdxdy=R3R3|Uε(x)|2μ,s|Uε(y)|2μ,s|xy|μdxdy2R3BδBδ|Uε(x)|2μ,s|Uε(y)|2μ,s|xy|μdxdyR3BδR3Bδ|Uε(x)|2μ,s|Uε(y)|2μ,s|xy|μdxdy:=C(μ)32sSH,L6μ2s2AB,

where

A=R3BδBδ|Uε(x)|2μ,s|Uε(y)|2μ,s|xy|μdxdy,B=R3BδR3Bδ|Uε(x)|2μ,s|Uε(y)|2μ,s|xy|μdxdy.

By direct computation, we have

(4.10) A=R3BδBδ|Uε(x)|2μ,s|Uε(y)|2μ,s|xy|μdxdy,=Cε6μR3BδBδ1(ε2b2+x2Ss1s)6μ21|xy|μ1(ε2b2+y2Ss1s)6μ2dxdyCε6μ(R3Bδ1(ε2b2+x2Ss1s)3dx)6μ6(Bδ1(ε2b2+y2Ss1s)3dy)6μ6Cε6μ(R3Bδ1|x|6Ss3sdx)6μ6(0Δr2(ε2b2+r2Ss1s)3dr)6μ6=O(ε6μ2)

and

(4.11) B=R3BδR3Bδ|Uε(x)|2μ,s|Uε(y)|2μ,s|xy|μdxdyCε6μR3BδR3Bδ1(ε2b2+x2Ss1s)6μ21|xy|μ1(ε2b2+y2Ss1s)6μ2dxdyCε6μ(R3Bδ1(ε2b2+x2Ss1s)3dx)6μ6(R3Bδ1(ε2b2+y2Ss1s)3dy)6μ6Cε6μ(R3Bδ1|x|6Ss3sdx)6μ6(R3Bδ1|y|6Ss3sdy)6μ6=O(ε6μ)

It follows from (4.9) to (4.11) that

R3R3|uε(x)|2μ,s|uε(y)|2μ,s|xy|μdxdyC(μ)32sSH,L6μ2sO(ε6μ2)O(ε6μ)=C(μ)32sSH,L6μ2sO(ε6μ2).

Then (4.7) follows. Now we prove (4.8). If q<2μ,s, we have

BδBδ|Uε(x)|q|Uε(y)|q|xy|μdxdy=O(ε6μq(32s))=Cεq(32s)BδBδ1(ε2b2+x2Ss1s)32s2q1|xy|μ1(ε2b2+y2Ss1s)32s2qdxdyCεq(32s)(Bδ1(ε2b2+x2Ss1s)32s2q66μdx)6μ6(1(ε2b2+y2Ss1s)32s2q66μdy)6μ6Cεq(32s)(0εr2(ε2b2+r2Ss32s2q66μdr)6μ3=O(ε6μq(32s)

Lemma 4.7

Suppose that (f1) − (f3) hold. Then the number cV0 satisfies that

0<cV0<2μ,s122μ,sSH,L2μ,s2μ,s1.

Proof

By the definition of cV0 it suffices to prove that there exists vN0 such that

(4.12) I0(v)<2μ,s122μ,sSH,L2μ,s2μ,s1.

By Lemma 4.1, there exists τε>0 such that τεuεN0. We claim that for ε > 0 small enough, there exist A1 and A2 independent of ε such that

(4.13) 0<A1τεA2<.

Indeed, note that N0 is bounded away from 0, we have that τεA1>0 using (4.5) and (4.6). Moreover, since I0(τεuε),τεuε=0 from (4.7) we have that

τε22μ,sC(τε2uε02+τε2q1uε02q1+τε2q2uε02q2+τεq1+2μ,suε0q1+2μ,s+τεq2+2μ,suε0q2+2μ,s).

Using (4.5) and (4.6) again, there exists A2>0 such that τεA2. Then (4.13) holds true.

Now we estimate I0(τεuε). Note that

(4.14) I0(τεuε)(τε22R3|(Δ)s2uε(x)|2dxτε22μ,s22μ,sR3R3|uε(y)|2μ,s|uε(x)|2μ,s|xy|μdydx)+(τε22R3V0uε2dx122μ,sR3R3F(τεuε(y))F(τεuε(x))|xy|μdydx):=I1+I2.

For I1, we set

A:=R3|(Δ)s2uε(x)|2dx,B=R3R3|uε(y)|2μ,s|uε(x)|2μ,s|xy|μdydx,

and consider the function θ:[0,)R defined by

θ(τ)=12Aτ2τε22μ,s22μ,sB,

we have that τ0=(AB)122μ,s2 is a maximum point of θ and

θ(τ0)=2μ,s122μ,sA2μ,s2μ,s1B112μ,s.

Combining with Lemma 4.3, (4.5) and (4.7) we have

(4.15) I12μ,s122μ,sSH,L2μ,s12μ,s+O(ε32s)+O(ε6μ2).

For I2, given A0 > 0, we invoke (f3) to obtain R=R(A0)>0 such that, for xR3,tR,

(4.16) F(x,t)A0t2μ,s1if3<4s,A0t2μ,s2s32s(logt)12if3=4s,A0t2μ,s2s32sif3>4s.

By (4.6) and (4.16), we need to estimate I2 in three cases. Since the argument is similar, we only consider the case that 3 < 4s. For |x|<ε<δ, noting that φ1 inBδ(0), by the definition of uε and (4.13), we get a constant β > 0 such that

τεuε(x)A1Uε(x)βε2s32.

Then we can choose ε1 > 0 such that τεuεR,for|x|<ε,0<ε<ε1. It follows from (4.16) that

F(x,τεuε(x))A0τε2μ,s1uε2μ,s1,

for |x|<ε,0<ε<ε1. Then for any 0<ε<ε1, by (4.8) we get

BεBεF(τεuε(x))F(τεuε(y))|xy|μdydxA02BεBε|τεuε(y)|2μ,s1|τεuε(x)|2μ,s1|xy|μdydxCA02BεBε|Uε(y)|2μ,s1|Uε(x)|2μ,s1|xy|μdydx=A02O(ε32s).

Note that F(u)0, (4.6) and (4.13), we have

(4.17) I2C|uε|22122μ,sBεBεF(τεuε(x))F(τεuε(y))|xy|μdydx(CC1A02)ε32s.

Inserting (4.15) and (4.17) into (4.14), we get

(4.18) I0(τεuε)2μ,s122μ,sSH,L2μ,s12μ,s+(C2+CC1A02)ε32s+O(ε6μ2).

Observe that 6μ2>32s for 3 < 4s, and A0 > 0 is arbitrary, we choose large enough A0 such that C + C2 C1A02<0 .Then for small ε > 0 we have v:=τεuε satisfies (4.12).

Theorem 4.1

Assume that (f1)−(f3) hold. Then autonomous problem (4.1) has a positive ground state solution u with I0(u)=cV0.

Proof

By Lemma 3.2 with V(x) = V0 and the Mountain Pass Theorem without (PS) condition (cf. [50]), there exists a (PS)cV0. -sequence {un}H0 of I0 with

cV0<2μ,s122μ,sSH,L2μ,s12μ,s.

By Lemma 3.5 and 4.3, {un} is bounded in Hs(ℝ3) and non-vanishing, namely there exist r,δ>0 and a sequence {yn}R3 such that

lim infnBr(yn)|un|2dxδ.

Up to s subsequence, there exists u ∈ Hs(ℝ3) such that

unuinHs(R3),unuinLlocr(R3),2r<2s,un(x)u(x)a.e.inR3.

As Lemma 2.5, we have I0(u)=0. Since I0andI0 are both invariant by translation, without lost of generality, we can assume that {yn} is bounded. Note that unuinLloc2(R3). Then u0.SouN0. Then

cV0I0(u)=I0(u)12I0(u),u=122μ,sR3R3G(u(y))|xy|μ(g(u(x))u(x)G(u(x)))dydx122μ,slim infnR3R3G(un(y))|xy|μ(g(un(x))un(x)G(un(x)))dydx=lim infn(I0(un)12I0(un),un)=cV0,

where we used Fatou Lemma and Lemma 2.4. Therefore, I0(u)=cV0, which means that u is a ground state solution for (4.1).

Next we prove that the solution u is positive, using u=max{u,0} as a test function in (4.1) we obtain

(4.19) R3(Δ)s2u(Δ)s2udx+R3V0|u|2dx=0.

On the other hand,

R3(Δ)s2u(Δ)s2udx=12C(s)R3×R3(u(x)u(y))(u(x)u(y))|xy|3+2sdxdy12C(s){u>0}×{u<0}(u(x)u(y))(u(y))|xy|3+2sdxdy+12C(s){u<0}×{u<0}(u(x)u(y))2|xy|3+2sdxdy+12C(s){u<0}×{u>0}(u(x)u(y))u(x)|xy|3+2sdxdy0.

Thus, it follows from (4.19) that u=0andu0. Rewriting the equation (4.1) in the form of

(Δ)su+V0u=(R3H(u(y))u(y)|xy|μdy)K(u)inR3,

where

H(u):=|u|2μ,s+F(u)u,K(u):=|u|2μ,s2μ+12μ,sf(u)L23μ(R3)+L63+2sμ(R3).

By Lemma 2.8, we know uLp(R3) for all p[2,18(3μ)(32s)). Using the growth assumption (f1) and the higher integrability of u, for some C > 0 we have

(4.20) |R3G(u(y))|xy|μ|C||u|2μ,s+|u|q1+|u|q2|33μC(|u|3(6μ)3μ2μ,s+|u|3q13μq1+|u|3213μq2),

which is finite since the various exponents live within the range [2,18(3μ)(32s)). Thus,

(Δ)su+V0uC(|u|2μ,s2μu+12μ,sf(u))inR3.

By the Moser iteration, similar arguments developed in Lemma 6.1 below, we can get uL(R3) and lim|x|+u(x)=0 uniformly in n. Then, by regularity theory [43], there exists α(0,1) such that uCloc0,α(R3). Therefore, if u(x0)=0 for some x0R3, we have that (Δ)su(x0)=0 and by [19, Lemma 3.2], we have

(Δ)su(x0)=C(s)2R3u(x0+y)+u(x0y)2u(x0)|y|3+2sdy,

therefore,

R3u(x0+y)+u(x0y)|y|3+2sdy=0,

yielding u0, a contradiction. Therefore, u is a positive solution of the equation (4.1) and the proof is completed.

The next result is a compactness result on autonomous problem which we will use later.

Lemma 4.8

Let {un}N0 be a sequence such that I0(un)cV0. Then {un} has a convergent subsequence in H0.

Proof

Since {un}N0, it follows from Lemma 4.1-(a3), Lemma 4.2-(b4) and the definition of cV0 that

vn=m1(un)=unun0S0+,nN,

and

Ψ0(vn)=I0(un)cV0=infS0+Ψ0.

Although S0+ is incomplete, due to Lemma 4.1, we can still apply the Ekeland’s variational principle [21] to the functional Θ0:HR{}, defined by Θ0(u)=Ψ0(u) if uS0+andΘ0(u)=ifuS0+, where H=S0+¯is the complete metric space equipped with the metric d(u,v):=uv0. In fact, take ε=1k2 in Theorem 1.1 of [21], we have a subsequence {vnk}{vn} such that

cV0Ψ(vnk)cV0+1k2.

From Theorem 1.1 in [21], for λ=1k, there exist a sequence {v~k}S0+ such that

Θ0(v~k)Θ0(vnk)<cV0+1k2

and

vnkv~k01k.

In particular, for any uS0+ we have

Ψ0(u)>Ψ0(v~k)1kuv~k0.

Hence, similar the proof for Theorem 3.1 in [21], we have that there exists λkR such that

Ψ^0(v~k)λkg0(v~k)01k,

where g0(u)=u021. Which means that

λk=1g0(v~k)02Ψ^0(v~k),g0(v~k)+ok(1),g0(v~k)=v~k.

From Lemma 4.2-(b1),

λk=Ψ^0(v~k),v~k+ok(1)=τv~kI0(tv~kv~k),v~k+ok(1)=ok(1).

Therefore, we can conclude there is a sequence {v~n}S0+ such that {v~n}isa(PS)cV0 sequence for Ψ0onS0+ and

unv~n0=on(1).

Now the remainder of the proof follows from Lemma 4.2, Theorem 4.1 and arguing as in the proof of Lemma 3.8.

5 Solutions for the penalized problem

In this section, we shall prove the existence and multiplicity of solutions. We begin showing the existence of the positive ground-state solution for the penalized problem (3.1).

Theorem 5.1

Suppose that the nonlinearity f satisfies (f1) − (f3) and that the potential function V satisfies assumptions (V1) − (V2). Then, for any ε > 0, problem (3.1) has a positive ground-state solution uε.

Proof

Similar to Lemma 3.2, we can prove that Jε also satisfies the Mountain Pass geometry. Let

cε:=infuHε{0}maxτ0Jε(τu)=infuNεJε(u).

Then, we know that there exists a (PS) sequence at cε, i.e.

Jε(un)0andJε(un)cε.

Therefore, by Lemma 3.7, the existence of ground state solution uε is guaranteed. Moreover, similarly to the proof in Theorem 4.1, we know that uε(x)>0 in ℝ3.

Next, we will relate the number of positive solutions of (3.1) to the topology of the set M. For this, we consider δ > 0 such that MδΩ and by Theorem 4.1, we can choose wN0 with I0(w)=cV0. Let η be a smooth nonincreasing cut-off function defined in [0,+) such that η(t)=1if0tδ2andη(t)=0iftδ. For each y ∈ 𝓜, let

Ψε,y(x)=η(|εxy|)w(εxyε).

Then for small ε > 0, one has Ψε,yHε{0}forallyM. In fact, using the change of variable z=xyε, one has

R3V(εx)Ψε,y2(x)dx=R3V(εx)η2(|εxy|)w2(εxyε)dx=R3V(εz+y)η2(|εz|)w2(z)dzCR3w2(z)dz<+.

Moreover, using the change of variable x=xyε,z=zyε we have

|(Δ)s2Ψε,y|22=12C(s)R3×R3|η(|εxy|)w(εxyε)η(|εzy|)w(εzyε)|2|xz|3+2sdxdz=12C(s)R3×R3|η(|εx|)w(x)η(|εz|)w(z)|2|xz|3+2sdxdz=|(Δ)s2η(|εx|)w(x)|22=|(Δ)s2ηεw|22,

where ηε(x)=η(|εx|). By Lemma 2.3, we see that ηεwDs,2(R3)asε0 and hence Ψε,yDs,2(R3) for ε > 0 small. Hence Ψε,yHε. Now we proof Ψε,y0. In fact,

R3Ψε,y2(x)dx=R3η2(|εxy|)w2(εxyε)dx=|εxy|<δη2(|εxy|)w2(εxyε)dx|z|δ2εη2(|εz|)w2(z)dzB0(δ2ε)w2(z)dzR3w2(z)dz>0

as ε → 0. Then Ψε,y0 for small ε > 0. Therefore, there exists unique τε>0 such that

maxτ0Iε(τΨε,y)=Iε(τεΨε,y)andτεΨε,yNε.

We introduce the map Φε:MNε by setting

Φε(y)=τεΨε,y.

By construction, Φε(y) has a compact support for any yMandΦε is a continuous map.

Lemma 5.1

The functional Φε(y) has the following property:

limε0Jε(Φε(y))=cV0uniformlyinyM.

Proof

Suppose that the result is false. Then, there exist some δ0>0,{yn}M and εn0 such that

(5.1) |Jεn(Φεn(yn))cV0|δ0.

By the definition of τεn we have

(5.2) 0<τεn2R3|(Δ)s2Ψεn,yn|2dx+τεn2R3V(εnx)Ψεn,yn2dx=12μ,sR3R3H(εnx,τεnΨεn,yn)h(εnx,τεnΨεn,yn)τεnΨεn,yn|xy|μdydx

It follows from (5.2) that τεn/0, then τεnτ0>0 for some τ0 > 0. If τεn+, by (f2) and the boundedness of Ψεn,yn, we get

(5.3) (R3|(Δ)s2Ψεn,yn|2dx+R3V(εnx)Ψεn,yn2dx)=12μ,sR3R3H(εnx,τεnΨεn,yn)h(εnx,τεnΨεn,yn)Ψεn,yn|xy|μτεndydx+

as n → +∞. But the left side of the above inequality is boundedness, which is impossible. Hence, 0<τ0 τεnC. Without loss of generality, we may assume that τεnT>0.

Next we claim that T = 1. By Lemma 2.3 and Lebesgue’s theorem we have

(5.4) limn+Ψεn,ynεn2=w02,limn+Σε(Ψεn,yn)=Σ0(w)

Moreover, from

(5.5) τεn2Ψεn,ynεn2=12μ,sR3R3H(εnx,τεnΨεn,yn)h(εnx,τεnΨεn,yn)τεnΨεn,yn|xy|μdydx

we can deduce that

(5.6) w02=limn12μ,sR3R3H(εnx,τεnΨεn,yn)h(εnx,τεnΨεn,yn)τεnΨεn,yn|xy|μdydx

Taking into account that w is a ground state solution to (4.1) and using (f2), we deduce that T = 1. It follows from (5.4), we have

(5.7) limn+Jεn(Φεn(yn))=J0(w)=cV0,

which is a contradiction with (5.1). This completes the proof.

Let ρ=ρ(δ)>0 be such that MδBρ(0). Consider χ:R3R3 be defined as χ(x)=xfor|x|ρ and χ(x)=ρx|x|for|x|ρ. Finally, let us consider the barycenter map βε:NεR3 given by

βε(u)=R3χ(εx)u2(x)dxR3u2(x)dxR3.

Lemma 5.2

The functional βε satisfies

limε0βε(Φε(y))=yuniformlyinyM.

Proof

Suppose by contradiction that there exist δ0>0,{yn}M and εn0 such that

(5.8) |βεn(Φεn(yn))yn|δ0.

Using the change of variables z=εnxynεn and the definition of βε, we have

βεn(Φεn(yn))=yn+R3(χ(εnz+y)yn)|η(|εnz|)w(z)|2dxR3|η(|εnz|)w(z)|2dx.

Since {yn}MBρ(0) andχ|Bρid, we conclude that

|βεn(Φεn(yn))yn|=on(1),

which contradicts (5.8) and the desired conclusion holds.

Lemma 5.3

Let εn0and{un}Nεn be such that Jεn(un)cV0. Then, there exists a sequence {y~n}R3 such that vn(x)=un(x+y~n) has a convergent subsequence in H0. Moreover, passing to a subsequence, yn:= εny~ny0M.

Proof

By Lemma 3.5, {un} is bounded in H0. Note that cV0>0, and since unεn0 would imply Jεn(un) 0, we can argue as in the proof of Lemma 4.3 to obtain a sequence {y~n}R3 and constants R, β > 0 such that

lim infn+BR(y~n)un2dxβ>0.

Define vn(x):=un(x+y~n), then {vn} is also bounded in H0 and up to a subsequence, we have

vnv0inH0.

Let τn>0 be such that v~n:=τnvnN0andsetyn=εny~n.By{un}Nεn, we have

cV0J0(v~n)=J0(τnun)Jεn(τnun)Jεn(un)=cV0+on(1).

Which implies that limn+J0(v~n)=cV0. In virtue of v~nN0, we obtain {v~n} is bounded in H0. It follows from the boundedness of {vn} in H0 that n} is bounded, without loss of generality, we may assume that τnτ00. If τ0 = 0, in view of the boundedness of {vn} in H0, we have v~n=τnvn0inH0. Hence J0(v~n)0, which contradicts cV0>0. Thus, τ0 > 0 and the weak limit of {v~n} is different from zero. Hence, up to a subsequence, we have v~nτ0v:=v~0 in H0 by the uniqueness of the weak limit. From Lemma 4.8, we know that v~nv~ in H0. Moreover, v~N0.

Now, we will show that {yn} is bounded in ℝ3. Suppose that after passing to a subsequence, |yn|+. Choosing R > 0 such that ΩBR(0). Without loss of generality we may assume that |yn| > 2R. Then, for all zBR/εn(o),

(5.9) |εnz+yn||yn||εnz|>R.

By the change of variable xz+tildeyn, using the fact that V0V(εx) and (5.9), we have

(5.10) vn02CR3H(εz+yn,vn)vndxCBR/εnG~(vn)vndz+CR3BR/εnG(vn)vndz.

Since G~(u)V0Kuandvnv in H0, we can see that (5.10) implies that

vn02=on(1),

that is vn0 in H0, which is a contradiction. Therefore, up to s subsequence, we may assume that yn y0R3. It remains to check that y0M. Clearly, if y0=Ω, then we can argue as before and we deduce that un0 in H0, which is impossible. Hence we only need to show that V(y0) = V0. Arguing by contradiction again, we suppose that V(y0) > V0. Then, by using v~nv~ in H0 and Fatou’s Lemma, we have

cV0=J0(v~)<lim infn(12R3|(Δ)s2v~|2dx+12R3V(y0)v~2dxΣ0(v~))lim infn(12R3|(Δ)s2v~n|2dx+12R3V(εnx+yn)v~n2dxΣ0(v~n))lim infnJεn(τnun)lim infnJεn(un)=cV0,

which yields a contradiction. So, y0M and the proof is completed.

Let h:R+R+ be any positive function satisfying h(ε)0+asε0+. Define the set

N~ε={uNε:Jε(u)cV0+h(ε)}.

Given yM, we conclude from Lemma 5.1 that h(ε)=supyM|Iε(Φε(y))cV0|0asε0+.Thus,Φε(y)N~ε and N~εforε>0. Moreover, we have the following Lemma.

Lemma 5.4

For any δ>0, there holds that

limε0supuN~εinfyMδ|βε(u)y|=0.

Proof

Let {εn}R+ be such that εn0. By definition, there exists {un}N~εn such that

infyMδ|βεn(un)y|=supuN~εninfyMδ|βεn(u)y|+on(1).

So, it suffices to find a sequence {yn}Mδ satisfying

(5.11) limn+|βεn(un)yn|=0.

Since unN~εnNεn, we get

cV0cεnJεn(un)cV0+h(εn).

It follows that Jεn(un)cV0. Thus, we can invoke Lemma 5.3 to obtain a sequence {y~n}R3 such that yn=εny~nMδ for n large enough. Then

βεn(un)=yn+R3(χ(εnx+yn)yn)|un(x+y~n)|2dxR3|un(x+y~n)|2dx.

For xR3 fixed, since εnx+ynyMδ, we have that the sequence {yn} satisfies (5.11). This completes the proof.

Next we prove our multiplicity result by presenting a relation between the topology of M the number of solutions of the modified problem (3.1), we will apply the Ljusternik-Schnirelmann abstract result in [44, 46].

Theorem 5.2

Assume that conditions (V1) − (V2) and (f1)− (f3) hold. Then, given δ > 0 there is ε^δ>0 such that for any ε(0,ε^δ), problem (3.1) has at least catMδ(M) positive solutions.

Proof

For yM set γε(y)=mε1(Φε(y)). It follows from Lemma 3.4 and Lemma 5.1 that

(5.12) limε0Ψε(γε(y))=limε0Iε(Φε(y))=cV0,

uniformly in yM. Let

S~ε+={wSε+:Ψε(w)cV0+h(ε)},

where h is given in the definition of N~ε. From (5.12), we know that there is a number ε^ such that S~ε+ for ε(0,ε^).

For a fixed δ>0, by Lemmas 3.3, 5.1-5.2 and 5.4, we know that there exists a ε^=ε^δ>0 such that for any ε(0,ε^δ), the diagram

MΦεN~εmε1S~ε+mεN~εβεMδ

is well defined. From Lemma 5.2, there is a function λ(ε, y) with |λ(ε,y)|<δ2 uniformly in yM, for all ε(0,ε^), such that βε(Φε(y)):=y+λ(ε,y)for allyM. Define H(t,y)=y+(1t)λ(ε,y). Then, H : [0,1]×MMδ is continuous. Obviously, H(0,y)=βε(Φε(y)),H(1,y)=yfor allyM. That is, H(t, y) is homotopy between βεΦε and the inclusion map id:MMδ. This fact and Lemma 4.3 in [7] implies that

catS~ε+γε(M)catMδ(M).

On the other hand, using the definition of N~ε and choosing ε^δ small if necessary, we see that Iε satisfies the (PS) condition in N~ε recalling Lemma 3.7. By Lemma 3.4 and 3.8, we obtain that Ψε satisfies the (PS) condition inS~ε+. Therefore, the standard Ljusternik-Schnirelmann theory provides at least catS~ε+γε(M) critical points of Ψε restricted to S~ε+. Using Lemma 3.7 again, we infer that Iε has at least catMδ(M) critical points. Using the same arguments contained in the proof Theorem 4.1, we see that the equation (3.1) has at least catMδ(M) positive solutions.

6 Proof of Theorem 1.1

In this section we will prove our main result. The idea is to show that the solutions obtained in Theorem 5.2 verify the following estimate uε(x)a,xΩεcforε small enough. This fact implies that these solutions are in fact solutions of the original problem (2.3). The key ingredient is the following result, whose proof uses an adaptation of the arguments found in [20], which are related to the Moser iteration method [35].

Lemma 6.1

Let εn0+andunN~εn be a solution of (3.1). Then up to a subsequence, vn=un(x+y~n) satisfies that vnL(R3) and there exists C > 0 such that

vnL(R3)C,nN,

where {y~n} is given in Lemma 5.3. Furthermore,

lim|x|vn(x)=0uniformlyinnN.

Proof

Rewriting the equation (3.1) in the form of

(Δ)su+V(εx)u=(R3H(u(y))u(y)|xy|μdy)K(u)inR3,

where

H(u):=|u|2μ,s+F(u)u,K(u):=|u|2μ,s2μ+12μ,sf(u)L23μ(R3)+L63+2sμ(R3).

By Lemma 2.8, we know uLp(R3)for allp[2,18(3μ)(32s)). Using the growth assumption (f1) and the higher integrability of u, for some C > 0 we have

|R3G(u(y))|xy|μ|C||u|2μ,s+|u|q1+|u|q2|33μC(|u|3(6μ)3μ2μ,s+|u|3q13μq1+|u|3213μq2),

which is finite since the various exponents live within the range [2,18(3μ)(32s)). Therefore, we have

(6.1) |h(εx,vn)|:=(R3H(εnx+εny~n,vn)|xy|μdy)h(εnx+εny~n,vn)V(εnx+εny~n)vnC(1+|vn|2s1)

for n large enough.

Let T > 0, we define

H(t)=0,ift0,tβ,if0<t<T,βTβ1(tT)+Tβ,iftT,

with β>1 to be determined later. Since H is Lipschitz with constant L0=βTβ1, we have

[H(vn)]Ds,2=(R3×R3|H(vn(x))H(vn(y))|2|xy|3+2sdxdy)12(R3×R3L02|vn(x)vn(y)|2|xy|3+2sdxdy)12=L0[vn]Ds,2.

Therefore, H(vn)Ds,2(R3). Moreover, by the definition of H, we know that H is a convex function, then we have

(6.2) (Δ)sH(vn)H(vn)(Δ)svn

in the weak sense. Thus, from H(vn)Ds,2(R3) and (6.1)-(6.2), we have

H(vn)2s2CR3|(Δ)s2H(vn)|2dx=CR3H(vn)(Δ)sH(vn)dxCR3H(vn)H(vn)(Δ)svndx=CR3H(vn)H(vn)h(εnx,vn)dxCR3H(vn)H(vn)dx+CR3H(vn)H(vn)vn2s1dx.

Using the fact that H(vn)H(vn)βvn2β1andvnH(vn)βH(vn), we have

(6.3) (R3(H(vn))2sdx)22s=Cβ(R3vn2β1dx+R3(H(vn))2vn2s2dx),

where C is a positive constant that does not depend on β. Notice that the last integral is well defined for T in the definition of H. Indeed

R3(H(vn))2vn2s2dx=vnT(H(vn))2vn2s2dx+vn>T(H(vn))2vn2s2dxT2β2R3vn2sdx+CR3vn2sdx<.

We choose now β in (6.3) such that 2β1=2s, and we name it β1, that is

(6.4) β1:=2s+12.

Let R^>0 to be fixed later. Attending to the last integral in (6.3) and applying the Holder’s inequality with exponents γ:=2s2andγ:=2s2s2,

(6.5) R3(H(vn))2vn2s2dx=vnR^(H(vn))2vn2s2dx+vn>R^(H(vn))2vn2s2dxvnR^(H(vn))2vnR^2s1dx+(R3(H(vn))2sdx)22s(vn>R^vn2sdx)2s22s.

By Lemma 4.8, we know that {vn} has a convergent subsequence in H0, therefore we can choose R^ large enough so that

(vn>R^vn2sdx)2s22s12Cβ1,

where C is the constant appearing in (6.3). Therefore, we can absorb the last term in (6.5) by the left hand side of (6.3) to get

(R3(H(vn))2sdx)22s2Cβ1(R3vn2sdx+R^2s1R3(H(vn))2vndx),

Now we use the fact that H(vn)vnβ1 and we take T, we obtain

(R3vn2sβ1dx)22s2Cβ1(R3vn2sdx+R^2s1R3vn2sdx).

and therefore

(6.6) vnL2sβ1(R3).

Let us suppose now β>β1. Thus, using that H(vn)vnβ in the right hand side of (6.3) and letting T we get

(6.7) (R3vn2sβdx)22sCβ(R3vn2β1dx+R^2s1R3vn2β+2s2dx).

Set c0:=2s(2s1)2(β1)andc1:=2β1c0. Notice that, since β>β1, then 0<c0<2s,c1>0. Hence, applying Young’s inequality with exponents γ:=2s/c0andγ:=2s/2sc0, we have

R3vn2β1dxc02sR3vn2sdx+2s2sc0R3vn2sc12sc0dxR3vn2sdx+R3vn2β+2s2dxC(1+R3vn2β+2s2dx),

with C > 0 independent of β. Plugging into (6.7),

(R3vn2sβdx)22sCβ(1+R3vn2β+2s2dx),

with C changing from line to line, but remaining independent of β. Therefore

(6.8) (1+R3vn2sβdx)12s(β1)(Cβ)12(β1)(1+R3vn2β+2s2dx)12(β1).

Repeating this argument we will define a sequence βm,m1 such that

2βm+1+2s2=2sβm.

Thus,

βm+11=(2s2)m(β11).

Replacing it in (6.8) one has

(1+R3vn2sβm+1dx)12s(βm+11)(Cβm+1)12(βm+11)(1+R3vn2sβmdx)12s(βm1).

Defining Cm+1:=Cβm+1 and

Am:=(1+R3vn2sβmdx)12s(βm1),

we conclude that there exists a constant C0 > 0 independent of m, such that

Amk=1mCk12(βk1)A1C0A1.

Thus,

(6.9) vnC0A1<,

uniformly in nN, thanks to (6.6). Now argue as in the proof of [3, Lemma 2.6], we conclude that

un(x)0as|x|,

uniformly in n ∈ ℕ. This finishes the proof of Lemma 6.1.

We are now ready to prove the main result of the paper.

Proof of Theorem 1.1. We fix a small δ > 0 such that MδΩ. We first claim that there exists some ε~δ>0 such that for any ε(0,ε~δ) and any solution uεN~ε of the problem (3.1), there holds

(6.10) uεL(R3Ωε)<a.

In order to prove the claim we argue by contradiction. So, suppose that for some sequence εn 0+ we can obtain unN~εn such that Iεn(un)=0 and

(6.11) unL(R3Ωε)a.

As in the proof of Lemma 5.3, we have that Jεn(un)cV0 and we can obtain a sequence {y~n}R3 such that εny~ny0M.

If we take r>0 such that Br(y0)B2r(y0)Ω we have that

Brεn(y0εn)=1εnBr(y0)Ωεn.

Moreover, for any zBrεn(y~n), there holds

|zy0εn|||zy~n|+|y~ny0εn|<1εn(r+on(1))<2rεn,

for n large. For these values of n we have that Brεn(y~n)Ωεn, that is, R3ΩεnR3Brεn(y~n). On the other hand, it follows from Lemma 6.1 that there is R>0 such that

un(x)<afor|x|RandnN,

from where it follows that

vn(xy~n)<aforxBRc(y~n)andnN.

Thus, there exists n0N such that for any nn0and rεn>R, there holds

R3ΩεnR3Brεn(y~n)R3BR(y~n).

Then, there holds

un(x)<axR3Ωεn,

which contradicts to (6.11) and the claim holds true.

Let ε^δ given by Theorem 5.2 and let εδ:=min{ε^δ,ε~δ}. We will prove the theorem for this choice of εδ. Let ε(0,εδ) be fixed. By using Theorem 5.2 we get catMδ(M) nontrivial solutions of problem (3.1). If uHε is one of these solutions, we have that uN~ε, and we can use (6.10) and the definition of g to conclude that H(,u)=G(u). Hence, u is also a solution of the problem (2.1). An easy calculation shows that ω(x)=u(xε) is a solution of the original problem (1.8). Then, (1.8) has at least catMδ(M) positive solutions.

Now we consider εn0+ and take a sequence unHεn of positive solutions of the problem (3.1) as above. In order to study the behavior of the maximum points of un, we first notice that, by the definition of H and (h1), (h2), there exists 0 < < a such that

(6.12) H(εnx,u)uV0Ku2,for allxR3,uγ.

Using a similar discussion above, we obtain R>0 and{y~n}R3 such that

(6.13) unL(BRc(y~n))<γ.

Up to a subsequence, we may assume that

(6.14) unL(BR(y~n))γ.

Indeed, if this is not the case, we have unL<γ, and therefore it follows from Jεn(un)=0 and (6.12) that

(6.15) unεn2V0KR3un2dx.

The above expression implies that unεn0asn, which leads to a contradiction. Thus, (6.14) holds.

By using (6.13) and (6.14) we conclude that the maximum points pn3 of un belongs to BR(y~n). Hence, pn=y~n+qn for some qn ∈ BR(0). Recalling that the associated solution of (1.8) is of the form ωn(x)=un(xε), we conclude that the maximum point ηε of vnisηε:=εny~n+εnqn. Since{qn}BR(0) is bounded and εny~ny0M, we obtain

limnV(ηε)=V(y0)=V0.

Thus, the proof of Theorem 1.1 is completed.

Acknowledgment

We would like to thank the anonymous referee for his/her careful readings of our manuscript and the useful comments made for its improvement.

This work is supported by NSFC (11771385), China.

References

[1] N. Ackermann. On a periodic Schrödinger equation with nonlocal superlinear part. Math. Z. 248(2):423–443, 2004.10.1007/s00209-004-0663-ySearch in Google Scholar

[2] C.O. Alves, F. Gao, M. Squassina, and M. Yang. Singularly perturbed critical Choquard equations. J. Differential Equations 263(7):3943–3988, 2017.10.1016/j.jde.2017.05.009Search in Google Scholar

[3] C.O. Alves and O.H. Miyagaki. Existence and concentration of solution for a class of fractional elliptic equation in RN via penalization method. Calc. Var. Partial Differential Equations 55(3):19, 2016.10.1007/s00526-016-0983-xSearch in Google Scholar

[4] C.O. Alves and M. Yang. Investigating the multiplicity and concentration behaviour of solutions for a quasi-linear Choquard equation via the penalization method. Proc. Roy. Soc. Edinburgh Sect. A 146(1):23–58, 2016.10.1017/S0308210515000311Search in Google Scholar

[5] A. Ambrosetti and A. Malchiodi. Concentration phenomena for nonlinear Schrödinger equations: recent results and new perspectives. In Perspectives in nonlinear partial differential equations volume 446 of Contemp. Math. pages 19–30. Amer. Math. Soc., Providence, RI, 2007.10.1090/conm/446/08624Search in Google Scholar

[6] P. Belchior, H. Bueno, O. H. Miyagaki, and G. A. Pereira. Remarks about a fractional Choquard equation: ground state, regularity and polynomial decay. Nonlinear Anal. 164:38–53, 2017.10.1016/j.na.2017.08.005Search in Google Scholar

[7] V. Benci and G. Cerami. Multiple positive solutions of some elliptic problems via the Morse theory and the domain topology. Calc. Var. Partial Differential Equations 2(1):29–48, 1994.10.1007/BF01234314Search in Google Scholar

[8] S. Bhattarai. On fractional Schrödinger systems of Choquard type. J. Differential Equations 263(6):3197–3229, 2017.10.1016/j.jde.2017.04.034Search in Google Scholar

[9] H. Brézis. Functional analysis, Sobolev spaces and partial differential equations Universitext. Springer, New York, 2011.10.1007/978-0-387-70914-7Search in Google Scholar

[10] H. Brézis and T. Kato. Remarks on the Schrödinger operator with singular complex potentials. J. Math. Pures Appl. (9) 58(2):137–151, 1979.Search in Google Scholar

[11] C. Bucur and E. Valdinoci. Nonlocal diffusion and applications volume 20 of Lecture Notes of the Unione Matematica Italiana Springer, 2016.10.1007/978-3-319-28739-3Search in Google Scholar

[12] D. Cassani and J. Zhang. Choquard-type equations with Hardy-Littlewood-Sobolev upper-critical growth. Adv. Nonlinear Anal. 8(1):1184–1212, 2019.10.1515/anona-2018-0019Search in Google Scholar

[13] S. Chen, Y. Li, and Z. Yang. Multiplicity and concentration of nontrivial nonnegative solutions for a fractional Choquard equation with critical exponent. Rev. R. Acad. Cienc. Exactas Fís. Nat. Ser. A Mat. RACSAM 114(1):Paper No. 33, 35, 2020.10.1007/s13398-019-00768-4Search in Google Scholar

[14] Y. Chen and C. Liu. Ground state solutions for non-autonomous fractional Choquard equations. Nonlinearity 29(6):1827–1842, 2016.10.1088/0951-7715/29/6/1827Search in Google Scholar

[15] S. Cingolani, M. Clapp, and S. Secchi. Multiple solutions to a magnetic nonlinear Choquard equation. Z. Angew. Math. Phys. 63(2):233–248, 2012.10.1007/s00033-011-0166-8Search in Google Scholar

[16] A. Cotsiolis and N. Tavoularis. Best constants for Sobolev inequalities for higher order fractional derivatives. J. Math. Anal. Appl. 295(1):225–236, 2004.10.1016/j.jmaa.2004.03.034Search in Google Scholar

[17] P. d’Avenia, G. Siciliano, and M. Squassina. On fractional Choquard equations. Math. Models Methods Appl. Sci. 25(8):1447–1476, 2015.10.1142/S0218202515500384Search in Google Scholar

[18] M. del Pino and P.L. Felmer. Local mountain passes for semilinear elliptic problems in unbounded domains. Calc. Var. Partial Differential Equations 4(2):121–137, 1996.10.1007/BF01189950Search in Google Scholar

[19] E. Di Nezza, G. Palatucci, and E. Valdinoci. Hitchhiker’s guide to the fractional Sobolev spaces. Bull. Sci. Math. 136(5):521–573, 2012.10.1016/j.bulsci.2011.12.004Search in Google Scholar

[20] S. Dipierro, M. Medina, and E. Valdinoci. Fractional elliptic problems with critical growth in the whole of Rn volume 15 of Appunti. Scuola Normale Superiore di Pisa (Nuova Serie) [Lecture Notes. Scuola Normale Superiore di Pisa (New Series)] Edizioni della Normale, Pisa, 2017.10.1007/978-88-7642-601-8Search in Google Scholar

[21] I. Ekeland. On the variational principle. J. Math. Anal. Appl. 47:324–353, 1974.10.1016/0022-247X(74)90025-0Search in Google Scholar

[22] F. Gao and M. Yang. The Brezis-Nirenberg type critical problem for the nonlinear Choquard equation. Sci. China Math. 61(7):1219–1242, 2018.10.1007/s11425-016-9067-5Search in Google Scholar

[23] Z. Gao, X. Tang, and S. Chen. On existence and concentration behavior of positive ground state solutions for a class of fractional Schrödinger–Choquard equations. Z. Angew. Math. Phys. 69(5):69:122, 2018.10.1007/s00033-018-1016-8Search in Google Scholar

[24] L. Guo and T. Hu. Existence and asymptotic behavior of the least energy solutions for fractional Choquard equations with potential well. Math. Methods Appl. Sci. 41(3):1145–1161, 2018.10.1002/mma.4653Search in Google Scholar

[25] N. Laskin. Fractional quantum mechanics and Lévy path integrals. Phys. Lett. A 268(4-6):298–305, 2000.10.1016/S0375-9601(00)00201-2Search in Google Scholar

[26] N. Laskin. Fractional Schrödinger equation. Phys. Rev. E (3) 66(5):056108, 2002.10.1142/9789813223806_0003Search in Google Scholar

[27] E.H. Lieb. Existence and uniqueness of the minimizing solution of Choquard’s nonlinear equation. Studies in Appl. Math. 57(2):93–105, 1976/77.10.1007/978-3-642-55925-9_37Search in Google Scholar

[28] E.H. Lieb and M. Loss. Analysis volume 14 of Graduate Studies in Mathematics American Mathematical Society, Providence, RI, second edition, 2001.10.1090/gsm/014Search in Google Scholar

[29] P.-L. Lions. The Choquard equation and related questions. Nonlinear Anal. 4(6):1063–1072, 1980.10.1016/0362-546X(80)90016-4Search in Google Scholar

[30] L. Ma and L. Zhao. Classification of positive solitary solutions of the nonlinear Choquard equation. Arch. Ration. Mech. Anal. 195(2):455–467, 2010.10.1007/s00205-008-0208-3Search in Google Scholar

[31] P. Ma and J. Zhang. Existence and multiplicity of solutions for fractional Choquard equations. Nonlinear Anal. 164:100–117, 2017.10.1016/j.na.2017.07.011Search in Google Scholar

[32] V. Moroz and J. Van Schaftingen. Groundstates of nonlinear Choquard equations: existence, qualitative properties and decay asymptotics. J. Funct. Anal. 265(2):153–184, 2013.10.1016/j.jfa.2013.04.007Search in Google Scholar

[33] V. Moroz and J. Van Schaftingen. Existence of groundstates for a class of nonlinear Choquard equations. Trans. Amer. Math. Soc. 367(9):6557–6579, 2015.10.1090/S0002-9947-2014-06289-2Search in Google Scholar

[34] V. Moroz and J. Van Schaftingen. Semi-classical states for the Choquard equation. Calc. Var. Partial Differential Equations 52(1-2):199–235, 2015.10.1007/s00526-014-0709-xSearch in Google Scholar

[35] J. Moser. A new proof of De Giorgi’s theorem concerning the regularity problem for elliptic differential equations. Comm. Pure Appl. Math. 13:457–468, 1960.10.1002/cpa.3160130308Search in Google Scholar

[36] T. Mukherjee and K. Sreenadh. Fractional Choquard equation with critical nonlinearities. NoDEA Nonlinear Differential Equations Appl. 24(6):Art. 63, 34, 2017.10.1007/s00030-017-0487-1Search in Google Scholar

[37] G. Palatucci and A. Pisante. Improved Sobolev embeddings, profile decomposition, and concentration-compactness for fractional Sobolev spaces. Calc. Var. Partial Differential Equations 50(3-4):799–829, 2014.10.1007/s00526-013-0656-ySearch in Google Scholar

[38] S. Pekar. Untersuchung über die Elektronentheorie der Kristalle Akademie Verlag, Berlin, 1954.10.1515/9783112649305Search in Google Scholar

[39] P.H. Rabinowitz. On a class of nonlinear Schrödinger equations. Z. Angew. Math. Phys. 43(2):270–291, 1992.10.1007/BF00946631Search in Google Scholar

[40] S. Secchi. Ground state solutions for nonlinear fractional Schrödinger equations in RN J. Math. Phys. 54(3):031501, 17, 2013.10.1063/1.4793990Search in Google Scholar

[41] R. Servadei and E. Valdinoci. The Brezis-Nirenberg result for the fractional Laplacian. Trans. Amer. Math. Soc. 367(1):67–102, 2015.10.1090/S0002-9947-2014-05884-4Search in Google Scholar

[42] Z. Shen, F. Gao, and M. Yang. Ground states for nonlinear fractional Choquard equations with general nonlinearities. Math. Methods Appl. Sci. 39(14):4082–4098, 2016.10.1002/mma.3849Search in Google Scholar

[43] L. Silvestre. Regularity of the obstacle problem for a fractional power of the Laplace operator. Comm. Pure Appl. Math. 60(1):67–112, 2007.10.1002/cpa.20153Search in Google Scholar

[44] M. Struwe. Variational methods Springer-Verlag, Berlin, 1990.10.1007/978-3-662-02624-3Search in Google Scholar

[45] A. Szulkin and T. Weth. Ground state solutions for some indefinite variational problems. J. Funct. Anal. 257(12):3802–3822, 2009.10.1016/j.jfa.2009.09.013Search in Google Scholar

[46] A. Szulkin and T. Weth. The method of Nehari manifold Int. Press, Somerville, MA, 2010.Search in Google Scholar

[47] F. Wang and M. Xiang. Multiplicity of solutions to a nonlocal Choquard equation involving fractional magnetic operators and critical exponent. Electron. J. Differential Equations pages Paper No. 306, 11, 2016.Search in Google Scholar

[48] Y. Wang and Y. Yang. Bifurcation results for the critical Choquard problem involving fractional p-Laplacian operator. Bound. Value Probl. pages No. 132, 11, 2018.10.1186/s13661-018-1050-7Search in Google Scholar

[49] J. Wei and M. Winter. Strongly interacting bumps for the Schrödinger-Newton equations. J. Math. Phys. 50(1):012905, 22, 2009.10.1063/1.3060169Search in Google Scholar

[50] M. Willem. Minimax theorems Birkhäuser Boston, Inc., Boston, MA, 1996.10.1007/978-1-4612-4146-1Search in Google Scholar

[51] H. Zhang, J. Wang, and F. Zhang. Semiclassical states for fractional Choquard equations with critical growth. Commun. Pure Appl. Anal. 18(1):519–538, 2019.10.3934/cpaa.2019026Search in Google Scholar

[52] W. Zhang and X. Wu. Nodal solutions for a fractional Choquard equation. J. Math. Anal. Appl. 464(2):1167–1183, 2018.10.1016/j.jmaa.2018.04.048Search in Google Scholar

Received: 2019-07-10
Accepted: 2020-09-06
Published Online: 2020-12-01

© 2021 Zhipeng Yang and Fukun Zhao, published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 19.4.2024 from https://www.degruyter.com/document/doi/10.1515/anona-2020-0151/html
Scroll to top button